It is often acknowledged that the search for life on Mars might produce false positive results, particularly via the detection of objects, patterns or substances that resemble the products of life in some way but are not biogenic. The success of major current and forthcoming rover missions now calls for significant efforts to mitigate this risk. Here, we review known processes that could have generated false biosignatures on early Mars. These examples are known largely from serendipitous discoveries rather than systematic research and remain poorly understood; they probably represent only a small subset of relevant phenomena. These phenomena tend to be driven by kinetic processes far from thermodynamic equilibrium, often in the presence of liquid water and organic matter, conditions similar to those that can actually give rise to, and support, life. We propose that strategies for assessing candidate biosignatures on Mars could be improved by new knowledge on the physics and chemistry of abiotic self-organization in geological systems. We conclude by calling for new interdisciplinary research to determine how false biosignatures may arise, focusing on geological materials, conditions and spatiotemporal scales relevant to the detection of life on Mars, as well as the early Earth and other planetary bodies.

Thematic collection: This article is part of the Astrobiology: Perspectives from the Geology of Earth and the Solar System collection available at: https://www.lyellcollection.org/cc/astrobiology-perspectives-from-geology-of-earth-and-solar-system

These are exciting times for astrobiology. The search for evidence of past life on Mars is the central scientific objective of two contemporary rover missions: NASA's Mars 2020 Mission (Perseverance rover, touchdown February 2021) and the ESA-RosCosmos ExoMars Mission (Rosalind Franklin rover, scheduled for launch in 2022). Both rovers will seek geological biosignatures in clay-rich, subaqueously deposited sedimentary rocks of Noachian age (c. 4 Ga), probably representing habitable palaeoenvironments (Horgan et al. 2020; Quantin-Nataf et al. 2021). Both rovers are equipped with sophisticated multispectral cameras and instruments for the analysis of these rocks and any organic matter they may contain.

The two missions differ slightly in their technical abilities. For example, the Rosalind Franklin rover carries a drill capable of extracting samples from 2 m below the surface, where the overlying rock may have protected organic molecules from oxidation or radiation damage. The Perseverance rover can detect organic matter, but not extract it from such depths nor characterize it in as much detail. However, Perseverance is equipped with a sample-caching system that will enable it to extract astrobiologically promising geological samples and seal them for subsequent collection and return to Earth, probably in a joint ESA–NASA operation (Muirhead et al. 2020). These will be the first samples brought back from Mars and will serve as a rich source of data for the investigation of martian geology and astrobiology. In theory, it will become possible to analyse rocks from Mars using almost the full range of techniques used to detect ancient biosignatures on Earth.

Studies of the Earth's biosphere and geological record have illuminated an enormous variety of biosignatures that astrobiologists might detect either in situ on Mars or in returned samples. For instance, samples may contain: morphological fossils as small as individual microbial cells, colonies or biofilms, or as large as stromatolites; organic materials with a compositional profile, complexity, modular molecular construction and/or stereoisomerism that suggest biosynthetic origins; or isotopic, geochemical or mineralogical anomalies that would normally be interpreted as evidence of life (Westall et al. 2015; Hays et al. 2017; Vago et al. 2017, 2019; McMahon et al. 2018).

Astrobiologists have amassed innumerable examples of these and other kinds of microbial biosignatures from ancient and modern Mars-analogue environments, providing an immensely rich set of reference data or ‘search images’. We might reasonably argue, therefore, that the risk of false negative results in the search for life is low, or at least lower for the forthcoming rover missions than it was for the Viking landers of the 1970s: if evidence of life is really present in the materials that either the Rosalind Franklin or Perseverance rover brings to light, there is now a good chance that we will find and recognize it. What concerns us here is the risk of false positive errors in the detection of life, specifically those arising from the misinterpretation of abiotic geological and chemical features that are misleadingly life-like (rather than analytical false positives or spacecraft contamination).

One reason for concern is that such errors have been frequent in the history of palaeobiology and astrobiology down to the present day. In the nineteenth century, intricate layered and tubular structures found in serpentine-rich metacarbonate rocks from the Canadian shield were interpreted as the fossilized tests of ancient forams, designated Eozoon canadense, the ‘dawn animal of Canada’; it took much heated debate, and some decades of careful work, to show that these were abiotic textures (Adelman 2007). Palaeontologists in the twentieth century repeatedly made similar misdiagnoses; examples listed by Cloud (1973) include spurious ‘worm tracks’ (drag marks), ‘medusae’ (pyrite rosettes), ‘arthropods’ (mud curls around desiccation cracks) and numerous Proterozoic ‘microfossils’ (crystallites, contaminating spores and preparation artefacts).

The difficulty of correctly determining the biogenicity of life-like geological substances and structures has continued to plague palaeontology, and the new field of astrobiology, often despite laudable efforts to practice the ‘care, patience, and critical attitude’ recommended by Cloud. For example, McKay et al. (1996) reported multiple lines of evidence in the martian meteorite ALH84001 that were suggestive of past life on Mars, making news bulletins around the world. These included carbonate globules and magnetite crystals resembling bacteriogenic precipitates, polyaromatic hydrocarbon compounds and worm-like microstructures interpreted as morphological fossils. The origins of these features are still unclear today, but abiotic explanations have been offered for all of them, and the overall case for life in ALH84001 no longer seems compelling (Martel et al. 2012). This example warns us that it is not enough to have ‘multiple lines of evidence’ for biogenicity (an oft-repeated mantra) if each line of evidence is ambiguous.

Other ongoing debates concern the oldest fossil and geochemical evidence of life on Earth. Schopf (1993) presented fossil ‘cyanobacteria’ from the 3.5-Gyr-old Apex Chert of Western Australia, which were later reinterpreted as abiotic carbon organized around spherulitic quartz growth and/or worm-like delaminated clay booklets (Brasier et al. 2002, 2005; Wacey et al. 2016a). Recent work highlights the possibility that organic matter in such cherts might be biogenic even if it has been abiotically redistributed (Duda et al. 2018), while Schopf et al. (2018) have claimed that morphology-specific carbon isotope values confirm the biogenicity of the Apex microstructures and reveal them to include methane-cycling archaea; this controversy continues (e.g. Alleon and Summons 2019). Nutman et al. (2016) described triangular features exposed on 3.7-Gyr-old metasedimentary rocks from the Isua supracrustal belt in Greenland and interpreted them as fossil stromatolites; these have been reinterpreted by some researchers as abiotic deformation features on the basis of 3D imaging and geochemical analysis (Allwood et al. 2018; Zawaski et al. 2020), although Nutman and co-workers stand by their original view (Nutman et al. 2021). Dodd et al. (2017) described hematite tubules in a c. 4-Gyr-old hydrothermal chert from the Nuvvuagittuq Greenstone Belt in Canada as the mineralized sheaths of ancient Fe-oxidizing bacteria; these have been queried as possible chemical gardens (McMahon 2019), but further analyses are needed.

These examples show that the return of samples from Mars will not necessarily solve once and for all the problem of the existence of (ancient) life on that planet. Although we must hope for definitive results, candidate biosignatures are likely to be at least somewhat ambiguous (García-Ruiz et al. 2002) and may spark debates that cannot be quickly resolved. To give an extreme example from Earth, the biogenicity of the filamentous mineral networks found in ‘moss agates’ has been unclear for more than 200 years (e.g. Daubenton 1782; MacCulloch 1814; Bowerbank 1842; Göppert 1848; Liesegang 1914; Brown 1957; Hopkinson et al. 1998; Hofmann and Farmer 2000; McMahon 2019; Götze et al. 2020). As in all the debates cited here, the key evidence will come from the investigation of abiotic physicochemical systems and their capacity to mimic the forms and properties of life. Yet this area of enquiry has received rather scant and unsystematic attention from astrobiologists, who have tended to focus their published work on expanding our knowledge of life's signatures rather than its abiotic mimics. This problem was lamented by Cady et al. (2003), who predicted that:

… images and spectroscopic data indicative of life will continue to accumulate in the peer-reviewed literature. At the same time, millions of images and spectra from dubiofossils (of unknown origin) and pseudofossils (abiotic mimics) will also continue to accumulate, yet they will rarely appear in publication.

With this motivation, we here attempt to summarize the physicochemical processes so far known to generate life-like morphologies, minerals, molecules and other phenomena, and consider how and where they may have taken place on early Mars, with a particular focus on the materials that the Rosalind Franklin and Perseverance rovers may encounter and sample at Oxia Planum and Jezero Crater, respectively. We note that, by definition, a biosignature is more than simply a phenomenon produced by life: it is a phenomenon that specifically requires a biological agent – that is, it could not have been produced naturally by non-living systems (e.g. Des Marais et al. 2008). The reliability of any detected biosignatures on Mars therefore depends crucially on our understanding of the abiotic processes that might mimic them (Brasier and Wacey 2012; Chan et al. 2019).

This paper is not intended to criticize or undermine the search for geological evidence of life on Mars (or the early Earth), but to facilitate the formulation and testing of abiotic ‘null hypotheses’ for the evaluation of candidate biosignatures if and when they are discovered. It is intended to complement the many existing reviews on the possible biosignatures that may be found on Mars (e.g. Westall et al. 2015; Hays et al. 2017; Vago et al. 2017; McMahon et al. 2018). For the most part, however, we do not attempt to provide ‘biogenicity criteria’ for evaluating the credibility of candidate biosignatures. Rather, we emphasize that such criteria will become more reliable as the formation of false biosignatures becomes better understood. Significant progress in this field will require the numerical, experimental and analytical skills and specialist knowledge of chemists, physicists, geologists, mineralogists and materials scientists, among others.

The search for ancient organic matter is a central element of the life-detection strategy for ExoMars, Mars 2020 and the anticipated joint ESA–NASA sample return mission. Indigenous martian organic matter with a chemical fingerprint sufficiently similar to (degraded) biomass would be an exciting potential biosignature in its own right and could significantly strengthen the credibility of any associated morphological, isotopic or mineralogical potential biosignatures (Summons et al. 2011). Despite the pervasively oxidizing and ionizing conditions at the martian surface, the Sample Analysis at Mars (SAM) instrument aboard the Curiosity rover has confirmed that ancient organic matter has survived where shielded and protected by minerals, particularly clays. This instrument measured trace amounts of aliphatic and aromatic molecular fragments in drill powder extracted from a Noachian–Hesperian mudstone, which had been exposed relatively recently by scarp retreat (Glavin et al. 2013; Farley et al. 2014; Ming et al. 2014; Freissinet et al. 2015; Eigenbrode et al. 2018). This organic matter has been altered diagenetically (e.g. sulfurized) and may have originally derived partly from carbonaceous meteorites and partly from indigenous sources on Mars (Eigenbrode et al. 2018; Franz et al. 2020). Interaction with perchlorates during sample pyrolysis chlorinated some hydrocarbons, further obscuring their original composition (Szopa et al. 2020). Ancient, probably indigenous, organic matter is also found in some basaltic martian meteorites, probably from abiotic sources (e.g. Steele et al. 2012).

Jezero Crater and Oxia Planum both host clay-rich Noachian sedimentary rocks with good potential to contain preserved organic matter, which both the Perseverance and Rosalind Franklin rovers would be capable of detecting and, to some extent, characterizing. Rosalind Franklin's Mars Organic Molecular Analyzer (MOMA) instrument can fingerprint organic molecules in drill powders with very high sensitivity and specificity, even allowing enantiomeric excesses to be measured in some compounds (Goesmann et al. 2017) (Table 1). Unfortunately, determining the biogenicity of any organic matter detected may not be straightforward. Noachian Mars, like the early Earth, was presumably supplied with a rich complement of organic matter derived from meteorites and from a variety of endogenous abiotic processes (Chyba and Sagan 1992). Among the latter, the most frequently discussed involve either the reduction of inorganic carbon catalysed on mineral surfaces under hydrothermal conditions (e.g. Fischer–Tropsch type synthesis, organosulfur pathways and electrochemical pathways), high-temperature carbonate decomposition or (mineral-catalysed) photochemical carbon fixation and cycling (e.g. McCollom 2013; Milesi et al. 2015; Dalai et al. 2016; Steele et al. 2018; Franz et al. 2020). Taken together, these various sources would have provided, inter alia, polyaromatic hydrocarbons, alkanes, amines, fatty acids/lipids, carboxylic acids, amino acids, nucleobases, aldehydes, ketones and carbohydrates, all of which are present – along with marked enantiomeric excesses – in carbonaceous chondrites (Botta and Bada 2002; Simoneit 2004; Myrgorodska et al. 2015; Dalai et al. 2016; Furukawa et al. 2019; Lai et al. 2019).

Table 1.

Descriptions of analytical and imaging instruments aboard the Perseverance and Rosalind Franklin rovers, with their respective spatial resolutions

Instrument nameDescriptionRelevant spatial resolution
Perseverance rover
 Mastcam-ZLong-range (2–100 m) multispectral, stereoscopic imager; wavelength selection using narrowband filters provides some mineralogical information150 μm per pixel to 7.4 mm per pixel depending on distance
 WATSONCamera for fine-scale imaging of textures and structures on the martian surface15.9 μm per pixel
 SHERLOCDeep UV Raman and fluorescence spectrometer for the detection and mapping of organic molecules and minerals; combined with a high-resolution camera (context imager)c. 50 μm per pixel (spectroscopy)
10.1 μm per pixel (context imager)
 SuperCamLaser-induced breakdown spectroscopy, Raman and time-resolved luminescence spectroscopy, and visible and infrared reflectance spectroscopy for remote (up to c. 12 m) mineralogy and elemental and molecular chemical characterization (including organic molecules); combined with a remote micro-imager0.5–5 mm per spot depending on technique and distance (spectroscopy)
c. 30 μm per pixel at 1.5 m (imager)
 PIXLX-ray fluorescence spectrometer to map the distribution and abundance variations of chemical elements at the sub-millimetre scale58 μm per pixel
Rosalind Franklin rover
 Panoramic camera (PanCAM)Two wide-angle stereo cameras, one high-resolution camerac. 170 μm per pixel at 2 m distance
 CLose UP Imager (CLUPI)Close-range (50 cm) ‘hand lens' imager for imaging rocks and drill powders7.8 μm per pixel
 Infrared Spectrometer for ExoMars (ISEM)Infrared spectrometer for remote mineralogical characterization and identification of rock targets3–10 cm per spot depending on distance
 Mars Multispectral Imager for Subsurface Studies (Ma_MISS)Visible and near-infrared spectroscopy to characterize the mineralogy of the walls of drilled boreholes120 μm spatial resolution
 MicrOmegaNear-infrared and visible light spectromicroscopy for spatially resolved characterization of minerals and organic molecules in drill powders20 μm per pixel
 Raman laser spectrometer (RLS)Raman spectrometry to identify mineralogy and organic matter in drill powdersSpot size c. 50 μm
 Mars Organic Molecule Analyzer (MOMA)UV laser desorption and pyrolysis (with or without derivatization agent), gas chromatography–mass spectrometry to identify and characterize organic molecules (alkanes, lipids, carboxylic acids, fatty acids, amino acids, nucleobases, amines, alcohols; chirality)N/A
Instrument nameDescriptionRelevant spatial resolution
Perseverance rover
 Mastcam-ZLong-range (2–100 m) multispectral, stereoscopic imager; wavelength selection using narrowband filters provides some mineralogical information150 μm per pixel to 7.4 mm per pixel depending on distance
 WATSONCamera for fine-scale imaging of textures and structures on the martian surface15.9 μm per pixel
 SHERLOCDeep UV Raman and fluorescence spectrometer for the detection and mapping of organic molecules and minerals; combined with a high-resolution camera (context imager)c. 50 μm per pixel (spectroscopy)
10.1 μm per pixel (context imager)
 SuperCamLaser-induced breakdown spectroscopy, Raman and time-resolved luminescence spectroscopy, and visible and infrared reflectance spectroscopy for remote (up to c. 12 m) mineralogy and elemental and molecular chemical characterization (including organic molecules); combined with a remote micro-imager0.5–5 mm per spot depending on technique and distance (spectroscopy)
c. 30 μm per pixel at 1.5 m (imager)
 PIXLX-ray fluorescence spectrometer to map the distribution and abundance variations of chemical elements at the sub-millimetre scale58 μm per pixel
Rosalind Franklin rover
 Panoramic camera (PanCAM)Two wide-angle stereo cameras, one high-resolution camerac. 170 μm per pixel at 2 m distance
 CLose UP Imager (CLUPI)Close-range (50 cm) ‘hand lens' imager for imaging rocks and drill powders7.8 μm per pixel
 Infrared Spectrometer for ExoMars (ISEM)Infrared spectrometer for remote mineralogical characterization and identification of rock targets3–10 cm per spot depending on distance
 Mars Multispectral Imager for Subsurface Studies (Ma_MISS)Visible and near-infrared spectroscopy to characterize the mineralogy of the walls of drilled boreholes120 μm spatial resolution
 MicrOmegaNear-infrared and visible light spectromicroscopy for spatially resolved characterization of minerals and organic molecules in drill powders20 μm per pixel
 Raman laser spectrometer (RLS)Raman spectrometry to identify mineralogy and organic matter in drill powdersSpot size c. 50 μm
 Mars Organic Molecule Analyzer (MOMA)UV laser desorption and pyrolysis (with or without derivatization agent), gas chromatography–mass spectrometry to identify and characterize organic molecules (alkanes, lipids, carboxylic acids, fatty acids, amino acids, nucleobases, amines, alcohols; chirality)N/A

On Earth, hydrothermal cherts as old as 3.5 Ga contain abundant kerogen-like carbonaceous matter suggested (although not conclusively demonstrated) to derive from Fischer–Tropsch type synthesis (Lindsay et al. 2005; Alleon et al. 2019) and it has been argued that ‘the abiotic organic output from [such] hydrothermal systems may overwhelm any early biospheric geochemical signal’ (Lindsay et al. 2005). On Mars, the structurally simple products of abiotic carbon fixation may be easier to interpret given the lack of metamorphic overprint compared with the Earth (assuming minimal degradation by diagenesis, oxidation and ionizing radiation).

Difficulty arises, however, from the fact that the relatively simple building blocks produced by abiotic carbon fixation can give rise to much more complex products, especially where thermodynamic disequilibrium is available to drive chemical self-organization; this much is evident from the existence of life on Earth. In other words, abiotic organic matter on Mars may have undergone transformations that are on the road to life (e.g. enantiomeric selection, metabolism-like reaction networks and autocatalytic cycles, the formation of complex polymers with repeating subunits, unexpected molecular weight distributions), but without reaching this destination. We lack a good understanding of these processes or of the diversity of molecular (or multimolecular) pseudofossils they might produce, although computational studies of prebiotic systems chemistry potentially offer some important clues (e.g. Wołos et al. 2020). If current models for the origin of life on Earth are any guide, then homochirality, polypeptides/proteins, RNA and DNA may all have emerged in prebiotic chemical reaction systems prior to the onset of Darwinian life (e.g. Orgel 2004; Brack 2007; Dalai et al. 2016; Erastova et al. 2017; Teichert et al. 2019; Xu et al. 2020), so even these are not necessarily unambiguous biosignatures by themselves.

None of this is to deny that anomalies or patterns in the overall molecular profile of martian organic matter could still be strongly suggestive of biosynthesis (despite pitfalls; even meteoritic carbon can apparently show a preference for odd-over-even carbon numbers; Westall et al. 2018). Some molecules – which, once life is established, it produces continuously – are too complex to assemble in large numbers without biologically evolved machinery (see Marshall et al. 2017 for a discussion of where the abiotic/biotic ‘threshold’ might be). If we are lucky, it may be possible to recognize such anomalies even if martian life is built from a significantly different core chemistry from life on Earth (Goesmann et al. 2017; Marshall et al. 2017). But timing is everything: sedimentary rocks at Jezero Crater and Oxia Planum were probably deposited within only a few hundred million years of the origin of life on Earth; even if they post-date an origin of life on Mars, it may be impossible to detect any biotic signal amid the complex products of a still largely ‘prebiotic’ carbon cycle (Lindsay et al. 2005).

Assessment of the biogenicity of organic matter (or any other candidate biosignature) will require close study of the geological context by the Rosalind Franklin or Perseverance rovers. An association between complex organic matter and evidence for local serpentinization, hydrothermal conditions favourable for Fischer–Tropsch type synthesis, or perhaps ferrocyanide salts (Sasselov et al. 2020), could be especially suggestive of abiotic/prebiotic chemistry rather than biotic chemistry. From this point of view, the lack of clear mineralogical evidence (so far) for local hydrothermal activity at the Oxia Planum landing site is reassuring. However, the large fluvial catchment area in this region may have concentrated organic matter of remote and diverse provenance and carbonate minerals in Jezero Crater could be related to serpentinization (Quantin-Nataf et al. 2021; Zastrow and Glotch 2021).

Conversely, any association between organic matter and other potential biosignatures (e.g. morphologies) could strengthen the case that life has been found. However, caution is needed again here. We lack a good understanding of how abiotic organic matter would have interacted with minerals and fluids on Noachian Mars and several forms of inorganic ‘pseudobiosignature’ might have adsorbed or passively entrained abiotic organic matter in some settings. In addition, organic matter itself can self-organize into life-like morphologies under some conditions (see following section on organic biomorphs).

Microbialites are microbially influenced organo-sedimentary structures that result from the trapping and binding of sediment and the authigenic precipitation of calcium carbonate and other minerals in benthic microbial communities (Burne and Moore 1987). The interplay between these biotically driven processes and extrinsic hydrodynamic, sedimentary and physicochemical conditions generates a wide variety of microbialite morphologies and textures, many of which have a fossil record on Earth reaching back to the Archean (Noffke et al. 2006; Baumgartner et al. 2019). The most abundant fossil microbialites are stromatolites and microbially induced sedimentary structures (MISS). Stromatolites are laminated growth structures that project upwards as cones, domes, ridges and columns, whereas MISS are less vertically extended and can be restricted to single bedding planes (Noffke et al. 2001). MISS include fossil microbial mats, under-mat textures and wrinkle structures formed in microbially biostabilized clastic sediments; these macroscopic forms are commonly accompanied by microtextural features such as oriented grains and crinkly organic-rich laminae (Noffke et al. 2001).

Microbialites have excellent preservation potential in carbonate, chert and siliciclastic facies and might easily be large enough to image at good spatial resolution using the Mars rover cameras and close-up imagers. Structures that resemble microbialites have been highlighted as attractive targets for sampling and return (Westall et al. 2015; Hays et al. 2017; Vago et al. 2017; McMahon et al. 2018). However, it is generally accepted that gross morphology alone is an unreliable indicator of the biogenicity of sedimentary structures and that (micro)textural and geochemical data will be necessary to confirm any tentative identification of microbialites on Mars (Noffke 2018). Relevant observations would include filamentous microfossils, crinkly carbon-rich laminae, fenestrae (fossilized gas bubbles, which can contain mineral grains indicative of local photosynthetic oxidation), oriented grains and grains ‘floating’ in an organic matrix (Noffke et al. 2001; Noffke 2018; Wilmeth et al. 2019). All of these features might (just) be large and distinct enough to resolve in situ using rover cameras (Table 1) if they are found in fresh drillcore (Noffke et al. 2001; Noffke 2018). Abiotic processes are also unlikely to mimic the conical form and regular spacing of some ancient stromatolites, features that are now understood to result from phototactic growth and competition for nutrients and light (e.g. Batchelor et al. 2004; Petroff et al. 2010). Microbialites that lack any of these features (or in which they cannot be detected by rover instruments) may be difficult to distinguish from pseudofossils. Indeed, there are several indications that pseudomicrobialites could have formed both on Earth and on early Mars.

The biogenicity of some stromatolite-like structures in very ancient (meta)sediments on Earth has often been contested (e.g. Grotzinger and Rothman 1996; Allwood et al. 2018; Zawaski et al. 2020). The presence of (pseudo)laminations can result from abiotic factors, including a variable energy/sediment supply during deposition and self-organization during crystallization, diagenesis, metamorphism and chemical weathering (Ortoleva 1994; Brasier et al. 2017, 2019; Allwood et al. 2018; Zawaski et al. 2020) (Fig. 1a). Domes and ridges can form via soft sediment or tectonic deformation (Antcliffe and McLoughlin 2008; Allwood et al. 2018). Grotzinger and Rothman (1996) showed that the laminations seen in some purported Precambrian stromatolites conform to the Kardar–Parisi–Zhang equation, which describes the growth of self-affine interfaces in terms of particle supply, surface-normal growth, random accretion and relaxation/diffusion. Because the Kardar–Parisi–Zhang equation captures the growth of a variety of interfaces, both biotic and abiotic, this result implies that organisms were not necessarily involved in shaping these laminations.

Fig. 1.

Pseudomicrobialites. (a) Laminated red siltstones draping the surface of gneiss boulders (Brasier et al. 2017). (b, c) Stromatolite-like laminated domes and branching columns formed by aerosolized colloidal paint sprayed onto flat surfaces (McLoughlin et al. 2008). Images reproduced with permissions.

Fig. 1.

Pseudomicrobialites. (a) Laminated red siltstones draping the surface of gneiss boulders (Brasier et al. 2017). (b, c) Stromatolite-like laminated domes and branching columns formed by aerosolized colloidal paint sprayed onto flat surfaces (McLoughlin et al. 2008). Images reproduced with permissions.

McLoughlin et al. (2008) showed experimentally that aerosolized colloidal paint sprayed onto a flat surface generates laminated deposits with remarkably stromatolite-like domes and columns, albeit on a small scale (column width <1 cm) (Fig. 1b, c). Many similar examples have been collected from factory floors where spray paint is applied to cars. The ballistic deposition of colloids (which can also be modelled using the Kardar–Parisi–Zhang equation) thus seems to be a plausible abiotic mechanism for the origin of stromatolite-like morphologies in the calcareous and siliceous deposits found in hot spring ‘splash zones’ on Earth. Such deposits commonly contain fossil bacteria, but this does not necessarily mean that bacteria were instrumental in producing the macroscopic stromatolite-like growth pattern (McLoughlin et al. 2008). Small-scale, dendritic stromatolite- (and thrombolite-) like forms can also result from diffusion-limited aggregation at high surface tension (Duarte-Neto et al. 2014).

Less is known about how pseudo-MISS might form, but Davies et al. (2016) emphasized that particle-sticking, loading, fluid escape, impression, shear and shrinkage can generate complex, ‘MISS-like’ sedimentary surface textures in sufficiently cohesive and plastic sediments. They provide examples from modern and ancient settings where abiotic factors have imbued sediments with high viscosity and cohesivity, mimicking some of the patterns and textures associated with biostabilization (Davies et al. 2016).

We speculate that two general attributes of depositional environments on early Mars may have been especially conducive to the formation of pseudomicrobialites. First, fine-grained clay-rich sediments that were presumably not bioturbated by animals would have been naturally cohesive and thus able to form and retain a variety of sedimentary structures that might be mistaken for fossil MISS (Davies et al. 2016). Direct evidence for a high degree of sediment cohesion on early Mars has been discovered at Gale Crater in the form of fossil subaqueous shrinkage cracks, a sedimentary structure sometimes associated with microbial biostabilization on Earth (Siebach et al. 2014; McMahon et al. 2017).

Second, aqueous fluids interacting with early martian sediments may commonly have been rich in silica, salts and other dissolved minerals liable to precipitate as cements and crusts near the sediment–water(–atmosphere) interface, further enhancing sediment cohesion and potentially generating intrastratal cracks, polygons, tepee structures and related macroscopic effects that may superficially mimic MISS (McLennan et al. 2005). Pseudomicrobialites might also have adsorbed and absorbed a certain amount of abiotic organic matter from the environment as carbon-rich laminae (e.g. from ‘primordial oil slicks’; Nilson 2002).

The effects of post-depositional processes may also confound the detection of true microbialites on Mars. Many martian outcrops are sculpted by wind abrasion and are unlikely to display well-preserved textures on bedding planes unless they have been very recently exposed. Irregular textures and forms that seem to resemble microbialites can result merely from the vagaries of erosion. Thus the MISS-like features identified by Noffke (2015) in Curiosity rover imagery from Gale Crater appear not to be primary sedimentary structures (Davies et al. 2016; McMahon et al. 2018).

Sinter-like hot spring silica deposits are unlikely to be encountered by the Perseverance or Rosalind Franklin rovers, but were previously observed by the Spirit rover in Gusev Crater (Ruff and Farmer 2016). This opaline deposit consists of porous, nodular material with millimetre- to centimetre-scale finger-like extensions, resembling fossil-microbe-bearing silica structures from hot springs on Earth described as ‘microbially mediated microstromatolites’ (Ruff and Farmer 2016; Ruff et al. 2020). A potential abiotic explanation for this morphology is that the silica precipitated evaporatively from sprayed or splashed droplets or accumulated from the ballistic deposition of suspended particles (McLoughlin et al. 2008). Experimental work to refine and test this hypothesis would be worthwhile.

Mars rovers are not designed or equipped to detect microfossils of bacterial cell size in situ (Table 1). On Earth, some Archean cellular fossils are >100 μm across, but these are carbonaceous compressions and cannot be visualized clearly until they have been macerated out of the host rock with hydrofluoric acid (Javaux et al. 2010). Many Proterozoic and younger cherts on Earth contain morphologically diverse filamentous and spheroidal microorganisms preserved by carbonaceous material, which can also be hundreds of micrometres in maximum dimension (e.g. Barghoorn and Tyler 1965; Schopf 1968; Garwood et al. 2020). However, these cannot be visualized clearly except in finely polished rock slabs or petrographic thin sections, which will not be produced by the rovers in situ. Nevertheless, pseudofossils on the scale of micrometres to centimetres may confound the search for life on Mars for two reasons. First, objects at the upper end of this size range will be visible to rover cameras and may be associated with other apparent lines of evidence for biogenicity (e.g. organic matter). Second, objects at the lower end may be detected by microscopy in samples cached by the Perseverance rover if and when they are eventually returned to Earth, or in future studies of martian meteorites.

The term ‘biomorph’ is particularly associated with carbonate–silica precipitates, but is used here in a wider sense to describe all small abiotic structures that bear a morphological resemblance to living or fossil microbes (including, for example, mineralized extracellular sheaths). Biomorphs have been discovered in numerous reaction–diffusion–precipitation systems, both in nature and in the laboratory, some of which are more familiar to palaeontologists than others. Chemical gardens were discovered in the seventeenth century and their potential to misleadingly resemble fossils has long been recognized (Liesegang 1914; Hawley 1926), but bears repeating (McMahon 2019). For reasons not altogether clear, candidate fossils are often compared with carbonate–silicate biomorphs, even when other known classes of biomorph are more similar (e.g. Dodd et al. 2017; Gan et al. 2021).

Classical chemical gardens

First described at the dawn of modern chemistry (Glauber 1646), chemical gardens are plant-like structures formed by reactions between transition metal salts and aqueous anionic solutions of, for example, silicate (Barge et al. 2015) (Fig. 2a–d). In the classic chemical garden experiment, a salt seed crystal begins to dissolve in a sodium silicate solution (c. pH 12), releasing acid that reacts spontaneously with the alkaline medium to produce a semipermeable membrane of gelatinous silica around the crystal (within seconds to minutes). Osmotic inflow increases the pressure inside the membrane until it ruptures, expelling a jet of acidic fluid that tends to rise buoyantly. New silica membrane ensheaths the expelled fluid, forming a thin tube that connects to the rupture point and is contiguous with the original membrane envelope. The tube may remain open at the tip and continue to grow for some time; it may also produce branches of equal or lesser thickness. Multiple tubes may develop from a single seed crystal (Fig. 2a). Growth of the chemical garden continues until the seed crystal has fully dissolved.

Fig. 2.

(a–d) Chemical gardens and (e–k) carbonate–silica biomorphs. (a) Rosettes of tubes emerging from seed crystals. (b) Serially branching, irregularly curved tubes. (c) Snail-like coiled tube. (d) Siliceous tube with rough Fe (oxyhydr)oxide-coated interior and hollow central cavity. (e, f) Worm-like braid, (g) helicoidal and (h) mushroom-like biomorphs. (i) Composite of globular, sheet-like and helicoidal morphologies. (j) Close-up of the area depicted by the white rectangle in part (i) showing dendrites produced by fractal growth. (k) Close-up of thin sinuous filaments and globular amorphous silica. Chemical garden images SM (parts a–c) or reproduced from McMahon (2019) under creative commons license (creativecommons.org/licenses/by/4.0/) (part d). Images of carbonate–silica biomorphs courtesy of J. Rouillard and S. Borenstazjn (Institut de Physique du Globe de Paris) (parts e–h) and P. Knoll and O. Steinbock (Florida State University) (parts i–k).

Fig. 2.

(a–d) Chemical gardens and (e–k) carbonate–silica biomorphs. (a) Rosettes of tubes emerging from seed crystals. (b) Serially branching, irregularly curved tubes. (c) Snail-like coiled tube. (d) Siliceous tube with rough Fe (oxyhydr)oxide-coated interior and hollow central cavity. (e, f) Worm-like braid, (g) helicoidal and (h) mushroom-like biomorphs. (i) Composite of globular, sheet-like and helicoidal morphologies. (j) Close-up of the area depicted by the white rectangle in part (i) showing dendrites produced by fractal growth. (k) Close-up of thin sinuous filaments and globular amorphous silica. Chemical garden images SM (parts a–c) or reproduced from McMahon (2019) under creative commons license (creativecommons.org/licenses/by/4.0/) (part d). Images of carbonate–silica biomorphs courtesy of J. Rouillard and S. Borenstazjn (Institut de Physique du Globe de Paris) (parts e–h) and P. Knoll and O. Steinbock (Florida State University) (parts i–k).

Although initially soft and flexible, chemical garden tubes harden and become brittle as metal (oxyhydr)oxides precipitate on their interior walls (McMahon 2019; Kotopoulou et al. 2021) (Fig. 2d). The tubes are typically highly circular in cross-section and can show a variety of interesting biomimetic morphologies formed during growth, including helical twisting, coiling (Fig. 2c), pinching, swelling, tapering, sinuously curving growth trajectories, serial bifurcation (branching) (Fig. 2b) and occasional anastomosis – that is, the reconnection of independent branches (Leduc 1911; McMahon 2019). Tube diameters range from micrometres to several millimetres depending on the experimental conditions, with a wall thickness controlled by the extent of metal oxyhydroxide and silica precipitation, which can largely occlude the internal space (resulting in a porous, but ‘filled’, filament rather than an open tube; McMahon 2019).

Many variations on this theme have been explored, including chemical gardens formed on the injection of acidic solutions rather than the dissolution of salts, ‘inverse' chemical gardens formed on the injection of the alkaline solution into the acid and quasi-2D chemical gardens confined in a narrow space between two flat plates (see Barge et al. 2015 for review). It is clear that chemical gardens can produce tubules from naturally occurring alkaline solutions and minerals, so they should not be dismissed as ‘exotic’ to natural environments (García-Ruiz 2000; García-Ruiz et al. 2017; McMahon 2019; McMahon et al. 2021).

As their name implies, chemical gardens are misleadingly life-like. Once solidified, they resemble fossil microorganisms, particularly Fe-mineralized fossil bacteria and fungi, in size, shape and composition. They may be relevant to the remarkable profusion of filaments composed of metal oxides, oxyhydroxides and (alumino)silicates found in siliceous and calcareous mineral deposits of all ages on Earth, including cavity fills in many volcanic and metavolcanic rocks (McMahon 2019; McMahon and Ivarsson 2019; McMahon et al. 2021). Many of these are probably fossils – indeed, fossils from a deep biosphere hosted in igneous rocks, with special relevance to some scenarios for life on Mars – but some are probably not and may result from chemical-garden-like processes and/or other types of self-organization that lead to filamentous crystals and aggregates (e.g. Hopkinson et al. 1998; Oaki and Imai 2003; Toramaru et al. 2003; Bonev et al. 2005).

The biotic interpretation receives support from the observed mineralizing behaviour of some modern organisms. Filamentous Fe-oxidizing bacteria, such as those in the SphaerotilusLeptothrix group, produce Fe-oxyhydroxide sheaths that conform to the filament surface, producing tubes that can bifurcate where cells divide, and are commonly vacated by cells leaving little or no organic residue (Emerson and Moyer 2002; Chan et al. 2016). The sheath exteriors can accumulate a flocculent Fe-oxyhydroxide coating up to several tens of micrometres thick (Schmidt et al. 2014). These microstructures are easily preserved in a silica matrix (chert) or in calcium carbonate and have a putative fossil record reaching back over million- to billion-year timescales (Dodd et al. 2017; Georgieva et al. 2021), although some of these purported fossils may be abiotic (Hopkinson et al. 1998; McMahon 2019; Johannessen et al. 2020; McMahon et al. 2021). Other Fe-oxidizing bacteria (e.g. Gallionella, Mariprofundus) secrete twisted, ribbon-like ‘stalks’ composed largely of Fe-oxhydroxides, which likewise can bifurcate during cell division (Chan et al. 2016). Again, however, there is a distinct potential for chemical gardens and other processes of crystallization and polycrystalline aggregate formation to generate twisted ribbons and other helical forms (Leduc 1911; García-Ruiz 2000; Oaki and Imai 2003; Jordan 2008).

Natural chemical gardens could plausibly have formed as a consequence of chemical weathering on Mars (García-Ruiz 2000; García-Ruiz et al. 2020; Sainz-Díaz et al. 2021). Soluble Fe and Mg sulfate salts have been identified at various martian localities (Johnson et al. 2007; Bishop et al. 2009; Ojha et al. 2015). Many of these likely formed in part through the dissolution and leaching of basaltic silicate minerals by sulfur-rich, acidic, oxidizing fumarolic fluids and vapours, which can be efficient even at very low temperatures (Tosca et al. 2004; Niles et al. 2017; Ruff et al. 2020); others probably formed by the aqueous oxidation of sulfide minerals (e.g. Dehouck et al. 2012). These sulfates would produce chemical gardens if introduced (as solids or in solution) to sufficiently silica-rich, alkaline fluids (e.g. in the subsurface).

Such fluids would have been supplied in some localities by serpentinization (i.e. the reaction of water with mafic and ultramafic minerals) and by the alteration of serpentinites, which, on Earth, can generate natural waters in which chemical gardens grow readily (García-Ruiz et al. 2017). Mineral assemblages suggestive of local serpentinization and serpentine alteration occur at several localities on Mars, including Jezero Crater (Ehlmann et al. 2010; Michalski et al. 2018; Brown et al. 2020; Zastrow and Glotch 2021). More generally, the copious production of silica-rich fluids on early Mars is attested by the widespread occurrence of hydrated silica deposits in fluviolacustrine and groundwater-influenced settings (Pan et al. 2021). It has been shown experimentally that chemical gardens can form even at temperatures and pressures characteristic of Amazonian and present-day Mars, close to the H2O triple point (Sainz-Díaz et al. 2021).

Carbonate–silica biomorphs

Carbonate–silica biomorphs are strikingly biomorphic inorganic objects composed of amorphous silica and crystalline carbonates of the alkaline earth metals (Ba, Sr or, more rarely, Ca) (Fig. 2e–k). First described by Juan Manuel García-Ruiz in the early 1980s (García-Ruiz and Amorós 1981a, b), they are formed in the laboratory through the slow crystallization of a carbonate phase under moderately alkaline conditions (c. pH 8.5–11) in the presence of silica, originally present as a silica gel or, in more recent work, in solution. They display dimensions ranging from a few micrometres to a few hundred micrometres and can adopt a wide diversity of life-like morphologies, most of which can be classified into three categories: (1) helicoidal filaments (Fig. 2g, i); (2) worm-like braids (Fig. 2e, f); and (3) leaf-like flat sheets (Fig. 2i) (Kellermeier et al. 2012c). Under some conditions, more atypical morphologies – such as flower-like forms, ‘trumpets’, ‘corals’, ‘moths’, ‘snails’ or ‘mushrooms’ (Fig. 2h) – can also be found (Opel et al. 2018; Rouillard et al. 2018).

In addition to their curved or sinuous overall shapes, in appearance more akin to the living than to the mineral world, carbonate–silica biomorphs also display complex internal structures and textures that are reminiscent of biological objects. They frequently possess a core–shell structure, with an enveloping ‘skin’ of amorphous silica and an internal core composed of aggregated carbonate nanocrystals. The following section on mineralogical signatures shows that the polycrystalline core of the carbonate–silica biomorphs can be described as a quasi-mesocrystal, a characteristic of minerals formed by living organisms.

The complex morphologies of the carbonate–silica biomorphs emerge from purely inorganic mineral growth mechanisms. In an initial fractal growth stage, dendrites (Fig. 2j), dumbbells, framboids and spheroids are formed. Fractal growth here is due to silica poisoning of the surface of the carbonate crystallites, causing them to repeatedly split at non-crystallographic angles. In a second stage called ‘curvilinear growth’, polycrystalline mineral sheets grow radially in two dimensions, forming flat structures such as discs and leaves. Curling at the margins of these sheets introduces curvature and leads to the development of twisted shapes, such as helicoidal filaments and braids (García-Ruiz et al. 2009; Kellermeier et al. 2012a; Rouillard et al. 2018). The growth of these curling sheets is fed by an autocatalytic process in which carbonate precipitation locally decreases the pH, triggering the polymerization of amorphous silica, coating and cementing the carbonate crystallites and preventing further growth. In turn, silica polymerization locally increases the pH, allowing the precipitation of new carbonate building blocks (Kellermeier et al. 2012b).

Although purely inorganic, the curved and sinuous shapes of the biomorphs evoke biological objects, such as helical and segmented filamentous microbes, protists and even plants and animals. Interestingly, the hydrophobic surfaces of carbonate–silica biomorphs readily incorporate or adsorb organic molecules (García-Ruiz et al. 2002; Opel et al. 2016), which, as already discussed here, may have an abiotic origin on Mars. Although they have (so far) not been observed to form spontaneously in the environment, carbonate–silica biomorphs were obtained experimentally from natural alkaline silica-rich spring waters derived from serpentinization environments (García-Ruiz et al. 2017) and they may have formed in some hydrothermal environments of the early Earth (García-Ruiz 1994; García-Ruiz et al. 2020). Similar geochemical conditions, also conducive to the formation of chemical gardens, may have prevailed in alkaline, silica-rich fluids associated with serpentinization on early Mars.

Fibrous crystals, trichites and other crystallites

Strong or irregular curvature induced during the growth or deformation of fibrous or acicular crystals can generate shapes reminiscent of filamentous microorganisms. Muscente et al. (2018) described curving Mn oxide (MnO) crystals of probable metasomatic origin preserved in Neoarchean chert from the Pilbara Craton of Australia and noted that their simple morphology could be compared with some purported filamentous microfossils. In addition to Mn oxides, many clay and serpentine group minerals known or predicted to exist on Mars also show a fibrous habit, including sepiolite–palygorskite, celadonite and serpentine (Bishop et al. 2008; Ehlmann et al. 2010; Bristow and Milliken 2011). In addition, spectacularly complex and varied thread-like crystallites (‘trichites’) of pyroxene and other igneous minerals are found in volcanic and impact-related glasses, formed by the quenching of molten rock. Glass fragments are expected to be widespread in wind-blown sediments on Mars, albeit in a chemically weathered condition (Horgan and Bell 2012).

On Earth, trichites in natural glass are readily visible in thin section, where their resemblance to fossil microorganisms was first noted by Cloud (1976). They can be tightly coiled, looped, branched and irregularly curving (worm-like); they are sometimes found in spider-like starburst arrangements (Ross 1962; Engelhardt et al. 1995) (see Fig. 4a). They are commonly subdivided into smaller aligned crystallites, giving them a segmented appearance. Trichites commonly nucleate on the walls of vesicles formed by trapped volatiles. Glass vesicles themselves can also resemble microfossils, as noted with respect to some scoria fragments found in the 3.4-Gyr-old Strelley Pool Formation of Western Australia, where vesicles are coated with kerogen-like carbonaceous material (Wacey et al. 2018).

It is unlikely that fibrous crystals, trichites and other crystallites would be large and distinct enough to resolve with Mars rover cameras (Table 1) and yet at the same time small enough to resemble microorganisms. They may be discovered in returned samples, but if they are well preserved, then it would then be straightforward to exclude them from consideration as microbial fossils because they can be recognized by their composition, crystallographic parameters and angular, faceted cross-sections. Diagnosis may be more difficult, however, where mineral particles have decomposed or altered to produce substances and morphologies more similar to mineralized microfossils, with softened/rounded cross-sections and hollow interiors. Thus, we tentatively suggest that solid and hollow filaments observed in impact glass by Sapers et al. (2014) could be modified trichites of some kind (cf. Pickersgill et al. 2021).

McMahon et al. (2021) described inorganic (Mg silicate + Fe oxide) filamentous dubiofossils occurring in calcite veins from the pervasively oxidized serpentinites of NW Italy and speculated that they might represent coated and recrystallized mineral fibres, among other possible explanations. A misleading association with organic matter can also complicate the interpretation of mineral grains. Horodyski (1981) noted that apatite grains in a Mesoproterozoic shale had acquired a coating of organic matter that created a superficial resemblance to fossil microorganisms. Similarly, Wacey et al. (2016b) showed that carbonaceous microstructures in the 3.5-Gyr-old Apex Chert, previously interpreted as fossil filamentous bacteria, are, in fact, altered particles of potassium mica, which were expanded into worm-like structures (the mineral vermiculite) by hydrous alteration and then coated at a late stage by carbon ‘moving around the system'. Hematite aggregates in the same chert unit are also filament-like and have been described as ‘pseudomicrofossils’ (Marshall et al. 2011). Even genuine microfossils can be secondarily enriched in carbon from such external sources (Rasmussen et al. 2021).

According to biogenicity determination protocols, life-like morphologies with organic composition score more highly than inorganic life-like morphologies (e.g. McLoughlin and Grosch 2015; Neveu et al. 2018). Nevertheless, organic materials are no less capable than mineral particles of forming a wide variety of biomorphs; indeed, structural self-organization in the organic milieu necessarily preceded the origin of life on Earth (and perhaps Mars). Known organic biomorphs can be divided into the recently discovered carbon–sulfur biomorphs and the morphologically simpler spheroids that result from hydrophobic interactions with water. We are aware that chemical gardens can also be produced from organic salts and solutions, but (so far) the known examples are of limited geological relevance (e.g. Bernini et al. 2021).

Carbon–sulfur biomorphs

Carbon–sulfur biomorphs are microscopic spherical and tubular objects composed of elemental sulfur encapsulated within an organic envelope (Fig. 3). Their sizes range from a few tens of nanometres to a few micrometres in diameter. Carbon–sulfur biomorphs form spontaneously in aqueous solutions in which hydrogen sulfide is oxidized in the presence of dissolved organic molecules (including simple prebiotic organic molecules such as glycine) (Cosmidis and Templeton 2016; Cosmidis et al. 2019). The elemental sulfur core of the biomorphs is the product of chemical sulfide oxidation, while their organic envelope is thought to form through the self-assembly of amphiphiles produced by the reaction of organic molecules with sulfides and polysulfides in solution (Cosmidis et al. 2019). Observations of the early stages of carbon–sulfur biomorph formation reveal numerous empty vesicles, suggesting that the organic envelope may form prior to the sulfur core (Fig. 3c).

Fig. 3.

Carbon–sulfur biomorphs. (a) Complex network formed by filaments and spheres. (b) Rosette. (c) Empty organic vesicles during early stages of carbon–sulfur biomorph formation. Small organic vesicles later coalesce into c. 1 μm spheres. (d) Aggregate of spheres and filaments. (e) Spheres with visible organic envelopes. (f) Straight and helical filaments. Images adapted from Nims et al. (2021) (parts a and b) or JC (parts c and f) and courtesy of C. Nims (University of Michigan) (parts d and e).

Fig. 3.

Carbon–sulfur biomorphs. (a) Complex network formed by filaments and spheres. (b) Rosette. (c) Empty organic vesicles during early stages of carbon–sulfur biomorph formation. Small organic vesicles later coalesce into c. 1 μm spheres. (d) Aggregate of spheres and filaments. (e) Spheres with visible organic envelopes. (f) Straight and helical filaments. Images adapted from Nims et al. (2021) (parts a and b) or JC (parts c and f) and courtesy of C. Nims (University of Michigan) (parts d and e).

Fig. 4.

Other types of organic and inorganic biomorphs. (a) Asteroidal trichites radiating from central magnetite grains in a natural volcanic glass (Cougar Mountain, Oregon). (b) Organic microspheres obtained from RNA and quartz mixtures under diagenetic conditions (Criouet et al. 2021). (c) Amphiphilic spherical and tube-like vesicles formed by mixed fatty acids and 1-alkanols (C10–C15) in 10 mM CaCl2. Courtesy of S.F. Jordan. (d) Coacervate microdroplets prepared in water by mixing PDDA with ATP (M. Li et al. 2014). (e, f) Chain-forming and twinning proteinoid microspheres obtained by heating mixtures of amino acids (Fox and Yuyama 1963). (g, h) Apatite particles formed by double diffusion in gelatin mimicking dividing cells and raised polygonal features in embryos (Crosby and Bailey 2017). (i) Rod-shaped silica particles formed on the walls of basalt fractures injected with supercritical CO2 (Schaef et al. 2011). (j) Cryogenic opal filaments and sheets formed in rapidly freezing silica-rich fluids (Channing and Butler 2007). (k) Cryogenic carbonate segmented filaments obtained by freezing of silica-rich alkaline brines (Fox-Powell and Cousins 2021). (l) Rod-shaped apatite particles forming within organic films in sediments (Mänd et al. 2018). (m) Fluoroapatite particle precipitated in the presence of citrate, mimicking dividing coccoid cells (Wu et al. 2010). Images reproduced with permissions (parts a, d–h, j–m) or under creative commons licenses (part b, creativecommons.org/licenses/by-nc-nd/4.0/; part i, creativecommons.org/licenses/by-nc-nd/3.0/).

Fig. 4.

Other types of organic and inorganic biomorphs. (a) Asteroidal trichites radiating from central magnetite grains in a natural volcanic glass (Cougar Mountain, Oregon). (b) Organic microspheres obtained from RNA and quartz mixtures under diagenetic conditions (Criouet et al. 2021). (c) Amphiphilic spherical and tube-like vesicles formed by mixed fatty acids and 1-alkanols (C10–C15) in 10 mM CaCl2. Courtesy of S.F. Jordan. (d) Coacervate microdroplets prepared in water by mixing PDDA with ATP (M. Li et al. 2014). (e, f) Chain-forming and twinning proteinoid microspheres obtained by heating mixtures of amino acids (Fox and Yuyama 1963). (g, h) Apatite particles formed by double diffusion in gelatin mimicking dividing cells and raised polygonal features in embryos (Crosby and Bailey 2017). (i) Rod-shaped silica particles formed on the walls of basalt fractures injected with supercritical CO2 (Schaef et al. 2011). (j) Cryogenic opal filaments and sheets formed in rapidly freezing silica-rich fluids (Channing and Butler 2007). (k) Cryogenic carbonate segmented filaments obtained by freezing of silica-rich alkaline brines (Fox-Powell and Cousins 2021). (l) Rod-shaped apatite particles forming within organic films in sediments (Mänd et al. 2018). (m) Fluoroapatite particle precipitated in the presence of citrate, mimicking dividing coccoid cells (Wu et al. 2010). Images reproduced with permissions (parts a, d–h, j–m) or under creative commons licenses (part b, creativecommons.org/licenses/by-nc-nd/4.0/; part i, creativecommons.org/licenses/by-nc-nd/3.0/).

As a result of their spherical and tubular morphologies, carbon–sulfur biomorphs are reminiscent of microbial spherical and filamentous cells. These structures display more complex morphologies than the spheroids and aggregates produced by other known biomorph systems (see previous section). The tubular biomorphs may branch at 90° and 45° angles (Cosmidis and Templeton 2016), similar to some filamentous bacteria (e.g. Streptomyces), and they can be helicoidal (Fig. 3f), similar to helical bacteria (e.g. cyanobacteria of the genus Oscillatoria, spirochaetes). Tubular carbon–sulfur biomorphs can also form rosettes (Fig. 3b), mimicking those formed by filamentous microbes such as Thiothrix.

Although carbon–sulfur biomorphs are originally composed of organic carbon and elemental sulfur, experiments have shown that their sulfur core rapidly diffuses away during early diagenesis in silica-rich environments, leaving behind empty organic envelopes (Nims et al. 2021). The morphologies of the biomorphs are finely preserved through this diagenetic process as a result of the rapid precipitation of nanocolloidal silica at the surfaces of the organic spheres and tubes (Nims et al. 2021). The resulting silicified organic microstructures are strikingly similar to coccoidal and filamentous organic microfossils preserved in Precambrian cherts by the rapid precipitation of silica at the seafloor, as found, for instance, in the 1.9-Gyr-old Gunflint Formation in Canada (Wacey et al. 2013) or the c. 2.4-Gyr-old Turee Creek Group in Western Australia (Schopf et al. 2015; Barlow and Kranendonk 2018).

Mars has long been considered a ‘sulfur-rich planet’ (King and McLennan 2010), but, as a result of the oxidizing conditions at the surface, the vast majority of this sulfur is currently present as sulfate minerals (Franz et al. 2019). However, in situ sulfur isotope analyses acquired by NASA's Curiosity rover in Gale Crater suggest the existence of an active and fully developed sulfur redox cycle on early Mars (Franz et al. 2017). Evolved gas analyses performed by Curiosity's SAM instrument showed that reduced sulfur phases are present within ancient fluviolacustrine sediments at Gale Crater (McAdam et al. 2014; Wong et al. 2020). SAM also detected evidence for sulfurized organic matter in the 3.5-billion-year-old Murray Formation, indicating interactions between organic molecules and reduced sulfur (Eigenbrode et al. 2018; Heinz and Schulze-Makuch 2020).

Although the full space of physicochemical parameters permitting the formation of carbon–sulfur biomorphs remains to be explored, they have been synthetized in the laboratory at temperatures ranging from 4°C to room temperature, pH values ranging from acidic (c. pH 4) to slightly basic (c. pH 8) and salinities ranging from c. 0% (deionized water) to 3.5% (seawater), encompassing plausible conditions for wet environments on early Mars (Fukushi et al. 2019). As stated earlier, carbon–sulfur biomorphs can be finely preserved through silicification in silica-rich depositional settings, mimicking fossilization in cherts, a major preserving medium for many Archean and Proterozoic putative microfossils on Earth (Schopf 2006; Javaux and Lepot 2018). On Mars, rapidly forming silica deposits may have formed in different types of ancient sedimentary environments (reviewed by McMahon et al. 2018), allowing the preservation of carbon–sulfur biomorphs if they were ever present on that planet.

Other organic biomorphs

Most organic compounds are hydrophobic and tend to aggregate into spheres in the presence of water, commonly with dimensions similar to microorganisms. Thus, spherical organic particles were probably fairly abundant in any prebiotic milieu on early Mars (and Earth). Some of them would have had non-negligible potential to be preserved in the rock record. For example, research in applied chemistry has found that complex organic matter subjected to ‘hydrothermal carbonization’ (devolatilization in water at c. 200°C) assembles into stable, solid spheres similar in size to bacteria, with hydrophobic aromatic cores and hydrophilic shells (‘hydrochar’; Sevilla and Fuertes 2009).

Criouet et al. (2021) found that similar spheres were formed when a mixture of RNA, quartz and water was held at 200°C under Ar for 20 days. These objects were c. 0.5–5 μm in diameter, variously textured and commonly connected together like dividing cells (Fig. 4b). Their composition, as probed, for example, by carbon X-ray absorption near-edge structure (C-XANES), resembles fossil bacteria more closely than it resembles the original RNA. Criouet et al. (2021) infer from previous experimental work on hydrochar that further diagenetic alteration of these structures could render them hollow and thus even more similar to fossil bacteria. The viability of an RNA world on early Mars has been explored – and broadly supported – in light of geochemical considerations (Mojarro et al. 2021), but, in any case, similar spheroids can very probably result from many other organic precursors. Indeed, similarly robust hollow hydrocarbon spheroids <1–50 μm in diameter were recovered from deep groundwater in the Witwatersrand Basin of South Africa, where they appear to have formed through the abiotic reorganization of thermally altered (ultimately biogenic) organic matter (Wanger et al. 2012). Sub-micron hollow organic ‘globules’ have also been identified in carbonaceous meteorites (Nakamura-Messenger et al. 2006).

Fatty acids produced by abiotic Fischer–Tropsch type reactions under hydrothermal conditions, like other amphiphilic compounds, readily self-assemble in water to form membranous vesicles from nanometres to hundreds of micrometres in size. Such vesicles may have played a part in the origin of life by encapsulating protocells (e.g. Deamer 2017). The stability of these vesicles under different conditions depends on the mixture of amphiphiles present. Experiments have shown that relatively long-chain (C10–C15) fatty acids and 1-alkanols generate vesicles in the warm, saline, alkaline waters characteristic of serpentinization-driven marine hydrothermal systems on the early Earth and Mars (Fig. 4c) (Jordan et al. 2019). Vesicle formation by fatty acids is also favoured over a wide pH range in non-marine hot spring settings (Deamer 2021). Lipid vesicles are typically spheroidal, but can extend into tubules and thread-like filaments (e.g. Zhu and Szostak 2009; Deamer 2021). The potential for any of these rather delicate structures to be preserved as morphological fossils is unclear, but has sometimes been mooted (e.g. Javaux et al. 2013).

Other structures discussed in prebiotic chemistry also have relevance as potential pseudofossils. Coacervates – microdroplets of viscous fluid that separate out from a colloidal solution (Fig. 4d) – can incorporate a variety of inclusion structures and were first suggested as possible protocells nearly a century ago (Oparin 1924). Coacervates have been produced from mononucleotides and peptides of surprisingly low molecular weight (Koga et al. 2011). Other coacervates based on similarly small molecules have been silicified in the laboratory to form solid spheroids with multiple sub-compartments, albeit through a rather non-naturalistic chemical process (Fothergill et al. 2014).

‘Proteinoid microspheres’ represent another spheroidal biomorph system once thought to be a plausible model for prebiotic protocells. These were discovered by Sidney Fox and co-workers, who showed that spheroids 1–2 μm in diameter self-assemble from thermally polymerized mixtures of dry amino acids (e.g. at >130°C for 3 h) after the addition of water (e.g. Fox and Yuyama 1963; Brooke and Fox 1977). These objects show a range of biomimetic features, including internal spheroids, sharply defined single or concentric double boundaries, ‘twinned’ growth strongly reminiscent of cells dividing, and aggregation into chains and networks (Fig. 4e, f). These pseudo-cells could presumably have arisen spontaneously on the early Earth or Mars if amino acids (e.g. from meteorites, atmospheric chemistry and hydrothermal reactions) became highly concentrated and subject to dry heat. It has been shown that some of these proteinoid microspheres are readily silicified to form pseudofossils with good preservation potential that can resemble eukaryotes as well as prokaryotes (Francis et al. 1978). Thus, as noted earlier, the availability of silica in both groundwaters and surface waters on early Mars could have assisted the preservation of both pseudofossils and fossils.

Very few natural or experimental systems have yet been studied in relation to their capacity to produce biomorphs in geochemically plausible settings, yet there are some tantalizing hints scattered across the scientific literature. For example, Oaki and Imai (2003) demonstrated the formation of smoothly curving filamentous dendritic aggregates from the precipitation of potassium dichromate in gelatin cooled from 100°C, a finding that may be relevant to the origin of similar structures found in some agates (which precipitate from silica gel). More generally, minerals grown in gels can sometimes adopt rounded and complex shapes (e.g. Dominguez Bella and García-Ruiz 1987; Crosby and Bailey 2017) (Fig. 4g, h).

Silica deposits formed on the walls of basalt fractures injected with wet, supercritical CO2 at sterilizing temperatures resemble rod-shaped and filamentous bacteria (Schaef et al. 2011) (Fig. 4i). Biomorphic particles of opal (Channing and Butler 2007) and carbonates (Fox-Powell and Cousins 2021) resembling microbial filaments precipitate from rapidly freezing mineral-laden waters (Fig. 4j, k), a process which may be particularly relevant for ancient hydrothermal systems on Mars. Abiotic silica spheroids found in hot spring ponds of the Dallol geothermal field, Ethiopia form botryoidal clusters that resemble coccoidal bacterial colonies (Belilla et al. 2019). All these systems require further investigation to elucidate new classes of geologically relevant biomorphs on Mars.

In addition to these processes, mineral structures can often adopt rounded, curved or complex life-like shapes when formed in the presence of organic molecules (which can be abiotic in origin), as already illustrated for carbon–sulfur biomorphs. More generally, ‘organominerals’ (i.e. minerals in which formation is mediated by organic molecules) (Défarge 2011) can present unusual, non-crystallographic morphologies as a result of the templating effect of the organic matrices on which they nucleate, or under the influence of organic molecules in solution during mineral growth. Soluble organic molecules can influence mineral shapes by poisoning some crystal faces after preferential adsorption, preventing further growth, or by stabilizing amorphous precursors and favouring their assembly into complex structures (Gower 2008). Organic molecules can promote the fractal growth of crystals, leading to the formation of rod-shaped particles (Fig. 4l), dumbbells, spherules, divided spheroids mimicking cell division (Fig. 4g, h, m) and ‘cauliflower-like’ structures (Busch et al. 1999; Wu et al. 2010; Sand et al. 2012; Crosby and Bailey 2017; Mänd et al. 2018).

A particularly relevant example for astrobiology is the observation of calcium carbonate particles precipitated in the presence of organic matter extracted from the Murchison CM2 meteorite. The minerals present rounded morphologies reminiscent of coccoid and rod-shaped bacteria (Reitner 2004) and similar to the calcareous structures previously interpreted as fossil microorganisms in the martian meteorite ALH84001 (McKay et al. 1996). More exotic mineral shapes – such as trumpets, helices, twisted ribbons and stars – can also be obtained through precipitation in the presence of large organic molecules, particularly polymers (Mukkamala and Powell 2004; Yu and Cölfen 2004).

Microorganisms can dissolve, etch and bore into solid substrates (Cockell and Herrera 2008). Bioalteration textures apparently resulting from these processes have been reported in various geological materials, most commonly basaltic glass, which is probably fairly widespread on Mars. Experiments have confirmed that bacteria etch cell-sized pits in volcanic glass in the laboratory (Thorseth et al. 1995). Alteration features in natural samples divide into two types: granular textures and tubular textures (extensively illustrated by Staudigel et al. 2006; Fisk and McLoughlin 2013). The granular type consists of closely packed, micrometre-scale spheroidal cavities, giving a spongy overall appearance. The tubular type consists of elongated cavities c. 1–6 μm wide, which extend up to hundreds of micrometres from altered glass surfaces and fractures (McLoughlin et al. 2019). Complex, biogenic organic material has been found in close association with both types (e.g. Fisk et al. 1998; Preston et al. 2011; Wacey et al. 2014), although some examples are less convincing on close inspection than they first appeared (McLoughlin and Grosch 2015; Wacey et al. 2017).

Experimental work has partially succeeded in mimicking the granular form abiotically by reacting basaltic glass with seawater at 150°C (a sterilizing temperature) for 48 days (McCollom and Donaldson 2019). Tubular textures, however, have never been reproduced in experiments, either with or without microbes. Some examples are more compellingly life-like than others, showing a variety of curving, twisting, branching, coiling and other elaborate morphologies suggestive of microbial growth. High-resolution microanalysis has shown that some of the simpler tubules are probably abiotic pseudofossils of uncertain origin (e.g. Pedersen et al. 2015). Titanite-filled ‘microborings’ in Archean basalts from South Africa and Western Australia have also now been convincingly reinterpreted as abiotic mineral growths, probably resulting from thermal metamorphism and seafloor hydrothermal recrystallization, respectively (Grosch and McLoughlin 2014, 2015; McLoughlin et al. 2020). These same processes could probably have generated analogous pseudofossils in basalts on Mars.

Relatedly, White et al. (2014) found linear microtextures and organic carbon associated with fracture-filling alteration minerals in the basaltic martian meteorite Yamato 000593 and interpreted them as possible evidence of microbial bioalteration. This view was rejected by McLoughlin et al. (2019), who showed that the textures were not tubular, but fracture-like, and that the carbon was most likely of hydrothermal origin. McLoughlin and Grosch (2015) present biogenicity criteria for alteration textures in (meta)volcanic rocks. Significantly, they argue that ‘no terrestrial example yet described’ achieves the maximum score of ‘Category 5’ necessary to demonstrate a biogenic origin.

Some features resembling ‘microborings’ are, in fact, ambient inclusion trails (AITs): snaking, tunnel-like cavities (which may be mineral-filled) formed by the propulsion of mineral grains through solid substrates, including quartz/chalcedony (Fig. 5a), phosphate minerals, chlorite and (rarely) altered volcanic glass (Wacey et al. 2008; Lepot et al. 2009, 2011; McLoughlin and Grosch 2015). AITs have also been found in clastic sedimentary rocks of Precambrian age (Fig. 5b). The cross-sectional shape and diameter of AITs are determined by the formative grain, most commonly a euhedral pyrite or magnetite crystal c. 0.5–10 μm across. Longitudinal marks seen in the trail walls are striations made by the grain vertices. The grain is sometimes still present at the end of the AIT. AITs are sometimes found in starburst patterns, indicating that the grains moved outward from a common centre (e.g. Knoll and Barghoorn 1974; Lepot et al. 2009). Examples in Paleoproterozoic chert were misdiagnosed as fossil algae by Gruner (1924) and re-classified as pseudofossils by Barghoorn and Tyler (1963).

Fig. 5.

Ambient inclusion trails (AITs). (a) AITS resembling microborings in agate (basalt-hosted chalcedony) from the Pentland Hills Volcanic Formation, Edinburgh, Lower Devonian. Note opaque terminal grains. Image: SM. (b) Titanite-filled AITs occurring in fine-grained sedimentary rock from the Late Ediacaran Gibbet Hill Formation of Newfoundland. Some trails terminate in pyrite crystals (not visible in this field of view). Image courtesy of J.J. Matthews (Oxford University Museum of Natural History).

Fig. 5.

Ambient inclusion trails (AITs). (a) AITS resembling microborings in agate (basalt-hosted chalcedony) from the Pentland Hills Volcanic Formation, Edinburgh, Lower Devonian. Note opaque terminal grains. Image: SM. (b) Titanite-filled AITs occurring in fine-grained sedimentary rock from the Late Ediacaran Gibbet Hill Formation of Newfoundland. Some trails terminate in pyrite crystals (not visible in this field of view). Image courtesy of J.J. Matthews (Oxford University Museum of Natural History).

AITs typically contain some organic matter and Wacey et al. (2016a) suggested that grain propulsion can be driven by microbially mediated decomposition of this material, which would make some AITs indirect biosignatures after all (Wacey et al. 2016b). However, there is evidence to suggest that the thermal decomposition of organic matter (which can be abiogenic) and metamorphic reactions involving chlorite might have formed some AITs abiotically – for example, in volcanic rocks free of biological influence (Lepot et al. 2009, 2011; Wacey et al. 2016a; Fig. 5a). Thus, these structures cannot be regarded as unambiguous biosignatures despite their life-like morphology and partly organic composition.

Several previous reviews have focused on the use of minerals as biosignatures (Benzerara and Menguy 2009; Benzerara et al. 2019), including in the context of the search for life on Mars (Banfield et al. 2001). Biominerals – that is, minerals whose formation is controlled by living organisms or, at a minimum, induced or influenced by them (see Lowenstam 1981; Dupraz et al. 2009 for definitions of these terms) – can possess specific morphological, structural, textural or chemical characteristics that may allow their discrimination from abiotically produced minerals.

The most obvious feature common to many biominerals is their unusual morphology. Unlike inorganically precipitated minerals, which typically adopt crystal-like prismatic shapes, with sharp angles and straight faces, biominerals often have rounded or curved surfaces and overall morphologies that can be extremely complex. Although this is especially true of eukaryotic biominerals (e.g. the delicate architecture of the silica frustule of a diatom), bacterial biominerals can also display elaborate, non-crystallographic shapes, such as, for instance, the twisted iron-rich stalks and tubular sheaths produced by some microaerophilic Fe-oxidizing bacteria (Chan et al. 2009, 2011) or the sinuous elemental sulfur filaments produced by the sulfur-oxidizing bacterium Candidatus Arcobacter sulfidicus (Taylor and Wirsen 1997; Sievert et al. 2007). Experimental diagenesis (Picard et al. 2015), as well as studies of the geological record (Slack et al. 2007; Hofmann et al. 2008), show that these morphological characteristics can be preserved in rocks. However, as demonstrated in the section on biomorphs, a wide range of abiotic processes can produce mineral, organic and organic–mineral structures with similar life-like shapes, making morphology an ambiguous indicator of biogenicity (García-Ruiz et al. 2002).

Organic–mineral associations

Biominerals are commonly intimately associated with organic molecules and biominerals can often be described as inorganic–organic composite materials. This is especially true of Metazoan biominerals, which tend to be formed within matrices or vesicles composed of organic macromolecules controlling mineral nucleation, crystallization and growth (Weiner 2008; Addadi and Weiner 2014). In microbial systems, biominerals may be enclosed within intracellular organelles, which can be lipidic – as in the case of the magnetosome, an invagination of the cytoplasmic membrane encapsulating magnetic Fe minerals in magnetotactic bacteria (Uebe and Schüler 2016) – or proteinic – as in the case of the envelope surrounding elemental sulfur minerals in some sulfur-oxidizing bacteria (Dahl 2020). Microbial biominerals may also form within the microbial cell wall (Benzerara et al. 2004; Cosmidis et al. 2015), on the cell surface (e.g. on S-layers; Schultze-Lam et al. 1992; Phoenix et al. 2005), on extracellular polymeric substances (Dupraz et al. 2009; Couasnon et al. 2020) and polysaccharidic stalks and sheaths (Chan et al. 2009) or within extracellular organic vesicles (Cron et al. 2019; Marnocha et al. 2019).

Diagenesis experiments have shown that some of the resulting biomineral–organic assemblages may withstand burial at high temperatures and pressures (J. Li et al. 2013, 2014) and observations from the geological record confirm that minerals precipitating on or within microbial cells can lead to well-preserved organic–mineral associations in rocks (e.g. Cosmidis et al. 2013).

The instruments onboard the most recent and future martian rovers have the capability to detect and characterize intimate mineral–organic associations at high spatial resolution, such as SHERLOC on the Perseverance rover and MicrOmega on the Rosalind Franklin rover (using deep UV Raman and near-infrared/visible light spectromicroscopy, respectively; Table 1) (Beegle et al. 2015; Bibring et al. 2017). However, such close associations are expected to be found in non-biological contexts whenever minerals form in the presence of organic molecules (which again may be abiotic in origin).

Organic molecules can be passively adsorbed onto minerals through different types of binding interactions (e.g. Lagaly et al. 2013). Under certain conditions, mineral surfaces can even participate in the polymerization of these adsorbed molecules into larger, life-like macromolecules (as, for instance, in the surface-aided polymerization of amino acids on silicates, oxides and sulfides; Lambert 2008). As shown with the example of carbon–sulfur biomorphs, mineral precipitation in the presence of organic molecules may lead to self-assembled organic–mineral structures mimicking biomineral encapsulation within organic vesicles (e.g. Cron et al. 2019; Marnocha et al. 2019). Different processes on Mars may have led to the preservation of organic molecules at the surface of, or within, mineral particles (reviewed in Fornaro et al. (2018), but for the reasons cited here it is unlikely that the resulting mineral–organic assemblages, even if combined with biological morphologies, may serve as convincing biosignatures.

Mineral structures

Biominerals may present crystal structures that differ from those of their abiotically precipitated counterparts and whose formation is thermodynamically unfavourable under normal temperature and pressure conditions. For example, unstable phases of elemental sulfur, the monoclinic allotropes β-S8 and ɣ-S8, can be formed as biominerals by sulfur-oxidizing bacteria (Cron et al. 2019) and have been proposed as potential biosignatures in astrobiological investigations (Gleeson et al. 2012).

The presence of metastable crystal phases in biominerals likely results from mineral formation on or within organic matrices acting like templates for nucleation (Falini et al. 1996), the stabilization of unstable mineral intermediates by organic molecules (e.g. amorphous calcium carbonate; Addadi et al. 2003) or modifications of the local chemical environment (for instance, increased supersaturation) within internal vesicles (Sviben et al. 2016; Uebe and Schüler 2016) or in the extracellular medium as a result of metabolic activity (Rodriguez-Navarro et al. 2007). Some aspects of the crystal structures of minerals present in martian rock and soil samples can be determined using different techniques such as Raman spectroscopy (SHERLOC and the Raman laser spectrometer onboard the Perseverance and Rosalind Franklin rovers, respectively; Table 1) (Beegle et al. 2015; Rull et al. 2017), although full crystallographic characterization will require sample return.

However, mineral precipitation in the presence of (non-biological) organic molecules can also induce the formation of metastable polymorphs, as demonstrated many times in laboratory experiments. For instance, a range of organic substances, such as alcohols (Sand et al. 2012), citrates (Tobler et al. 2015), amino acids (Hood et al. 2014; Tobler et al. 2014) and polymers (Xu et al. 2008; Sonobe et al. 2015) can impact calcium carbonate polymorphism and induce the formation of unstable carbonate phases such as aragonite and vaterite. The high-temperature sulfur allotropes β-S8 and ɣ-S8 can form abiotically at room temperature when sulfide is oxidized in the presence of different organic molecules, including non-biogenic organic molecules such as glycine (Cosmidis et al. 2019).

Precipitation under inorganic conditions can also produce thermodynamically unstable polymorphs. In the calcium carbonate system, a number of geochemical factors – such as rapid CO2 degassing, high supersaturation or the presence of Mg and other divalent cations – can induce the formation of aragonite instead of calcite in natural springs (Jones 2017). Although rare on Earth, vaterite was found in an extremely cold and dry environment that may be a good analogue for carbonate deposits on other planetary bodies, but its formation and persistence was thought to be abiogenic and favoured by the extreme cold and high-pH waters (Grasby 2003).

These examples illustrate the fact that in order to interpret the presence of unstable polymorphs or other information on crystal structure as biosignatures, the geochemical context of mineral formation (including the fluid chemistry and presence and nature of organic molecules) must be fully known. This may be extremely difficult when investigating ancient sediments on early Mars. Other processes that may alter mineral structure after their formation (e.g. recrystallization) must also be taken into account – for instance, ageing, desiccation and exposure to high-temperature conditions (due to impact and volcanism, even in the absence of a strong geothermal influence during burial).

Beyond general crystal structure, some crystallographic features of biominerals have been proposed as potential biosignatures – for instance, crystal purity and the absence of defects in biogenic magnetites (Fischer et al. 2011; Li et al. 2015). The crystallographic properties of bacterial Mn oxides, such as an abundance of lattice vacancies, have been proposed as mineralogical biosignatures that could be detected on Mars using electron paramagnetic resonance spectroscopy (Kim et al. 2011). More work is needed on a wider range of biominerals and abiotic systems to validate these approaches.

Primary mineral textures

The texture of biominerals, defined here as crystallite size and organization, often differs strongly from that of abiotically precipitated minerals. This is partly due to their crystallization mode: although classical crystal growth proceeds by ion-by-ion addition, biomineralization often proceeds through the addition of pre-formed mineral nanoparticles and, as a result, many biominerals are actually polycrystalline materials (De Yoreo et al. 2015). When the nanoscale building blocks share a common crystallographic orientation, the resulting mineral is described as a mesocrystal (Sturm and Cölfen 2016). Examples are found in coral skeletons, the nacre layer of shells, sea urchin spines and bones (Oaki et al. 2006; Olszta et al. 2007; Seto et al. 2012). Mesocrystals have also been described in calcium carbonates forming in microbial mats, stromatolites and laboratory cultures (Benzerara et al. 2010; Peng and Jones 2013; Han et al. 2017).

Intracellular magnetic iron biominerals in magnetotactic bacteria are usually elongated along specific crystallographic directions and aligned into a chain within the microbial cell (Amor et al. 2020) and hence may be described as 1D mesocrystals (Bergström et al. 2015). These specific textural characteristics are encountered in a fraction of the magnetite nanocrystals present in the martian meteorite ALH84001 and they have thus been described as ‘magnetofossils’ (Friedmann et al. 2001; Thomas-Keprta et al. 2002), although others have argued that their crystallographic and textural signatures could be mimicked abiotically (Golden et al. 2001, 2004; Bell 2007; Martel et al. 2012). Studies of ancient eukaryotic biominerals show that nanocrystalline textures can be preserved in the geological record (Gilbert et al. 2019) and the mesocrystal texture of calcite particles has been used as evidence of microbial influence in the formation of Paleogene micrite deposits (Bralower et al. 2020).

However, polycrystalline mineral textures mimicking biominerals, including mesocrystals, can be reproduced in the laboratory, notably through precipitation with organic additives. Organic molecules appear to play an important part in crystal growth by oriented particle attachment, by stabilizing the nano-building blocks, and guiding and orienting their assembly (Cuif and Dauphin 2005; Oaki et al. 2006; Nouet et al. 2012). Hydrophilic organic polymers are particularly effective in this process and induce the formation of mesocrystals of carbonates, phosphates and metal oxides in laboratory precipitation experiments (Mukkamala and Powell 2004; Yu and Cölfen 2004; Yu et al. 2005; Meldrum and Cölfen 2008) (Fig. 6a), but simple inorganic additives such as Mg2+ can also lead to mesocrystal-like polycrystalline structures (Liu et al. 2020) (Fig. 6c). Magnetite mesocrystals were also obtained through the hydrolysis of organically bound ferric iron (Cai et al. 2014).

Fig. 6.

Biomorphic nanocrystalline structures and mesocrystals. (a) Nanocrystalline calcite ‘microtrumpet’ formed in the presence of a carboxylic-rich organic molecule (Mukkamala and Powell 2004). (b) Co-oriented carbonate nanorods at the surface of a carbonate–silica biomorph (Kellermeier et al. 2012a). (c) Nanocrystalline calcite sphere precipitated in the presence of Mg2+ (Liu et al. 2020). (d) Fe oxide mesocrystal formed by droplet evaporation (Agthe et al. 2014). Images reproduced with permission (parts a–c) or under creative commons license (creativecommons.org/licenses/by/3.0/) (part d).

Fig. 6.

Biomorphic nanocrystalline structures and mesocrystals. (a) Nanocrystalline calcite ‘microtrumpet’ formed in the presence of a carboxylic-rich organic molecule (Mukkamala and Powell 2004). (b) Co-oriented carbonate nanorods at the surface of a carbonate–silica biomorph (Kellermeier et al. 2012a). (c) Nanocrystalline calcite sphere precipitated in the presence of Mg2+ (Liu et al. 2020). (d) Fe oxide mesocrystal formed by droplet evaporation (Agthe et al. 2014). Images reproduced with permission (parts a–c) or under creative commons license (creativecommons.org/licenses/by/3.0/) (part d).

Organization of nanocrystallites at the mesoscale can also emerge from inorganic self-assembly. Carbonate–silica biomorphs are, for instance, composed of carbonate nanorods, uniform in shape and size, that are almost parallel to each other and aligned with respect to their c-axis (the slight misalignment between the nanorods occurs at fixed angles and allows for the radial growth of the sheaths composing the biomorphs) (Kellermeier et al. 2012c) (Fig. 6b). Mesocrystals can also be obtained inorganically by crystallization within silica gels (Dominguez Bella and García-Ruiz 1987) or through hydrothermal processes or solvent evaporation (Agthe et al. 2014; Sturm and Cölfen 2016; Sun et al. 2019) (Fig. 6d).

It has also been suggested that the very small sizes of microbial biominerals forming from highly supersaturated solutions as a result of microbial activity could be used as a biosignature (Banfield et al. 2001). Different microbial biomineralization processes can indeed result in the formation of intra- and extracellular iron, sulfur or carbonate mineral nanoparticles (i.e. with at least one dimension <100 nm) (Mansor and Xu 2020). Crystallites in the nanoparticulate range can also form through nucleation and growth within an organic matrix, as observed in carbonate microbialites (Benzerara and Menguy 2009). However, because mineral nanoparticules can precipitate chemically from highly supersaturated fluids, or as a result of multiple nucleation events on organic or inorganic seeding materials, they do not appear to be robust biosignatures.

Similar to the non-equilibrium mineral structures described in the previous section, signatures based on mineral textures will likely be difficult to interpret, especially in deposits where organics are abundant and/or in the absence of information on their original geochemical formation environment. However, Mars is more suitable than Earth for the long-term preservation of nanocrystalline mineral textures (e.g. the survival of opaline silica over multi-billion-year timescales at many localities on Mars) as a result of its low surface temperature and geothermal gradient and less prevalent fluid circulation. The possibility that such textures will be discovered is thus considerable and more effort is needed to better constrain their biogenic and abiogenic formation mechanisms in Mars-like contexts.

Chemical signatures in minerals

Aside from their structure, the chemical composition of biominerals may be different from that expected to form at equilibrium with the surrounding fluid geochemistry. Organisms can preferentially accumulate or exclude certain elements from their intracellular medium or internal compartments. As a result, intracellular biominerals may display enrichments of specific elements compared with the chemistry of the extracellular environment (e.g. Cam et al. 2016) or high levels of chemical purity (Amor et al. 2015). The PIXL instrument onboard the Perseverance rover can map the distribution and abundance variations of major and trace chemical elements at the sub-millimetre scale in rock and soil samples (Allwood et al. 2020). This technique may enable the detection of fine chemical biosignatures on Mars, such as enrichments of certain trace metals associated with organic matter (Hickman-Lewis et al. 2020).

However, a thorough understanding of the geochemical context for mineral formation, as well as diagenetic history, are necessary to interpret mineral chemistry in terms of biosignatures. For instance, the Mg–Ca carbonate dolomite forms in an environment associated with microbial mats (e.g. DiLoreto et al. 2019), although, unlike calcium carbonates, this mineral is very difficult to precipitate in the laboratory at low temperatures. The presence of dolomite in microbial mats is explained by organic templating effects and the interaction of Mg2+ with the carboxylic groups of extracellular polymeric substances (Krause et al. 2012; Roberts et al. 2013; Petrash et al. 2017). Dolomite can also form abiotically at high temperatures by precipitation from hydrothermal fluids or by the replacement of other primary carbonates during diagenesis or metamorphism (Bontognali 2019), or at low temperatures under certain geochemical conditions (e.g. in evaporative saline alkaline lakes; Wanas and Sallam 2016), preventing its use as a robust biosignature. More work is required from experiments under simulated martian conditions or from studies of analogue environments on Earth (e.g. Hays et al. 2017) to identify relevant and unambiguous mineralogical biosignatures that might be detected on Mars.

Neither the Rosalind Franklin nor Perseverance rover is equipped to measure the isotopic compositions of martian samples in situ, but samples returned to Earth will undoubtedly be analysed with isotope mass spectrometry in due course. Unidirectional, enzyme-catalysed metabolic reactions tend to select the lighter isotopes of carbon, sulfur, nitrogen, iron and many other elements because of their lower bond energies. This kinetic effect generates isotopically light products while leaving the substrates depleted in the light isotopes (i.e. relatively enriched in heavier isotopes. In some heavy elements this pattern is actually reversed for reasons to do with nuclear size and shape (e.g. uranium; Brown et al. 2018).

The interpretation of isotopic anomalies as biosignatures in geological materials (including fluids, minerals and organic matter) is a well-established practice on Earth, but relies critically on a thorough understanding of the history and context of the sample as well as the (bio)geochemical cycle of the relevant element – that is, the isotopic composition of the major reservoirs and the fractionations associated with the main fluxes between them (which, in the case of biotic processes, depend on the particular enzymes involved) – and changes in these values across geological time. As others have noted, such an understanding is not available for Mars, which limits (but does not completely eliminate) the potential for discovering definitive isotopic biosignatures on that planet (Summons et al. 2011; Vago et al. 2017).

More generally, false positive isotopic biosignatures can occur where the interpreter does not take sufficient account of the abiotic pathways of isotope fractionation involving either kinetic or equilibrium processes (van Zuilen et al. 2003; Anbar 2004; Thomazo et al. 2009; Brown et al. 2018). This broad caveat applies to both bulk measurements and to more sophisticated analyses of the isotopic structure of molecules (e.g. clumped or site-specific isotope compositions), although the latter hold great promise for astrobiological applications (Eiler et al. 2018). We briefly discuss how false isotopic biosignatures may arise in the carbon, sulfur, nitrogen and iron systems.

Carbon isotopes

Biological uptake of carbon causes a kinetic isotope fractionation such that biologically fixed carbon exhibits a relatively negative δ13C value (the 13C/12C ratio of a sample expressed as a per mil deviation relative to a standard). On Earth, organic matter in the marine sedimentary rock record is characterized by δ13C values 20–30‰ lower than carbonates of the same age. The persistence and ubiquity of this offset over c. 3.5 Gyr of global geological history is considered strong evidence for the antiquity of life (Schidlowski 2001).

However, attributing individual samples of ancient carbonaceous matter to biotic or abiotic carbon fixation on the basis of δ13C values alone can be problematic even on Earth. On the one hand, the biological fractionation effect can be small in some circumstances (e.g. Takai et al. 2008; Reeves and Fiebig 2020). On the other hand, kinetic isotope effects associated with several abiotic pathways yield CH4 and/or other organic compounds with highly negative δ13C values (Horita 2005). For example, hydrocarbons produced by experimental Fischer–Tropsch type synthesis (at 250°C and 325–350 bar using formic acid as a carbon source) are reportedly depleted in 13C by as much as 36‰ (McCollom and Seewald 2006; McCollom et al. 2010).

Some researchers have noted that ‘abiotic photosynthesis’ (CO2 reduction photo-catalysed on mineral surfaces, with water as an electron donor) may also have produced isotopically light organic matter on early Mars (Franz et al. 2020), although experimental work is needed to test this hypothesis and to explore the isotopic effects of abiotic organic synthesis more generally. Van Zuilen (2008) has pointed out two further challenges for the identification of carbon isotopic biosignatures on Mars. First, martian sedimentary rocks may preserve debris from carbonaceous chondrites, which contain reduced carbon phases (some of them macromolecular) with δ13C values of about −15 to −20‰, similar to biomass on Earth. This material may be difficult to discriminate from biological organic matter. Second, the carbon isotopic composition of CO2 in the martian atmosphere probably changed significantly with the decline of volcanic outgassing and the drawdown and loss of CO2 early in martian history. To establish the isotopic fractionation associated with the formation of any endogenous organic carbon in recovered samples, it will therefore be necessary to identify carbonate minerals representative of contemporaneous atmospheric CO2, which may be challenging.

More generally, the isotopic composition of martian carbon is poorly understood. Martian meteorites contain reduced magmatic carbon characterized by δ13C values of about −20‰ (i.e. about 15‰ lighter than magmatic carbon in basalts on Earth, for unclear reasons), as well as carbonates with widely varying positive δ13CPDB values, up to about +60‰ (Grady et al. 2004; Steele et al. 2016). Pyrolysis of aeolian fines by the SAM instrument suite onboard the Curiosity rover yielded CO2 with δ13C values between c. −6 and +20‰ (Leshin et al. 2013). Subsequently, Noachian–Hesperian siliciclastic sedimentary rocks analysed by the same instrument yielded CO2 with δ13C values ranging from −25 ± 20 to +56 ± 11‰ (Franz et al. 2020). Both datasets suggest the presence of multiple carbon-bearing phases, including carbonates and reduced carbon, but little is known about their age and identity.

Sulfur isotopes

Variations in sulfur isotope ratios have been investigated as potential biosignatures in martian meteorites (Farquhar et al. 2000b) and sedimentary rocks (Franz et al. 2017) based on the fact that some microbial metabolisms produce mass-dependent fractionations of the sulfur isotopes (Ono 2008). In particular, microbial sulfate reduction can produce sulfide significantly depleted in the heavier 34S (i.e. relatively enriched in 32S) compared with the substrate (sulfate), resulting in isotopic fractionations of several tens of δ34S units (‰) (Kaplan and Rittenberg 1964; Canfield and Teske 1996). High-temperature abiotic sulfate reduction also fractionates sulfur isotopes (up to 20‰ at 100°C; Thomazo et al. 2009), but can probably be ruled out for many samples obtainable from the martian near-surface, which have not been exposed to high temperatures.

Microbial sulfur disproportionation and sulfide oxidation also fractionate sulfur isotopes and it can be difficult to disentangle the effects of these different metabolisms in sediments (Canfield et al. 1998; Philippot et al. 2007; Sim et al. 2011; Pellerin et al. 2019), particularly after further sulfur-cycling by post-depositional and diagenetic processes (Fike et al. 2015). Isotopically light sulfide has been interpreted as evidence of microbial sulfur-cycling microorganisms in Archean rocks on Earth, often in combination with other biosignatures such as microfossils or stromatolites (e.g, Wacey et al. 2011; Bontognali et al. 2012; Marin-Carbonne et al. 2018). However, this interpretation requires ruling out contributions from abiogenic processes that can also fractionate sulfur isotopes. For instance, isotopic exchange with sulfate in hydrothermal contexts can endow sulfide with δ34S values similar to microbially reduced sulfide (Ohmoto and Lasaga 1982).

Atmospheric processes such as SO2 photolysis can lead to mass-independent fractionation of sulfur isotopes (Farquhar et al. 2001). Signatures of mass-independent fractionation are commonly recorded in Archean rocks (Farquhar et al. 2000a), but are rarely observed today on Earth because the creation and preservation of mass-independent fractionation signals is unlikely in an oxygen-rich atmosphere (Pavlov and Kasting 2002; Halevy 2013).

Sulfur in martian meteorites records clear isotopic evidence of atmospheric photochemical mass-independent fractionation effects (Farquhar et al. 2000b; Franz et al. 2014). The Curiosity rover's SAM instrument suite measured δ34S values ranging from −47 to +28‰ in samples collected in Gale Crater, some of which are ‘broadly consistent with' the effects of microbial sulfur-cycling (Franz et al. 2017). However, these isotopic compositions could also be explained by the effects of equilibrium fractionation in the impact-driven hydrothermal system of the crater, combined with atmospheric chemical and photochemical processes (Franz et al. 2017).

High-precision measurements of the abundances of all four sulfur isotopes (32S, 33S, 34S and 36S) introduce new Δ33S and Δ36S isotopic ratios, which, in combination with geological contextual analysis, may help to discriminate the effects of biological, hydrothermal and photochemical fractionation processes (Philippot et al. 2007; Ono 2008; Kamyshny et al. 2014; Aoyaa and Ueno 2018). Such techniques will no doubt be applied to returned samples. The study of sulfur-rich Mars-analogue environments suggest that combining δ34S, Δ33S and Δ36S measurements with steady-state models of sulfur-cycling may be a possible diagnostic tool for biosignature detection (Moreras-Marti et al. 2021). However, this method requires an in-depth understanding of the martian geological, geochemical and atmospheric context, which, despite significant recent advances (Franz et al. 2019), may still be out of reach in the near future.

Nitrogen isotopes

Nitrogen undergoes a highly complex biogeochemical cycle on Earth, in which several important metabolic pathways impart isotopic fractionations of different amplitudes and directions (reported as δ15N, with Earth's atmospheric N2 providing the usual standard). The δ15N signature of marine biomass (about +5‰ today) can be preserved in fossil organic matter and as mineral-bound NH4+ derived from organic decay, with minimal diagenetic modification (Boyd and Philippot 1998; Thomazo et al. 2009). Organic matter in Archean metasediments records a systematic variation in δ15N over geological time, probably reflecting changes in the preponderance of different microbial metabolisms under increasingly oxidizing conditions (Ader et al. 2016); there are some interpretative difficulties relating to metamorphic overprint, but these will not be an issue on Mars.

The martian atmosphere is c. 3% nitrogen today and nitrogen-bearing organic molecules of uncertain origin have already been found in martian meteorites (Koike et al. 2020); they may soon be encountered in returned samples. It will be challenging to interpret δ15N in this material without a broader understanding of Noachian–Hesperian nitrogen cycling, particularly in relation to the isotopic effects of atmospheric escape and declining volcanism over geological time; NH4+ in clay minerals may provide a contemporaneous standard (van Zuilen 2008).

The risk of isotopic false positives arising in this context is not yet well understood, but large δ15N fractionations are already known to result from several relevant abiotic processes. Nitrate fixed by lightning may be isotopically light (Stüeken 2016). Organic nitrogen produced experimentally from NH3 in Miller–Urey reactions with CH4 and H2 is reportedly enriched in 15N by 10–12‰, while plasma discharge in mixtures of N2, CH4, CO and H2 causes fractionation of up to 25‰ in the opposite direction (Kung et al. 1979; Thomazo et al. 2009; Kuga et al. 2014). Nitrogen isotope fractionation may also occur during Fischer–Tropsch type organic synthesis (Kung et al. 1979; Thomazo et al. 2009). Chondritic meteoritic debris may also be present in returned samples and can contain organic nitrogen with δ15N values ranging from +25 to +150‰ (Sephton et al. 2003).

Iron isotopes

Fe isotopes have been investigated as potential signatures of microbial Fe cycling (Johnson et al. 2008; Poitrasson 2015), including in the context of Mars (Anand et al. 2006; van Zuilen 2008), although candidates for Fe isotopic biosignatures have not yet been detected in martian materials.

It has been reported that magnetite biomineralized by the magnetotactic bacterium Magnetospirillum magneticum AMB-1 is strongly depleted in heavy Fe isotopes (Amor et al. 2016). Moreover, mass-independent fractionation in 57Fe was observed during magnetite biomineralization, but not in the even Fe isotopes (54Fe, 56Fe and 58Fe), implying a magnetic isotope effect. This signature is apparently not produced abiotically or by other Fe-metabolizing bacteria (Amor et al. 2016), but it remains to be seen whether it can be detected in the rock record.

The oxidation of aqueous Fe(II) precipitates solid Fe(III) phases that are typically slightly enriched in the heavy Fe isotopes regardless of whether Fe(II) oxidation was biologically mediated or not (e.g. Bullen et al. 2001; Balci et al. 2006), so it is generally admitted that microbial Fe oxidation does not leave any robust isotopic biosignature (e.g. van Zuilen 2008; Thomazo et al. 2009). However, microbial dissimilatory Fe reduction, which couples the reduction of Fe(III) to the oxidation of organic matter, leads to an enrichment of the Fe(II) product in the light Fe isotopes (e.g. Percak-Dennett et al. 2011). Under certain conditions (in particular, low pH), the Fe isotope fractionation induced by microbial dissimilatory Fe reduction may exceed in amplitude the fractionation associated with the abiotic reductive dissolution of Fe(III) (e.g. Johnson et al. 2008; Thomazo et al. 2009; Chanda et al. 2021).

On Earth, isotopic signals related to microbial dissimilatory Fe reduction are, for instance, thought to be recorded in some banded iron formations (Johnson et al. 2008). However, a number of processes may overprint or even erase biologically driven Fe isotopic signals, such as the preferential binding of heavy Fe isotopes by organic ligands in solution (e.g. Lotfi-Kalahroodi et al. 2021) or isotopic exchange between Fe minerals during diagenesis or metamorphism (Whitehouse and Fedo 2007; Hyslop et al. 2008). Here again, the use of Fe isotopes to search for traces of biological activity necessarily requires an in-depth understanding of the mineralogical and geochemical context during primary sediment deposition and subsequent history, and their interpretation should be conducted in conjunction with other mineralogical, chemical and isotopic indicators (Poitrasson 2015).

In writing this review, we have been struck by the following five points. First, abiotic processes can mimic not only morphological biosignatures, but also chemical/molecular, mineralogical, isotopic and textural biosignatures; a critical attitude is required in all cases and morphological data are not necessarily less reliable than other possible lines of evidence for life (cf. García-Ruiz et al. 2002). Indeed, independent categories of false biosignature may often coincide in the same sample and thus ‘multiple lines of evidence’ suggestive of biology can be found in some non-biological systems.

Second, many abiotic materials that resemble life are produced by the dissipation of energy in systems characterized by strong thermodynamic and/or chemical gradients and the presence of liquid water (e.g. the hydrothermal synthesis of complex organic matter, the hydrothermal fractionation of carbon isotopes; chemical gardens, carbon–sulfur biomorphs and carbonate–silica biomorphs). Although in many ways unsurprising, this has paradoxical implications. Environments conducive to the origin and maintenance of life may also, by their very nature, be conducive to the formation of false biosignatures. Therefore a cell-like, organic, ambiguously biogenic structure found in a tenuously habitable, energy-poor environment (e.g. a cold desert) may be more credible as a biosignature than a similar structure found in the midst of a reducing hydrothermal system because the former is more difficult to explain abiotically than the latter. It follows that the search for evidence of life should not focus only on environments where life is most likely to have arisen (cf. Longo and Damer 2020) because it is here that the evidence will be most ambiguous.

Third, silica precipitation plays a key part in the formation of a wide variety of false biosignatures (e.g. carbonate–silicate biomorphs, chemical gardens, cryogenic opal biomorphs and, potentially, the formation of pseudomicrobialites in splash zones). Siliceous materials might be favoured targets for analysis and sample return because silica can help to preserve organic matter and morphological fossils as a cement and entombing medium (e.g. McMahon et al. 2018). It will be especially important to consider how self-organization might have influenced the textural and morphological features of these samples.

Fourth, interactions between water and basaltic lava, although probably contributing to habitability on early Mars in various important ways, is also associated with a number of pseudobiosignatures (e.g. quench crystallites, alteration textures, the production of chemical garden ‘seed’ minerals and the production of alkaline, silica-rich fluids that can generate biomorphs). Neither the potential occurrence of biosignatures nor the potential occurrence of pseudobiosignatures should be overlooked in the sampling and analysis of aqueously altered volcanic materials (Ivarsson et al. 2020, 2021).

Fifth, the study of false biosignatures requires input from colleagues and methods across several disciplines. Physical chemists, condensed matter physicists and materials scientists have long been interested in symmetry-breaking, self-assembly and the emergence of complex structures and materials under far from equilibrium conditions. There is a vast literature – only a tiny sample of which has been cited in this review – on the numerical and experimental interrogation of these processes and the analytical characterization of their products. This expertise now needs to be translated into a geological and astrobiological context by focusing on the minerals, fluids and pressure–temperature conditions present on Mars today and in the geological past (e.g. Sainz-Díaz et al. 2021). Some important theoretical foundations have been laid (e.g. Ortoleva 1994; García-Ruiz et al. 2020), but there is more work to be done, both to identify relevant sources of false biosignatures and to characterize them in detail. Ideally, multiple high-resolution analytical techniques should be used to facilitate in-depth comparisons with candidate biosignatures. Recent studies of probable Precambrian fossils have presented rich, multi-scale, multi-proxy datasets combining morphological, textural and compositional (chemical, mineralogical and isotopic) information (e.g. Alleon et al. 2018; Lepot et al. 2019; Hickman-Lewis et al. 2020; Marin-Carbonne et al. 2020). Studies of false biosignatures should be similarly comprehensive.

What about biogenicity criteria?

A number of protocols and strategies have been devised to assess the credibility of candidate biosignatures (e.g. Buick 1990; Brasier and Wacey 2012; McLoughlin and Grosch 2015; Vago et al. 2017; Neveu et al. 2018; Rouillard et al. 2021). Most of these schemes use multiple, nested criteria to assess biogenicity: did the object (or population of objects) form in a demonstrably habitable (palaeo)environment, with appropriate evidence of endogeneity and syngenicity? If so, is its morphology consistent with a biotic origin and inconsistent with an abiotic origin? If so, is its chemical composition distinctively life-like? And so on. The more definitely and completely the object (including its geological context) meets the criteria, the higher the biogenicity score.

Biogenicity determination protocols meet an important need, but most real fossils (whether on Earth or Mars) are likely to fall short of ideal standards and achieve only middling scores, along with pseudofossils. Thus a failure to meet some of the relevant biogenicity criteria is expected and can often be explained away by researchers claiming to have discovered biosignatures. A deeper worry is that biogenicity criteria are unable to discriminate sensitively and reliably between biosignatures and pseudobiosignatures unless they are grounded in extensive knowledge and understanding of both classes of phenomena. Even protocols that do not explicitly rely on binary criteria must necessarily appeal to reference data from relevant abiotic systems (e.g. Rouillard et al. 2021). However, most known varieties of pseudobiosignature have not been characterized or understood in sufficient detail for this to be possible. Moreover, given the haphazard and unsystematic way in which varieties of false biosignature have so far been identified, we can only assume that many others remain undiscovered. Here, the expertise and methods of mineralogists, chemists, physicists and materials scientists are as valuable to biosignature science as those of microbiologists and palaeontologists.

Although the discovery of new forms of pseudobiosignature will tend to undermine existing biogenicity protocols (e.g. McMahon 2019; Nims et al. 2021), in the long run this will lead to better protocols (e.g. McLoughlin et al. 2007; Johannessen et al. 2020).

If we are lucky, plain and unequivocal biosignatures will be discovered on Mars in the coming decades. But in light of the many cautionary tales in the history of palaeontology and astrobiology, it seems prudent to anticipate more ambiguous results. In interpreting these, it will be important to understand how misleadingly life-like objects and substances might have formed abiotically on Mars. We have summarized some of the examples that have already been discovered in the laboratory, in terrestrial rocks and in meteorites. We have argued that many other examples probably await discovery because abiotic self-organization processes relevant to the geology of Mars have not been explored systematically. Because life itself is presumed to be the product of self-organization in abiotic geochemical reactions, the complexity of abiotic natural products should not be underestimated. Nevertheless, we are optimistic that the problem of false biosignatures is not intractable. The better these phenomena are understood, the more sensitively we will be able to discriminate between true evidence of life and these impostors.

Relevant false biosignatures should therefore now be sought systematically in Mars-analogue field and experimental systems and characterized in detail with multiple analytical instruments and at multiple scales to provide datasets that are as rich as those available for biosignatures. Further work on the physics and chemistry of far-from-equilibrium systems will ultimately reveal the limits of abiotic self-organization and may even lead to new discoveries on the organizing principles at the origin of life on Earth. Such efforts have an important part to play in the interpretation of Earth's rock record as well as the search for life on Mars and elsewhere in the solar system.

The perspectives in this article were shaped in part by SM's participation in the EU Cost Action CA17120 Chemobrionics. We thank an anonymous reviewer and Mark van Zuilen for helpful comments that improved this paper.

SM: conceptualization (lead), investigation (equal), writing – original draft (lead), writing – review & editing (equal); JC: conceptualization (supporting), investigation (equal), writing – original draft (supporting), writing – review & editing (equal).

This research received no specific grant from any funding agency in the public, commercial, or not-for-profit sectors.

Data sharing is not applicable to this article as no datasets were generated or analysed during the current study.

Scientific editing by Peter Grindrod

1.
Addadi
,
L.
and
Weiner
,
S
.
2014
.
Biomineralization: mineral formation by organisms
.
Physica Scripta
 ,
89
,
098003
, https://doi.org/10.1088/0031-8949/89/9/098003
2.
Addadi
,
L.
,
Raz
,
S.
and
Weiner
,
S
.
2003
.
Taking advantage of disorder: amorphous calcium carbonate and its roles in biomineralization
.
Advanced Materials
 ,
15
,
959
970
, https://doi.org/10.1002/adma.200300381
3.
Adelman
,
J
.
2007
.
Eozoön: debunking the dawn animal
.
Endeavour
 ,
31
,
94
98
, https://doi.org/10.1016/j.endeavour.2007.07.002
4.
Ader
,
M.
,
Thomazo
,
C.
et al.
2016
.
Interpretation of the nitrogen isotopic composition of Precambrian sedimentary rocks: assumptions and perspectives
.
Chemical Geology
 ,
429
,
93
110
, https://doi.org/10.1016/j.chemgeo.2016.02.010
5.
Agthe
,
M.
,
Wetterskog
,
E.
,
Mouzon
,
J.
,
Salazar-Alvarez
,
G.
and
Bergström
,
L
.
2014
.
Dynamic growth modes of ordered arrays and mesocrystals during drop-casting of iron oxide nanocubes
.
CrystEngComm
 ,
16
,
1443
1450
, https://doi.org/10.1039/C3CE41871E
6.
Alleon
,
J.
and
Summons
,
R.E
.
2019
.
Organic geochemical approaches to understanding early life
.
Free Radical Biology and Medicine
 ,
140
,
103
112
, https://doi.org/10.1016/j.freeradbiomed.2019.03.005
7.
Alleon
,
J.
,
Bernard
,
S.
,
Le Guillou
,
C.
,
Beyssac
,
O.
,
Sugitani
,
K.
and
Robert
,
F
.
2018
.
Chemical nature of the 3.4 Ga Strelley Pool microfossils
.
Geochemical Perspectives Letters
 ,
7
,
37
42
, https://doi.org/10.7185/geochemlet.1817
8.
Alleon
,
J.
,
Flannery
,
D.T.
,
Ferralis
,
N.
,
Williford
,
K.S.H.
,
Zhang
,
Y.
,
Schuessler
,
J.A.
and
Summons
,
R.E.
2019
.
Organo-mineral associations in chert of the 3.5 Ga Mount Ada Basalt raise questions about the origin of organic matter in Paleoarchean hydrothermally influenced sediments
.
Scientific Reports
 ,
9
,
16712
, https://doi.org/10.1038/s41598-019-53272-5
9.
Allwood
,
A.C.
,
Rosing
,
M.T.
,
Flannery
,
D.T.
,
Hurowitz
,
J.A.
and
Heirwegh
,
C.M
.
2018
.
Reassessing evidence of life in 3700-million-year-old rocks of Greenland
.
Nature
 ,
563
,
241
244
, https://doi.org/10.1038/s41586-018-0610-4
10.
Allwood
,
A.C.
,
Wade
,
L.A.
et al.
2020
.
PIXL: Planetary instrument for X-ray lithochemistry
.
Space Science Reviews
 ,
216
,
134
, https://doi.org/10.1007/s11214-020-00767-7
11.
Amor
,
M.
,
Busigny
,
V.
et al.
2015
.
Chemical signature of magnetotactic bacteria
.
Proceedings of the National Academy of Sciences of the United States of America
 ,
112
,
1699
1703
, https://doi.org/10.1073/pnas.1414112112
12.
Amor
,
M.
,
Busigny
,
V.
et al.
2016
.
Mass-dependent and -independent signature of Fe isotopes in magnetotactic bacteria
.
Science
 ,
352
,
705
708
, https://doi.org/10.1126/science.aad7632
13.
Amor
,
M.
,
Mathon
,
F.P.
,
Monteil
,
C.L.
,
Busigny
,
V.
and
Lefevre
,
C.T
.
2020
.
Iron-biomineralizing organelle in magnetotactic bacteria: function, synthesis and preservation in ancient rock samples
.
Environmental Microbiology
 ,
22
,
3611
3632
, https://doi.org/10.1111/1462-2920.15098
14.
Anand
,
M.
,
Russell
,
S.S.
,
Blackhurst
,
R.L.
and
Grady
,
M.M
.
2006
.
Searching for signatures of life on Mars: an Fe-isotope perspective
.
Philosophical Transactions of the Royal Society B: Biological Sciences
 ,
361
,
1715
1720
, https://doi.org/10.1098/rstb.2006.1899
15.
Anbar
,
A.D
.
2004
.
Iron stable isotopes: beyond biosignatures
.
Earth and Planetary Science Letters
 ,
217
,
223
236
, https://doi.org/10.1016/S0012-821X(03)00572-7
16.
Antcliffe
,
J.
and
McLoughlin
,
N.
2008
. Deciphering fossil evidence for the origin of life and the origin of animals.
Cellular Origin, Life in Extreme Habitats and Astrobiology
,
12
,
211
229
, https://doi.org/10.1007/978-1-4020-8837-7_10
17.
Aoyaa
,
S.
and
Ueno
,
Y
.
2018
.
Multiple sulfur isotope constraints on microbial sulfate reduction below an Archean seafloor hydrothermal system
.
Geobiology
 ,
16
,
107
120
, https://doi.org/10.1111/gbi.12268
18.
Balci
,
N.
,
Bullen
,
T.D.
,
Witte-Lien
,
K.
,
Shanks
,
W.C.
,
Motelica
,
M.
and
Mandernack
,
K.W
.
2006
.
Iron isotope fractionation during microbially stimulated Fe(II) oxidation and Fe(III) precipitation
.
Geochimica et Cosmochimica Acta
 ,
70
,
622
639
, https://doi.org/10.1016/j.gca.2005.09.025
19.
Banfield
,
J.F.
,
Moreau
,
J.W.
,
Chan
,
C.S.
,
Welch
,
S.A.
and
Little
,
B
.
2001
.
Mineralogical biosignatures and the search for life on Mars
.
Astrobiology
 ,
1
,
447
465
, https://doi.org/10.1089/153110701753593856
20.
Barge
,
L.M.
,
Cardoso
,
S.S.S.
et al.
2015
.
From chemical gardens to chemobrionics
.
Chemical Reviews
 ,
115
,
8652
8703
, https://doi.org/10.1021/acs.chemrev.5b00014
21.
Barghoorn
,
E.S.
and
Tyler
,
S.A
.
1963
.
Fossil organisms from Precambrian sediments
.
Annals of the New York Academy of Sciences
 ,
108
,
451
452
, https://doi.org/10.1111/j.1749-6632.1963.tb13399.x
22.
Barghoorn
,
E.S.
and
Tyler
,
S.A
.
1965
.
Microorganisms from the Gunflint Chert: these structurally preserved Precambrian fossils from Ontario are the most ancient organisms known
.
Science
 ,
147
,
563
575
, https://doi.org/10.1126/science.147.3658.563
23.
Barlow
,
E.V.
and
Kranendonk
,
M.J.V
.
2018
.
Snapshot of an early Paleoproterozoic ecosystem: two diverse microfossil communities from the Turee Creek Group, Western Australia
.
Geobiology
 ,
16
,
449
475
, https://doi.org/10.1111/gbi.12304
24.
Batchelor
,
M.T.
,
Burne
,
R.V.
,
Henry
,
B.I.
and
Jackson
,
M.J
.
2004
.
A case for biotic morphogenesis of coniform stromatolites
.
Physica A: Statistical Mechanics and its Applications
 ,
337
,
319
326
, https://doi.org/10.1016/j.physa.2004.01.065
25.
Baumgartner
,
R.J.
,
Kranendonk
,
M.J.V.
et al.
2019
.
Nano-porous pyrite and organic matter in 3.5-billion-year-old stromatolites record primordial life
.
Geology
 ,
47
,
1039
1043
, https://doi.org/10.1130/G46365.1
26.
Beegle
,
L.
,
Bhartia
,
R.
et al.
2015
.
SHERLOC: scanning habitable environments with Raman and luminescence for organics and chemicals
. Paper presented at the 2015 IEEE Aerospace Conference, Big Sky, MT, USA,
1
11
, https://doi.org/10.1109/AERO.2015.7119105
27.
Belilla
,
J.
,
Moreira
,
D.
et al.
2019
.
Hyperdiverse Archaea near life limits at the polyextreme geothermal Dallol area
.
Nature Ecology & Evolution
 ,
3
,
1552
1561
, https://doi.org/10.1038/s41559-019-1005-0
28.
Bell
,
J.F.
,
Maki
,
J.N.
et al.
2021
.
The Mars 2020 Perseverance rover mast camera zoom (Mastcam-Z) multispectral, stereoscopic imaging investigation
.
Space Science Reviews
 ,
217
,
24
, https://doi.org/10.1007/s11214-020-00755-x
29.
Bell
,
M.S
.
2007
.
Experimental shock decomposition of siderite and the origin of magnetite in Martian meteorite ALH 84001
.
Meteoritics & Planetary Science
 ,
42
,
935
949
, https://doi.org/10.1111/j.1945-5100.2007.tb01142.x
30.
Benzerara
,
K.
and
Menguy
,
N
.
2009
.
Looking for traces of life in minerals
.
Comptes Rendus Palevol
 ,
8
,
617
628
, https://doi.org/10.1016/j.crpv.2009.03.006
31.
Benzerara
,
K.
,
Menguy
,
N.
,
Guyot
,
F.
,
Skouri
,
F.
,
de Luca
,
G.
,
Barakat
,
M.
and
Heulin
,
T
.
2004
.
Biologically controlled precipitation of calcium phosphate by Ramlibacter tataouinensis
.
Earth and Planetary Science Letters
 ,
228
,
439
449
, https://doi.org/10.1016/j.epsl.2004.09.030
32.
Benzerara
,
K.
,
Meibom
,
A.
,
Gautier
,
Q.
,
Kazmierczak
,
J.
,
Stolarski
,
J.
,
Menguy
,
N.
and
Brown
,
G.E
.
2010
.
Nanotextures of aragonite in stromatolites from the quasi-marine Satonda crater lake, Indonesia
.
Geological Society, London, Special Publications
 ,
336
,
211
224
, https://doi.org/10.1144/SP336.10
33.
Benzerara
,
K.
,
Bernard
,
S.
and
Miot
,
J.
2019
. Mineralogical identification of traces of life.
Advances in Astrobiology and Biogeophysics
,
123
144
, https://doi.org/10.1007/978-3-319-96175-0_6
34.
Bergström
,
L.
,
Sturm née Rosseeva
,
E.V.
,
Salazar-Alvarez
,
G.
and
Cölfen
,
H.
2015
.
Mesocrystals in biominerals and colloidal arrays
.
Accounts of Chemical Research
 ,
48
,
1391
1402
, https://doi.org/10.1021/ar500440b
35.
Bernini
,
F.
,
Castellini
,
E.
et al.
2021
.
Self-assembled structures from solid cadmium(II) acetate in thiol/ethanol solutions: a novel type of organic chemical garden
.
ChemSystemsChem
 ,
3
,
e2000048
, https://doi.org/10.1002/syst.202000048
36.
Bibring
,
J.-P.
,
Hamm
,
V.
,
Pilorget
,
C.
,
Vago
,
J.L.
and
the MicrOmega Team
2017
.
The MicrOmega investigation onboard ExoMars
.
Astrobiology
 ,
17
,
621
626
, https://doi.org/10.1089/ast.2016.1642
37.
Bishop
,
J.L.
,
Dobrea
,
E.Z.N.
et al.
2008
.
Phyllosilicate diversity and past aqueous activity revealed at Mawrth Vallis, Mars
.
Science
 ,
321
,
830
833
, https://doi.org/10.1126/science.1159699
38.
Bishop
,
J.L.
,
Parente
,
M.
et al.
2009
.
Mineralogy of Juventae Chasma: sulfates in the light-toned mounds, mafic minerals in the bedrock, and hydrated silica and hydroxylated ferric sulfate on the plateau
.
Journal of Geophysical Research: Planets
 ,
114
, https://doi.org/10.1029/2009JE003352.
39.
Bonev
,
I.K.
,
García-Ruiz
,
J.M.
,
Atanassova
,
R.
,
Otalora
,
F.
and
Petrussenko
,
S
.
2005
.
Genesis of filamentary pyrite associated with calcite crystals
.
European Journal of Mineralogy
 ,
17
,
905
913
, https://doi.org/10.1127/0935-1221/2005/0017-0905
40.
Bontognali
,
T.R.R
.
2019
.
Anoxygenic phototrophs and the forgotten art of making dolomite
.
Geology
 ,
47
,
591
592
, https://doi.org/10.1130/focus062019.1
41.
Bontognali
,
T.R.R.
,
Sessions
,
A.L.
,
Allwood
,
A.C.
,
Fischer
,
W.W.
,
Grotzinger
,
J.P.
,
Summons
,
R.E.
and
Eiler
,
J.M
.
2012
.
Sulfur isotopes of organic matter preserved in 3.45-billion-year-old stromatolites reveal microbial metabolism
.
Proceedings of the National Academy of Sciences of the United States of America
 ,
109
,
15146
15151
, https://doi.org/10.1073/pnas.1207491109
42.
Botta
,
O.
and
Bada
,
J.L
.
2002
.
Extraterrestrial organic compounds in meteorites
.
Surveys in Geophysics
 ,
23
,
411
467
, https://doi.org/10.1023/A:1020139302770
43.
Bowerbank
,
J.S
.
1842
.
IV.—On the spongeous origin of moss agates and other siliceous bodies
.
Annals and Magazine of Natural History
 ,
10
,
9
18
, https://doi.org/10.1080/03745484209445187
44.
Boyd
,
S.R.
and
Philippot
,
P
.
1998
.
Precambrian ammonium biogeochemistry: a study of the Moine metasediments, Scotland
.
Chemical Geology
 ,
144
,
257
268
, https://doi.org/10.1016/S0009-2541(97)00135-6
45.
Brack
,
A
.
2007
.
From interstellar amino acids to prebiotic catalytic peptides: a review
.
Chemistry & Biodiversity
 ,
4
,
665
679
, https://doi.org/10.1002/cbdv.200790057
46.
Bralower
,
T.J.
,
Cosmidis
,
J.
et al.
2020
.
Origin of a global carbonate layer deposited in the aftermath of the Cretaceous–Paleogene boundary impact
.
Earth and Planetary Science Letters
 ,
548
,
116476
, https://doi.org/10.1016/j.epsl.2020.116476
47.
Brasier
,
A.T.
,
Culwick
,
T.
,
Battison
,
L.
,
Callow
,
R.H.T.
and
Brasier
,
M.D
.
2017
.
Evaluating evidence from the Torridonian Supergroup (Scotland, UK) for eukaryotic life on land in the Proterozoic
.
Geological Society, London, Special Publications
 ,
448
,
121
144
, https://doi.org/10.1144/SP448.13
48.
Brasier
,
A.T.
,
Dennis
,
P.F.
et al.
2019
.
Detecting ancient life: investigating the nature and origin of possible stromatolites and associated calcite from a one billion year old lake
.
Precambrian Research
 ,
328
,
309
320
, https://doi.org/10.1016/j.precamres.2019.04.025
49.
Brasier
,
M.D.
and
Wacey
,
D
.
2012
.
Fossils and astrobiology: new protocols for cell evolution in deep time
.
International Journal of Astrobiology
 ,
11
,
217
228
, https://doi.org/10.1017/S1473550412000298
50.
Brasier
,
M.D.
,
Green
,
O.R.
et al.
2002
.
Questioning the evidence for Earth's oldest fossils
.
Nature
 ,
416
,
76
81
, https://doi.org/10.1038/416076a
51.
Brasier
,
M.D.
,
Green
,
O.
,
Lindsay
,
J.
,
McLoughlin
,
N.
,
Steele
,
A.
and
Stoakes
,
C
.
2005
.
Critical testing of Earth's oldest putative fossil assemblage from the c. 3.5 Ga Apex chert, Chinaman Creek, Western Australia
.
Precambrian Research
 ,
140
,
55
102
, https://doi.org/10.1016/j.precamres.2005.06.008
52.
Bristow
,
T.F.
and
Milliken
,
R.E
.
2011
.
Terrestrial perspective on authigenic clay mineral production in ancient Martian lakes
.
Clays and Clay Minerals
 ,
59
,
339
358
, https://doi.org/10.1346/CCMN.2011.0590401
53.
Brooke
,
S.
and
Fox
,
S.W
.
1977
.
Compartmentalization in proteinoid microspheres
.
Biosystems
 ,
9
,
1
22
, https://doi.org/10.1016/0303-2647(77)90028-4
54.
Brown
,
A.J.
,
Viviano
,
C.E.
and
Goudge
,
T.A
.
2020
.
Olivine–carbonate mineralogy of the Jezero Crater region
.
Journal of Geophysical Research: Planets
 ,
125
,
e2019JE006011
, https://doi.org/10.1029/2019JE006011
55.
Brown
,
R.W
.
1957
.
Plantlike features in thunder-eggs and geodes
.
Annual Report of the Board of Regents of the Smithsonian Institution for 1956
,
329
339
.
56.
Brown
,
S.T.
,
Basu
,
A.
,
Ding
,
X.
,
Christensen
,
J.N.
and
DePaolo
,
D.J
.
2018
.
Uranium isotope fractionation by abiotic reductive precipitation
.
Proceedings of the National Academy of Sciences of the United States of America
 ,
115
,
8688
8693
, https://doi.org/10.1073/pnas.1805234115
57.
Buick
,
R
.
1990
.
Microfossil recognition in Archean rocks: an appraisal of spheroids and filaments from a 3500 m.y. old chert–barite unit at North Pole, Western Australia
.
PALAIOS
 ,
5
,
441
459
, https://doi.org/10.2307/3514837
58.
Bullen
,
T.D.
,
White
,
A.F.
,
Childs
,
C.W.
,
Vivit
,
D.V.
and
Schulz
,
M.S
.
2001
.
Demonstration of significant abiotic iron isotope fractionation in nature
.
Geology
 ,
29
,
699
702
, https://doi.org/10.1130/0091-7613(2001)029<0699:DOSAII>2.0.CO;2
59.
Busch
,
S.
,
Dolhaine
,
H.
et al.
1999
.
Biomimetic morphogenesis of fluorapatite–gelatin composites: fractal growth, the question of intrinsic electric fields, core/shell assemblies, hollow spheres and reorganization of denatured collagen
.
European Journal of Inorganic Chemistry
 ,
1999
,
1643
1653
, https://doi.org/10.1002/(SICI)1099-0682(199910)1999:10<1643::AID-EJIC1643>3.0.CO;2-J
60.
Burne
,
R.V.
and
Moore
,
L.S
.
1987
.
Microbialites; organosedimentary deposits of benthic microbial communities
.
PALAIOS
 ,
2
,
241
254
, https://doi.org/10.2307/3514674
61.
Cady
,
S.L.
,
Farmer
,
J.D.
,
Grotzinger
,
J.P.
,
Schopf
,
J.W.
and
Steele
,
A
.
2003
.
Morphological biosignatures and the search for life on Mars
.
Astrobiology
 ,
3
,
351
368
, https://doi.org/10.1089/153110703769016442
62.
Cai
,
J.
,
Chen
,
S.
,
Ji
,
M.
,
Hu
,
J.
,
Ma
,
Y.
and
Qi
,
L
.
2014
.
Organic additive-free synthesis of mesocrystalline hematite nanoplates via two-dimensional oriented attachment
.
CrystEngComm
 ,
16
,
1553
1559
, https://doi.org/10.1039/C3CE41716F
63.
Cam
,
N.
,
Benzerara
,
K.
et al.
2016
.
Selective uptake of alkaline earth metals by cyanobacteria forming intracellular carbonates
.
Environmental Science & Technology
 ,
50
,
11654
11662
, https://doi.org/10.1021/acs.est.6b02872
64.
Canfield
,
D.E.
and
Teske
,
A
.
1996
.
Late Proterozoic rise in atmospheric oxygen concentration inferred from phylogenetic and sulphur-isotope studies
.
Nature
 ,
382
,
127
132
, https://doi.org/10.1038/382127a0
65.
Canfield
,
D.E.
,
Thamdrup
,
B.
and
Fleischer
,
S
.
1998
.
Isotope fractionation and sulfur metabolism by pure and enrichment cultures of elemental sulfur-disproportionating bacteria
.
Limnology and Oceanography
 ,
43
,
253
264
, https://doi.org/10.4319/lo.1998.43.2.0253
66.
Chan
,
C.S.
,
Fakra
,
S.C.
,
Edwards
,
D.C.
,
Emerson
,
D.
and
Banfield
,
J.F
.
2009
.
Iron oxyhydroxide mineralization on microbial extracellular polysaccharides
.
Geochimica et Cosmochimica Acta
 ,
73
,
3807
3818
, https://doi.org/10.1016/j.gca.2009.02.036
67.
Chan
,
C.S.
,
Fakra
,
S.C.
,
Emerson
,
D.
,
Fleming
,
E.J.
and
Edwards
,
K.J
.
2011
.
Lithotrophic iron-oxidizing bacteria produce organic stalks to control mineral growth: implications for biosignature formation
.
The ISME Journal
 ,
5
,
717
727
, https://doi.org/10.1038/ismej.2010.173
68.
Chan
,
C.S.
,
McAllister
,
S.M.
,
Leavitt
,
A.H.
,
Glazer
,
B.T.
,
Krepski
,
S.T.
and
Emerson
,
D
.
2016
.
The architecture of iron microbial mats reflects the adaptation of chemolithotrophic iron oxidation in freshwater and marine environments
.
Frontiers in Microbiology
 ,
7
, https://doi.org/10.3389/fmicb.2016.00796
69.
Chan
,
M.A.
,
Hinman
,
N.W.
et al.
2019
.
Deciphering biosignatures in planetary contexts
.
Astrobiology
 ,
19
,
1075
1102
, https://doi.org/10.1089/ast.2018.1903
70.
Chanda
,
P.
,
Amenabar
,
M.J.
,
Boyd
,
E.S.
,
Beard
,
B.L.
and
Johnson
,
C.M
.
2021
.
Stable Fe isotope fractionation during dissimilatory Fe(III) reduction by a thermoacidophile in acidic hydrothermal environments
.
Geochimica et Cosmochimica Acta
 ,
292
,
427
451
, https://doi.org/10.1016/j.gca.2020.09.025
71.
Channing
,
A.
and
Butler
,
I.B
.
2007
.
Cryogenic opal-A deposition from Yellowstone hot springs
.
Earth and Planetary Science Letters
 ,
257
,
121
131
, https://doi.org/10.1016/j.epsl.2007.02.026
72.
Chyba
,
C.
and
Sagan
,
C
.
1992
.
Endogenous production, exogenous delivery and impact-shock synthesis of organic molecules: an inventory for the origins of life
.
Nature
 ,
355
,
125
132
, https://doi.org/10.1038/355125a0
73.
Cloud
,
P
.
1973
.
Pseudofossils: a plea for caution
.
Geology
 ,
1
,
123
127
, https://doi.org/10.1130/0091-7613(1973)1 < 123:PAPFC>2.0.CO;2
74.
Cloud
,
P.
1976
.
Beginnings of Biospheric Evolution and Their Biogeochemical Consequences
.
Paleobiology
 ,
2
,
351
387
.
75.
Coates
,
A.J.
,
Jaumann
,
R.
et al.
2017
.
The PanCam instrument for the ExoMars rover
.
Astrobiology
 ,
17
,
511
541
, https://doi.org/10.1089/ast.2016.1548
76.
Cockell
,
C.S.
and
Herrera
,
A
.
2008
.
Why are some microorganisms boring?
Trends in Microbiology
 ,
16
,
101
106
, https://doi.org/10.1016/j.tim.2007.12.007
77.
Cosmidis
,
J.
and
Templeton
,
A.S
.
2016
.
Self-assembly of biomorphic carbon/sulfur microstructures in sulfidic environments
.
Nature Communications
 ,
7
,
12812
, https://doi.org/10.1038/NCOMMS12812
78.
Cosmidis
,
J.
,
Benzerara
,
K.
,
Gheerbrant
,
E.
,
Estève
,
I.
,
Bouya
,
B.
and
Amaghzaz
,
M
.
2013
.
Nanometer-scale characterization of exceptionally preserved bacterial fossils in Paleocene phosphorites from Ouled Abdoun (Morocco)
.
Geobiology
 ,
11
,
139
153
, https://doi.org/10.1111/gbi.12022
79.
Cosmidis
,
J.
,
Benzerara
,
K.
et al.
2015
.
Calcium-phosphate biomineralization induced by alkaline phosphatase activity in Escherichia coli: localization, kinetics, and potential signatures in the fossil record
.
Frontiers in Earth Science
 ,
3
, https://doi.org/10.3389/feart.2015.00084
80.
Cosmidis
,
J.
,
Nims
,
C.W.
,
Diercks
,
D.
and
Templeton
,
A.S
.
2019
.
Formation and stabilization of elemental sulfur through organomineralization
.
Geochimica et Cosmochimica Acta
 ,
247
,
59
82
, https://doi.org/10.1016/j.gca.2018.12.025
81.
Couasnon
,
T.
,
Alloyeau
,
D.
,
Ménez
,
B.
,
Guyot
,
F.
,
Ghigo
,
J.-M.
and
Gélabert
,
A
.
2020
.
In situ monitoring of exopolymer-dependent Mn mineralization on bacterial surfaces
.
Science Advances
 ,
6
,
eaaz3125
, https://doi.org/10.1126/sciadv.aaz3125
82.
Criouet
,
I.
,
Viennet
,
J.-C.
,
Jacquemot
,
P.
,
Jaber
,
M.
and
Bernard
,
S
.
2021
.
Abiotic formation of organic biomorphs under diagenetic conditions
.
Geochemical Perspectives Letters
 ,
16
,
40
46
, https://doi.org/10.7185/geochemlet.2102
83.
Cron
,
B.
,
Henri
,
P.
,
Chan
,
C.S.
,
Macalady
,
J.L.
and
Cosmidis
,
J
.
2019
.
Elemental sulfur formation by Sulfuricurvum kujiense is mediated by extracellular organic compounds
.
Frontiers in Microbiology
 ,
10
,
2710
, https://doi.org/10.3389/fmicb.2019.02710
84.
Crosby
,
C.H.
and
Bailey
,
J.V
.
2017
.
Experimental precipitation of apatite pseudofossils resembling fossil embryos
.
Geobiology
 ,
16
,
80
87
,https://doi.org/10.1111/gbi.12264
85.
Cuif
,
J.-P.
and
Dauphin
,
Y
.
2005
.
The two-step mode of growth in the scleractinian coral skeletons from the micrometre to the overall scale
.
Journal of Structural Biology
 ,
150
,
319
331
, https://doi.org/10.1016/j.jsb.2005.03.004
86.
Dahl
,
C.
2020
. Bacterial intracellular sulphur globules.
Microbiology Monographs
,
34
,
19
51
, https://doi.org/10.1007/978-3-030-60173-7_2
87.
Dalai
,
P.
,
Kaddour
,
H.
and
Sahai
,
N
.
2016
.
Incubating life: prebiotic sources of organics for the origin of life
.
Elements
 ,
12
,
401
406
, https://doi.org/10.2113/gselements.12.6.401
88.
Daubenton
,
L.J.M
.
1782
.
Sur les causes qui produisent trois sortes d'herborisations dans les pierres
.
Mémoires de l'Académie Royale des Sciences
 ,
667
673
.
89.
Davies
,
N.S.
,
Liu
,
A.G.
,
Gibling
,
M.R.
and
Miller
,
R.F
.
2016
.
Resolving MISS conceptions and misconceptions: a geological approach to sedimentary surface textures generated by microbial and abiotic processes
.
Earth-Science Reviews
 ,
154
,
210
246
, https://doi.org/10.1016/j.earscirev.2016.01.005
90.
De Sanctis
,
M.C.
,
Altieri
,
F.
et al.
2017
.
Ma_MISS on ExoMars: mineralogical characterization of the martian subsurface
.
Astrobiology
 ,
17
,
612
620
, https://doi.org/10.1089/ast.2016.1541
91.
De Yoreo
,
J.J.
,
Gilbert
,
P.U.P.A.
et al.
2015
.
Crystallization by particle attachment in synthetic, biogenic, and geologic environments
.
Science
 ,
349
,
aaa6760
, https://doi.org/10.1126/science.aaa6760
92.
Deamer
,
D
.
2017
.
The role of lipid membranes in life's origin
.
Life
 ,
7
,
5
, https://doi.org/10.3390/life7010005
93.
Deamer
,
D
.
2021
.
Where did life begin? Testing ideas in prebiotic analogue conditions
.
Life
 ,
11
,
134
, https://doi.org/10.3390/life11020134
94.
Défarge
,
C.
2011
. Organomineralization. In:
Reitner
,
J.
and
Thiel
,
V.
(eds)
Encyclopedia of Geobiology
 .
Springer
,
697
701
, https://doi.org/10.1007/978-1-4020-9212-1_159
95.
Dehouck
,
E.
,
Chevrier
,
V.
,
Gaudin
,
A.
,
Mangold
,
N.
,
Mathé
,
P.-E.
and
Rochette
,
P
.
2012
.
Evaluating the role of sulfide-weathering in the formation of sulfates or carbonates on Mars
.
Geochimica et Cosmochimica Acta
 ,
90
,
47
63
, https://doi.org/10.1016/j.gca.2012.04.057
96.
Des Marais
,
D.J.
,
Nuth
,
J.A.
et al.
2008
.
The NASA astrobiology roadmap
.
Astrobiology
 ,
8
,
715
730
, https://doi.org/10.1089/ast.2008.0819
97.
DiLoreto
,
Z.A.
,
Bontognali
,
T.R.R.
et al.
2019
.
Microbial community composition and dolomite formation in the hypersaline microbial mats of the Khor Al-Adaid sabkhas, Qatar
.
Extremophiles
 ,
23
,
201
218
, https://doi.org/10.1007/s00792-018-01074-4
98.
Dodd
,
M.S.
,
Papineau
,
D.
et al.
2017
.
Evidence for early life in Earth's oldest hydrothermal vent precipitates
.
Nature
 ,
543
,
60
64
, https://doi.org/10.1038/nature21377
99.
Dominguez Bella
,
S.
and
García-Ruiz
,
J.M.
1987
.
Banding structures in induced morphology crystal aggregates of CaCO3
.
Journal of Materials Science
 ,
22
,
3095
3102
, https://doi.org/10.1007/BF01161169
100.
Duarte-Neto
,
P.
,
Stošić
,
T.
,
Stošić
,
B.
,
Lessa
,
R.
and
Milošević
,
M.V
.
2014
.
Interplay of model ingredients affecting aggregate shape plasticity in diffusion-limited aggregation
.
Physical Review E
 ,
90
,
012312
, https://doi.org/10.1103/PhysRevE.90.012312
101.
Duda
,
J.-P.
,
Thiel
,
V.
, et al.
2018
.
Ideas and perspectives: hydrothermally driven redistribution and sequestration of early Archaean biomass – the “hydrothermal pump hypothesis”
.
Biogeosciences
 ,
15
,
1535
1548
, https://doi.org/10.5194/bg-15-1535-2018
102.
Dupraz
,
C.
,
Reid
,
R.P.
,
Braissant
,
O.
,
Decho
,
A.W.
,
Norman
,
R.S.
and
Visscher
,
P.T
.
2009
.
Processes of carbonate precipitation in modern microbial mats
.
Earth-Science Reviews
 ,
96
,
141
162
, https://doi.org/10.1016/j.earscirev.2008.10.005
103.
Ehlmann
,
B.L.
,
Mustard
,
J.F.
and
Murchie
,
S.L
.
2010
.
Geologic setting of serpentine deposits on Mars
.
Geophysical Research Letters
 ,
37
, https://doi.org/10.1029/2010GL042596
104.
Eigenbrode
,
J.L.
,
Summons
,
R.E.
et al.
2018
.
Organic matter preserved in 3-billion-year-old mudstones at Gale crater, Mars
.
Science
 ,
360
,
1096
1101
, https://doi.org/10.1126/science.aas9185
105.
Eiler
,
J.M.
,
Clog
,
M.
,
Lawson
,
M.
,
Lloyd
,
M.
,
Piasecki
,
A.
,
Ponton
,
C.
and
Xie
,
H
.
2018
.
The isotopic structures of geological organic compounds
.
Geological Society, London, Special Publications
 ,
468
,
53
81
, https://doi.org/10.1144/SP468.4
106.
Emerson
,
D.
and
Moyer
,
C.L
.
2002
.
Neutrophilic Fe-oxidizing bacteria are abundant at the Loihi Seamount hydrothermal vents and play a major role in Fe oxide deposition
.
Applied and Environmental Microbiology
 ,
68
,
3085
3093
, https://doi.org/10.1128/AEM.68.6.3085-3093.2002
107.
Engelhardt
,
W.V.
,
Arndt
,
J.
,
Fecker
,
B.
and
Pankau
,
H.G
.
1995
.
Suevite breccia from the Ries crater, Germany: origin, cooling history and devitrification of impact glasses
.
Meteoritics
 ,
30
,
279
293
, https://doi.org/10.1111/j.1945-5100.1995.tb01126.x
108.
Erastova
,
V.
,
Degiacomi
,
M.T.G.
,
Fraser
,
D.
and
Greenwell
,
H.C.
2017
.
Mineral surface chemistry control for origin of prebiotic peptides
.
Nature Communications
 ,
8
,
2033
, https://doi.org/10.1038/s41467-017-02248-y
109.
Falini
,
G.
,
Albeck
,
S.
,
Weiner
,
S.
and
Addadi
,
L
.
1996
.
Control of aragonite or calcite polymorphism by mollusk shell macromolecules
.
Science
 ,
271
,
67
69
, https://doi.org/10.1126/science.271.5245.67
110.
Farley
,
K.A.
,
Malespin
,
C.
et al.
2014
.
In situ radiometric and exposure age dating of the martian surface
.
Science
 ,
343
, https://doi.org/10.1126/science.1247166
111.
Farquhar
,
J.
,
Bao
,
H.
and
Thiemens
,
M.H
.
2000
a.
Atmospheric influence of Earth's earliest sulfur cycle
.
Science
 ,
289
,
756
759
, https://doi.org/10.1126/science.289.5480.756
112.
Farquhar
,
J.
,
Savarino
,
J.
,
Jackson
,
T.L.
and
Thiemens
,
M.H
.
2000
b.
Evidence of atmospheric sulphur in the martian regolith from sulphur isotopes in meteorites
.
Nature
 ,
404
,
50
52
, https://doi.org/10.1038/35003517
113.
Farquhar
,
J.
,
Savarino
,
J.
,
Airieau
,
S.
and
Thiemens
,
M.H
.
2001
.
Observation of wavelength-sensitive mass-independent sulfur isotope effects during SO2 photolysis: implications for the early atmosphere
.
Journal of Geophysical Research: Planets
 ,
106
,
32829
32839
, https://doi.org/10.1029/2000JE001437
114.
Fike
,
D.A.
,
Bradley
,
A.S.
and
Rose
,
C.V
.
2015
.
Rethinking the ancient sulfur cycle
.
Annual Review of Earth and Planetary Sciences
 ,
43
,
593
622
, https://doi.org/10.1146/annurev-earth-060313-054802
115.
Fischer
,
A.
,
Schmitz
,
M.
,
Aichmayer
,
B.
,
Fratzl
,
P.
and
Faivre
,
D
.
2011
.
Structural purity of magnetite nanoparticles in magnetotactic bacteria
.
Journal of the Royal Society Interface
 ,
8
,
1011
1018
, https://doi.org/10.1098/rsif.2010.0576
116.
Fisk
,
M.
and
McLoughlin
,
N
.
2013
.
Atlas of alteration textures in volcanic glass from the ocean basins
.
Geosphere
 ,
9
,
317
341
, https://doi.org/10.1130/GES00827.1
https://doi.org/10.1130/GES00827.1
117.
Fisk
,
M.R.
,
Giovannoni
,
S.J.
and
Thorseth
,
I.H
.
1998
.
Alteration of oceanic volcanic glass: textural evidence of microbial activity
.
Science
 ,
281
,
978
980
, https://doi.org/10.1126/science.281.5379.978
118.
Fornaro
,
T.
,
Steele
,
A.
and
Brucato
,
J.R
.
2018
.
Catalytic/protective properties of martian minerals and implications for possible origin of life on Mars
.
Life
 ,
8
, https://doi.org/10.3390/life8040056
119.
Fothergill
,
J.
,
Li
,
M.
,
Davis
,
S.A.
,
Cunningham
,
J.A.
and
Mann
,
S
.
2014
.
Nanoparticle-based membrane assembly and silicification in coacervate microdroplets as a route to complex colloidosomes
.
Langmuir
 ,
30
,
14591
14596
, https://doi.org/10.1021/la503746u
120.
Fox
,
S.W.
and
Yuyama
,
S
.
1963
.
Abiotic production of primitive protein and formed microparticles
.
Annals of the New York Academy of Sciences
 ,
108
,
487
494
, https://doi.org/10.1111/j.1749-6632.1963.tb13404.x
121.
Fox-Powell
,
M.G.
and
Cousins
,
C.R.
2021
.
Partitioning of crystalline and amorphous phases during freezing of simulated Enceladus ocean fluids
.
Journal of Geophysical Research: Planets
 ,
126
,
e2020JE006628
, https://doi.org/10.1029/2020JE006628
122.
Francis
,
S.
,
Margulis
,
L.
and
Barghoorn
,
E.S
.
1978
.
On the experimental silicification of microorganisms II. On the time of appearance of eukaryotic organisms in the fossil record
.
Precambrian Research
 ,
6
,
65
100
, https://doi.org/10.1016/0301-9268(78)90055-4
123.
Franz
,
H.B.
,
Kim
,
S.-T.
et al.
2014
.
Isotopic links between atmospheric chemistry and the deep sulphur cycle on Mars
.
Nature
 ,
508
,
364
368
, https://doi.org/10.1038/nature13175
124.
Franz
,
H.B.
,
McAdam
,
A.C.
et al.
2017
.
Large sulfur isotope fractionations in Martian sediments at Gale crater
.
Nature Geoscience
 ,
10
,
658
662
, https://doi.org/10.1038/ngeo3002
125.
Franz
,
H.B.
,
King
,
P.L.
and
Gaillard
,
F.
2019
. Sulfur on Mars from the atmosphere to the core. In:
Filiberto
,
J.
and
Schwenzer
,
S.P.
(eds)
Volatiles in the Martian Crust
 .
Elsevier
,
119
183
, https://doi.org/10.1016/B978-0-12-804191-8.00006-4
126.
Franz
,
H.B.
,
Mahaffy
,
P.R.
et al.
2020
.
Indigenous and exogenous organics and surface–atmosphere cycling inferred from carbon and oxygen isotopes at Gale crater
.
Nature Astronomy
 ,
4
,
526
532
, https://doi.org/10.1038/s41550-019-0990-x
127.
Freissinet
,
C.
,
Glavin
,
D.P.
et al.
2015
.
Organic molecules in the Sheepbed Mudstone, Gale Crater, Mars
.
Journal of Geophysical Research: Planets
 ,
120
,
495
514
, https://doi.org/10.1002/2014JE004737
128.
Friedmann
,
E.I.
,
Wierzchos
,
J.
,
Ascaso
,
C.
and
Winklhofer
,
M
.
2001
.
Chains of magnetite crystals in the meteorite ALH84001: evidence of biological origin
.
Proceedings of the National Academy of Sciences of the United States of America
 ,
98
,
2176
2181
, https://doi.org/10.1073/pnas.051514698
129.
Fukushi
,
K.
,
Sekine
,
Y.
,
Sakuma
,
H.
,
Morida
,
K.
and
Wordsworth
,
R
.
2019
.
Semiarid climate and hyposaline lake on early Mars inferred from reconstructed water chemistry at Gale
.
Nature Communications
 ,
10
,
4896
, https://doi.org/10.1038/s41467-019-12871-6
130.
Furukawa
,
Y.
,
Chikaraishi
,
Y.
et al.
2019
.
Extraterrestrial ribose and other sugars in primitive meteorites
.
Proceedings of the National Academy of Sciences of the United States of America
 ,
116
,
24440
24445
, https://doi.org/10.1073/pnas.1907169116
131.
Gan
,
T.
,
Luo
,
T.
et al.
2021
.
Cryptic terrestrial fungus-like fossils of the early Ediacaran Period
.
Nature Communications
 ,
12
,
641
, https://doi.org/10.1038/s41467-021-20975-1
132.
García-Ruiz
,
J
.
1994
.
Inorganic self-organization in Precambrian charts
.
Origins of Life and Evolution of the Biosphere
 ,
24
,
451
467
, https://doi.org/10.1007/BF01582030
133.
García-Ruiz
,
J.M
.
2000
. Geochemical scenarios for the precipitation of biomimetic inorganic carbonates.
SEPM, Special Publications
,
67
,
75
89.
134.
García-Ruiz
,
J.M.
and
Amorós
,
J.L
.
1981
a.
Morphological aspects of some symmetrical crystal aggregates grown by silica gel technique
.
Journal of Crystal Growth
 ,
55
,
379
383
, https://doi.org/10.1016/0022-0248(81)90038-5
135.
García-Ruiz
,
J.M.
and
Amorós
,
J.L
.
1981
b.
Crystal aggregates with induced morphologies grown by silica gel technique
.
Bulletin de Minéralogie
 ,
104
,
107
113
, https://doi.org/10.3406/bulmi.1981.7442
136.
García-Ruiz
,
J.M.
,
Carnerup
,
A.
,
Christy
,
A.G.
,
Welham
,
N.J.
and
Hyde
,
S.T
.
2002
.
Morphology: an ambiguous indicator of biogenicity
.
Astrobiology
 ,
2
,
353
369
, https://doi.org/10.1089/153110702762027925
137.
García-Ruiz
,
J.M.
,
Melero-Garcia
,
E.
and
Hyde
,
S.T
.
2009
.
Morphogenesis of self-assembled nanocrystalline materials of barium carbonate and silica
.
Science
 ,
323
,
362
365
, https://doi.org/10.1126/science.1165349
138.
García-Ruiz
,
J.M.
,
Nakouzi
,
E.
,
Kotopoulou
,
E.
,
Tamborrino
,
L.
and
Steinbock
,
O
.
2017
.
Biomimetic mineral self-organization from silica-rich spring waters
.
Science Advances
 ,
3
,
e1602285
, https://doi.org/10.1126/sciadv.1602285
139.
García-Ruiz
,
J.M.
,
van Zuilen
,
M.A.
and
Bach
,
W
.
2020
.
Mineral self-organization on a lifeless planet
.
Physics of Life Reviews
 ,
34
–35,
62
82
, https://doi.org/10.1016/j.plrev.2020.01.001
140.
Garwood
,
R.J.
,
Oliver
,
H.
and
Spencer
,
A.R.T
.
2020
.
An introduction to the Rhynie chert
.
Geological Magazine
 ,
157
,
47
64
, https://doi.org/10.1017/S0016756819000670
141.
Georgieva
,
M.N.
,
Little
,
C.T.S.
,
Maslennikov
,
V.V.
,
Glover
,
A.G.
,
Ayupova
,
N.R.
and
Herrington
,
R.J
.
2021
.
The history of life at hydrothermal vents
.
Earth-Science Reviews
 ,
217
,
103602
, https://doi.org/10.1016/j.earscirev.2021.103602
142.
Gilbert
,
P.U.P.A.
,
Porter
,
S.M.
,
Sun
,
C.-Y.
,
Xiao
,
S.
,
Gibson
,
B.M.
,
Shenkar
,
N.
and
Knoll
,
A.H
.
2019
.
Biomineralization by particle attachment in early animals
.
Proceedings of the National Academy of Sciences of the United States of America
 ,
116
,
17659
17665
, https://doi.org/10.1073/pnas.1902273116
143.
Glauber
,
J.R.
1646
. LXXXV. Wie man in diesem Liquore von allen Metallen in wenig Stunden Bäume mit Farben soll wachsen machen. In:
Furni Novi Philosophici
 , T. Williams (London),
186
189
.
144.
Glavin
,
D.P.
,
Freissinet
,
C.
et al.
2013
.
Evidence for perchlorates and the origin of chlorinated hydrocarbons detected by SAM at the Rocknest aeolian deposit in Gale Crater
.
Journal of Geophysical Research: Planets
 ,
118
,
1955
1973
, https://doi.org/10.1002/jgre.20144
145.
Gleeson
,
D.F.
,
Pappalardo
,
R.T.
,
Anderson
,
M.S.
,
Grasby
,
S.E.
,
Mielke
,
R.E.
,
Wright
,
K.E.
and
Templeton
,
A.S
.
2012
.
Biosignature detection at an Arctic analog to Europa
.
Astrobiology
 ,
12
,
135
150
, https://doi.org/10.1089/ast.2010.0579
146.
Goesmann
,
F.
,
Brinckerhoff
,
W.B.
et al.
2017
.
The Mars organic molecule analyzer (MOMA) instrument: characterization of organic material in martian sediments
.
Astrobiology
 ,
17
,
655
685
, https://doi.org/10.1089/ast.2016.1551
147.
Golden
,
D.C.
,
Ming
,
D.W.
et al.
2001
.
A simple inorganic process for formation of carbonates, magnetite, and sulfides in Martian meteorite ALH84001
.
American Mineralogist
 ,
86
,
370
375
, https://doi.org/10.2138/am-2001-2-321
148.
Golden
,
D.C.
,
Ming
,
D.W.
et al.
2004
.
Evidence for exclusively inorganic formation of magnetite in Martian meteorite ALH84001
.
American Mineralogist
 ,
89
,
681
695
, https://doi.org/10.2138/am-2004-5-602
149.
Göppert
,
J.H.R
.
1848
. Ü
ber pflanzenähnliche Einschlüsse in den Chalcedonen
.
Flora
 ,
16
,
257
266
.
150.
Götze
,
J.
,
Hofmann
,
B.
et al.
2020
.
Biosignatures in subsurface filamentous fabrics (SFF) from the Deccan Volcanic Province, India
.
Minerals
 ,
10
,
540
, https://doi.org/10.3390/min10060540
151.
Gower
,
L.B
.
2008
.
Biomimetic model systems for investigating the amorphous precursor pathway and its role in biomineralization
.
Chemical Reviews
 ,
108
,
4551
4627
, https://doi.org/10.1021/cr800443h
152.
Grady
,
M.M.
,
Verchovsky
,
A.B.
and
Wright
,
I.P
.
2004
.
Magmatic carbon in Martian meteorites: attempts to constrain the carbon cycle on Mars
.
International Journal of Astrobiology
 ,
3
,
117
124
, https://doi.org/10.1017/S1473550404002071
153.
Grasby
,
S.E
.
2003
.
Naturally precipitating vaterite (μ-CaCO3) spheres: unusual carbonates formed in an extreme environment
.
Geochimica et Cosmochimica Acta
 ,
67
,
1659
1666
, https://doi.org/10.1016/S0016-7037(02)01304-2
154.
Grosch
,
E.G.
and
McLoughlin
,
N
.
2014
.
Reassessing the biogenicity of Earth's oldest trace fossil with implications for biosignatures in the search for early life
.
Proceedings of the National Academy of Sciences of the United States of America
 ,
111
,
8380
8385
, https://doi.org/10.1073/pnas.1402565111
155.
Grosch
,
E.G.
and
McLoughlin
,
N
.
2015
.
Questioning the biogenicity of titanite mineral trace fossils in Archean pillow lavas
.
Proceedings of the National Academy of Sciences of the United States of America
 ,
112
,
E3090
E3091
, https://doi.org/10.1073/pnas.1506995112
156.
Grotzinger
,
J.P.
and
Rothman
,
D.H
.
1996
.
An abiotic model for stromatolite morphogenesis
.
Nature
 ,
383
,
423
425
, https://doi.org/10.1038/383423a0
157.
Gruner
,
J.W
.
1924
.
Contributions to the geology of the Mesabi range: with special reference to the magnetites of the iron-bearing formation west of Mesaba
.
Minnesota Geological Survey Bulletin
 ,
19
,
1
71
.
158.
Halevy
,
I
.
2013
.
Production, preservation, and biological processing of mass-independent sulfur isotope fractionation in the Archean surface environment
.
Proceedings of the National Academy of Sciences of the United States of America
 ,
110
,
17644
17649
, https://doi.org/10.1073/pnas.1213148110
159.
Han
,
Z.
,
Meng
,
R.
et al.
2017
.
Calcium carbonate precipitation by Synechocystis sp. PCC6803 at different Mg/Ca molar ratios under the laboratory condition
.
Carbonates and Evaporites
 ,
32
,
561
575
, https://doi.org/10.1007/s13146-016-0322-5
160.
Hawley
,
J.E
.
1926
.
An evaluation of the evidence of life in the Archean
.
The Journal of Geology
 ,
34
,
441
461
, https://doi.org/10.1086/623331
161.
Hays
,
L.E.
,
Graham
,
H.V.
et al.
2017
.
Biosignature preservation and detection in Mars analog environments
.
Astrobiology
 ,
17
,
363
400
, https://doi.org/10.1089/ast.2016.1627
162.
Heinz
,
J.
and
Schulze-Makuch
,
D
.
2020
.
Thiophenes on Mars: biotic or abiotic origin?
Astrobiology
 ,
20
,
552
561
, https://doi.org/10.1089/ast.2019.2139
163.
Hickman-Lewis
,
K.
,
Cavalazzi
,
B.
et al.
2020
.
Metallomics in deep time and the influence of ocean chemistry on the metabolic landscapes of Earth's earliest ecosystems
.
Scientific Reports
 ,
10
,
4965
, https://doi.org/10.1038/s41598-020-61774-w
164.
Hofmann
,
B.A.
and
Farmer
,
J.D
.
2000
.
Filamentous fabrics in low-temperature mineral assemblages: are they fossil biomarkers? Implications for the search for a subsurface fossil record on the early Earth and Mars
.
Planetary and Space Science
 ,
48
,
1077
1086
, https://doi.org/10.1016/S0032-0633(00)00081-7
165.
Hofmann
,
B.A.
,
Farmer
,
J.D.
,
von Blanckenburg
,
F.
and
Fallick
,
A.E
.
2008
.
Subsurface filamentous fabrics: an evaluation of origins based on morphological and geochemical criteria, with implications for exopaleontology
.
Astrobiology
 ,
8
,
87
117
, https://doi.org/10.1089/ast.2007.0130
166.
Hood
,
M.A.
,
Landfester
,
K.
and
Muñoz-Espí
,
R
.
2014
.
The role of residue acidity on the stabilization of vaterite by amino acids and oligopeptides
.
Crystal Growth & Design
 ,
14
,
1077
1085
, https://doi.org/10.1021/cg401580y
167.
Hopkinson
,
L.
,
Roberts
,
S.
,
Herrington
,
R.
and
Wilkinson
,
J
.
1998
.
Self-organization of submarine hydrothermal siliceous deposits: evidence from the TAG hydrothermal mound, 26°N Mid-Atlantic Ridge
.
Geology
 ,
26
,
347
350
, https://doi.org/10.1130/0091-7613(1998)026<0347:SOOSHS>2.3.CO;2
168.
Horgan
,
B.
and
Bell
,
J.F
.
2012
.
Widespread weathered glass on the surface of Mars
.
Geology
 ,
40
,
391
394
, https://doi.org/10.1130/G32755.1
169.
Horgan
,
B.H.N.
,
Anderson
,
R.B.
,
Dromart
,
G.
,
Amador
,
E.S.
and
Rice
,
M.S
.
2020
.
The mineral diversity of Jezero crater: evidence for possible lacustrine carbonates on Mars
.
Icarus
 ,
339
,
113526
, https://doi.org/10.1016/j.icarus.2019.113526
170.
Horita
,
J
.
2005
.
Some perspectives on isotope biosignatures for early life
.
Chemical Geology
 ,
218
,
171
186
, https://doi.org/10.1016/j.chemgeo.2005.01.017
171.
Horodyski
,
R.J
.
1981
.
Pseudomicrofossils and altered microfossils from a middle Proterozoic shale belt supergroup, Montana
.
Precambrian Research
 ,
16
,
143
154
, https://doi.org/10.1016/0301-9268(81)90009-7
172.
Hyslop
,
E.V.
,
Valley
,
J.W.
,
Johnson
,
C.M.
and
Beard
,
B.L
.
2008
.
The effects of metamorphism on O and Fe isotope compositions in the Biwabik Iron Formation, northern Minnesota
.
Contributions to Mineralogy and Petrology
 ,
155
,
313
328
, https://doi.org/10.1007/s00410-007-0244-2
173.
Ivarsson
,
M.
,
Drake
,
H.
,
Neubeck
,
A.
,
Sallstedt
,
T.
,
Bengtson
,
S.
,
Roberts
,
N.M.W.
and
Rasmussen
,
B
.
2020
.
The fossil record of igneous rock
.
Earth-Science Reviews
 ,
210
,
103342
, https://doi.org/10.1016/j.earscirev.2020.103342
174.
Ivarsson
,
M.
,
Drake
,
H.
,
Neubeck
,
A.
,
Snoeyenbos-West
,
O.
,
Belivanova
,
V.
and
Bengtson
,
S
.
2021
.
Introducing palaeolithobiology
.
GFF
 , https://doi.org/10.1080/11035897.2021.1895302
175.
Javaux
,
E.J.
and
Lepot
,
K
.
2018
.
The Paleoproterozoic fossil record: implications for the evolution of the biosphere during Earth's middle-age
.
Earth-Science Reviews
 ,
176
,
68
86
, https://doi.org/10.1016/j.earscirev.2017.10.001
176.
Javaux
,
E.J.
,
Marshall
,
C.P.
and
Bekker
,
A
.
2010
.
Organic-walled microfossils in 3.2-billion-year-old shallow-marine siliciclastic deposits
.
Nature
 ,
463
,
934
938
, https://doi.org/10.1038/nature08793
177.
Javaux
,
E.J.
,
Asael
,
D.
et al.
2013
.
Identifying early Earth microfossils in unsilicified sediments
. Abstract presented at the
EGU General Assembly Conference,
EGU2013-7748, Vienna, Austria
.
178.
Johannessen
,
K.C.
,
McLoughlin
,
N.
,
Vullum
,
P.E.
and
Thorseth
,
I.H
.
2020
.
On the biogenicity of Fe-oxyhydroxide filaments in silicified low-temperature hydrothermal deposits: implications for the identification of Fe-oxidizing bacteria in the rock record
.
Geobiology
 ,
18
,
31
53
, https://doi.org/10.1111/gbi.12363
179.
Johnson
,
C.M.
,
Beard
,
B.L.
and
Roden
,
E.E
.
2008
.
The iron isotope fingerprints of redox and biogeochemical cycling in modern and ancient Earth
.
Annual Review of Earth and Planetary Sciences
 ,
36
,
457
493
, https://doi.org/10.1146/annurev.earth.36.031207.124139
180.
Johnson
,
J.R.
,
Bell
,
J.F.
et al.
2007
.
Mineralogic constraints on sulfur-rich soils from Pancam spectra at Gusev crater, Mars
.
Geophysical Research Letters
 ,
34
, https://doi.org/10.1029/2007GL029894
181.
Jones
,
B
.
2017
.
Review of calcium carbonate polymorph precipitation in spring systems
.
Sedimentary Geology
 ,
353
,
64
75
, https://doi.org/10.1016/j.sedgeo.2017.03.006
182.
Jordan
,
D.
2008
.
Aspects of Landscape Evolution, Lineaments and Fault Zone Mineralisation in Southeast Ireland
 . PhD thesis, Trinity College Dublin.
183.
Jordan
,
S.F.
,
Rammu
,
H.
,
Zheludev
,
I.N.
,
Hartley
,
A.M.
,
Maréchal
,
A.
and
Lane
,
N
.
2019
.
Promotion of protocell self-assembly from mixed amphiphiles at the origin of life
.
Nature Ecology & Evolution
 ,
3
,
1705
1714
, https://doi.org/10.1038/s41559-019-1015-y
184.
Josset
,
J.-L.
,
Westall
,
F.
et al.
2017
.
The close-up imager onboard the ESA ExoMars rover: objectives, description, operations, and science validation activities
.
Astrobiology
 ,
17
,
595
611
, https://doi.org/10.1089/ast.2016.1546
185.
Kamyshny
,
A.
,
Druschel
,
G.
,
Mansaray
,
Z.F.
and
Farquhar
,
J
.
2014
.
Multiple sulfur isotopes fractionations associated with abiotic sulfur transformations in Yellowstone National Park geothermal springs
.
Geochemical Transactions
 ,
15
,
7
, https://doi.org/10.1186/1467-4866-15-7
186.
Kaplan
,
I.R.
and
Rittenberg
,
S.C
.
1964
.
Microbiological fractionation of sulphur isotopes
.
Journal of General Microbiology
 ,
34
,
195
212
, https://doi.org/10.1099/00221287-34-2-195
187.
Kellermeier
,
M.
,
Eiblmeier
,
J.
,
Melero-García
,
E.
,
Pretzl
,
M.
,
Fery
,
A.
and
Kunz
,
W
.
2012
a.
Evolution and control of complex curved form in simple inorganic precipitation systems
.
Crystal Growth & Design
 ,
12
,
3647
3655
, https://doi.org/10.1021/cg3004646
188.
Kellermeier
,
M.
,
Melero-García
,
E.
,
Kunz
,
W.
and
García-Ruiz
,
J.M
.
2012
b.
Local autocatalytic co-precipitation phenomena in self-assembled silica–carbonate materials
.
Journal of Colloid and Interface Science
 ,
380
,
1
7
, https://doi.org/10.1016/j.jcis.2012.05.009
189.
Kellermeier
,
M.
,
Cölfen
,
H.
and
García-Ruiz
,
J.M
.
2012
c.
Silica biomorphs: complex biomimetic hybrid materials from ‘sand and chalk’
.
European Journal of Inorganic Chemistry
 ,
2012
,
5123
5144
, https://doi.org/10.1002/ejic.201201029
190.
Kim
,
S.S.
,
Bargar
,
J.R.
et al.
2011
.
Searching for biosignatures using electron paramagnetic resonance (EPR) analysis of manganese oxides
.
Astrobiology
 ,
11
,
775
786
, https://doi.org/10.1089/ast.2011.0619
191.
King
,
P.L.
and
McLennan
,
S.M
.
2010
.
Sulfur on Mars
.
Elements
 ,
6
,
107
112
, https://doi.org/10.2113/gselements.6.2.107
192.
Knoll
,
A.H.
and
Barghoorn
,
E.S
.
1974
.
Ambient pyrite in Precambrian chert: new evidence and a theory
.
Proceedings of the National Academy of Sciences of the United States of America
 ,
71
,
2329
2331
, https://doi.org/10.1073/pnas.71.6.2329
193.
Koga
,
S.
,
Williams
,
D.S.
,
Perriman
,
A.W.
and
Mann
,
S
.
2011
.
Peptide–nucleotide microdroplets as a step towards a membrane-free protocell model
.
Nature Chemistry
 ,
3
,
720
724
, https://doi.org/10.1038/nchem.1110
194.
Koike
,
M.
,
Nakada
,
R.
,
Kajitani
,
I.
,
Usui
,
T.
,
Tamenori
,
Y.
,
Sugahara
,
H.
and
Kobayashi
,
A
.
2020
.
In-situ preservation of nitrogen-bearing organics in Noachian Martian carbonates
.
Nature Communications
 ,
11
,
1988
, https://doi.org/10.1038/s41467-020-15931-4
195.
Korablev
,
O.I.
,
Dobrolensky
,
Y.
et al.
2017
.
Infrared spectrometer for ExoMars: a mast-mounted instrument for the rover
.
Astrobiology
 ,
17
,
542
564
, https://doi.org/10.1089/ast.2016.1543
196.
Kotopoulou
,
E.
,
Lopez-Haro
,
M.
,
Gamez
,
J.J.C.
and
García-Ruiz
,
J.M.
2021
.
Nanoscale anatomy of iron–silica self-organized membranes: implications for prebiotic chemistry
.
Angewandte Chemie
 ,
133
,
1416
1422
, https://doi.org/10.1002/ange.202012059
197.
Krause
,
S.
,
Liebetrau
,
V.
,
Gorb
,
S.
,
Sánchez-Román
,
M.
,
McKenzie
,
J.A.
and
Treude
,
T
.
2012
.
Microbial nucleation of Mg-rich dolomite in exopolymeric substances under anoxic modern seawater salinity: new insight into an old enigma
.
Geology
 ,
40
,
587
590
, https://doi.org/10.1130/G32923.1
198.
Kuga
,
M.
,
Carrasco
,
N.
et al.
2014
.
Nitrogen isotopic fractionation during abiotic synthesis of organic solid particles
.
Earth and Planetary Science Letters
 ,
393
,
2
13
, https://doi.org/10.1016/j.epsl.2014.02.037
199.
Kung
,
C.-C.
,
Hayatsu
,
R.
,
Studier
,
M.H.
and
Clayton
,
R.N
.
1979
.
Nitrogen isotope fractionations in the Fischer–Tropsch synthesis and in the Miller–Urey reaction
.
Earth and Planetary Science Letters
 ,
46
,
141
146
, https://doi.org/10.1016/0012-821X(79)90072-4
200.
Lagaly
,
G.
,
Ogawa
,
M.
and
Dékány
,
I.
2013
. Clay mineral–organic interactions. Developments in Clay Science,
5
,
435
505
, https://doi.org/10.1016/B978-0-08-098258-8.00015-8
201.
Lai
,
J.C.-Y.
,
Pearce
,
B.K.D.
,
Pudritz
,
R.E.
and
Lee
,
D
.
2019
.
Meteoritic abundances of fatty acids and potential reaction pathways in planetesimals
.
Icarus
 ,
319
,
685
700
, https://doi.org/10.1016/j.icarus.2018.09.028
202.
Lambert
,
J.-F
.
2008
.
Adsorption and polymerization of amino acids on mineral surfaces: a review
.
Origins of Life and Evolution of Biospheres
 ,
38
,
211
242
, https://doi.org/10.1007/s11084-008-9128-3
203.
Leduc
,
S.
1911
.
The Mechanism of Life
 , Rebman (London).
204.
Lepot
,
K.
,
Philippot
,
P.
,
Benzerara
,
K.
and
Wang
,
G.-Y
.
2009
.
Garnet-filled trails associated with carbonaceous matter mimicking microbial filaments in Archean basalt
.
Geobiology
 ,
7
,
393
402
, https://doi.org/10.1111/j.1472-4669.2009.00208.x
205.
Lepot
,
K.
,
Benzerara
,
K.
and
Philippot
,
P
.
2011
.
Biogenic versus metamorphic origins of diverse microtubes in 2.7 Gyr old volcanic ashes: multi-scale investigations
.
Earth and Planetary Science Letters
 ,
312
,
37
47
, https://doi.org/10.1016/j.epsl.2011.10.016
206.
Lepot
,
K.
,
Williford
,
K.H.
et al.
2019
.
Extreme 13C-depletions and organic sulfur content argue for S-fueled anaerobic methane oxidation in 2.72 Ga old stromatolites
.
Geochimica et Cosmochimica Acta
 ,
244
,
522
547
, https://doi.org/10.1016/j.gca.2018.10.014
207.
Leshin
,
L.A.
,
Mahaffy
,
P.R.
et al.
2013
.
Volatile, isotope, and organic analysis of martian fines with the Mars Curiosity rover
.
Science
 ,
341
,
1238937
, https://doi.org/10.1126/science.1238937
208.
Li
,
J.
,
Benzerara
,
K.
,
Bernard
,
S.
and
Beyssac
,
O
.
2013
.
The link between biomineralization and fossilization of bacteria: insights from field and experimental studies
.
Chemical Geology
 ,
359
,
49
69
, https://doi.org/10.1016/j.chemgeo.2013.09.013
209.
Li
,
J.
,
Bernard
,
S.
,
Benzerara
,
K.
,
Beyssac
,
O.
,
Allard
,
T.
,
Cosmidis
,
J.
and
Moussou
,
J
.
2014
.
Impact of biomineralization on the preservation of microorganisms during fossilization: an experimental perspective
.
Earth and Planetary Science Letters
 ,
400
,
113
122
, https://doi.org/10.1016/j.epsl.2014.05.031
210.
Li
,
J.
,
Menguy
,
N.
et al.
2015
.
Crystal growth of bullet-shaped magnetite in magnetotactic bacteria of the Nitrospirae phylum
.
Journal of the Royal Society Interface
 ,
12
,
20141288
, https://doi.org/10.1098/rsif.2014.1288
211.
Li
,
M.
,
Huang
,
X.
,
Tang
,
T.-Y.D.
and
Mann
,
S
.
2014
.
Synthetic cellularity based on non-lipid micro-compartments and protocell models
.
Current Opinion in Chemical Biology
 ,
22
,
1
11
, https://doi.org/10.1016/j.cbpa.2014.05.018
212.
Liesegang
,
R.E.
1914
. Die achate. In:
d’ Achiardi
,
G.
,
Amberg
,
R.
et al.
(eds)
Silicate: Band II Erste Hälfte
 , Springer.
186
190
, https://doi.org/10.1007/978-3-642-49866-4_8
213.
Lindsay
,
J.F.
,
Brasier
,
M.D.
,
McLoughlin
,
N.
,
Green
,
O.R.
,
Fogel
,
M.
,
Steele
,
A.
and
Mertzman
,
S.A
.
2005
.
The problem of deep carbon—an Archean paradox
.
Precambrian Research
 ,
143
,
1
22
, https://doi.org/10.1016/j.precamres.2005.09.003
214.
Liu
,
Z.
,
Zhang
,
Z.
et al.
2020
.
Shape-preserving amorphous-to-crystalline transformation of CaCO3 revealed by in situ TEM
.
Proceedings of the National Academy of Sciences of the United States of America
 ,
117
,
3397
3404
, https://doi.org/10.1073/pnas.1914813117
215.
Lotfi-Kalahroodi
,
E.
,
Pierson-Wickmann
,
A.-C.
,
Rouxel
,
O.
,
Marsac
,
R.
,
Bouhnik-Le Coz
,
M.
,
Hanna
,
K.
and
Davranche
,
M
.
2021
.
More than redox, biological organic ligands control iron isotope fractionation in the riparian wetland
.
Scientific Reports
 ,
11
,
1933
, https://doi.org/10.1038/s41598-021-81494-z
216.
Longo
,
A.
and
Damer
,
B
.
2020
.
Factoring origin of life hypotheses into the search for life in the solar system and beyond
.
Life
 ,
10
,
52
, https://doi.org/10.3390/life10050052
217.
Lowenstam
,
H
.
1981
.
Minerals formed by organisms
.
Science
 ,
211
,
1126
1131
, https://doi.org/10.1126/science.7008198
218.
MacCulloch
,
J.
1814
.
XXIII. On vegetable remains preserved in chalcedony
.
Transactions of the Geological Society of London
 ,
S1-2
,
510
527
, https://doi.org/10.1144/transgsla.2.510
219.
Mänd
,
K.
,
Kirsimäe
,
K.
et al.
2018
.
Authigenesis of biomorphic apatite particles from Benguela upwelling zone sediments off Namibia: the role of organic matter in sedimentary apatite nucleation and growth
.
Geobiology
 ,
16
,
640
658
, https://doi.org/10.1111/gbi.12309
220.
Mansor
,
M.
and
Xu
,
J
.
2020
.
Benefits at the nanoscale: a review of nanoparticle-enabled processes favouring microbial growth and functionality
.
Environmental Microbiology
 ,
22
,
3633
3649
, https://doi.org/10.1111/1462-2920.15174
221.
Marin-Carbonne
,
J.
,
Remusat
,
L.
,
Sforna
,
M.C.
,
Thomazo
,
C.
,
Cartigny
,
P.
and
Philippot
,
P.
2018
.
Sulfur isotope's signal of nanopyrites enclosed in 2.7 Ga stromatolitic organic remains reveal microbial sulfate reduction
.
Geobiology
 ,
16
,
121
138
, https://doi.org/10.1111/gbi.12275
222.
Marin-Carbonne
,
J.
,
Busigny
,
V.
et al.
2020
.
In situ Fe and S isotope analyses in pyrite from the 3.2 Ga Mendon Formation (Barberton Greenstone Belt, South Africa): evidence for early microbial iron reduction
.
Geobiology
 ,
18
,
306
325
, https://doi.org/10.1111/gbi.12385
223.
Marnocha
,
C.L.
,
Sabanayagam
,
C.R.
et al.
2019
.
Insights into the mineralogy and surface chemistry of extracellular biogenic S0 globules produced by Chlorobaculum tepidum
.
Frontiers in Microbiology
 ,
10
,
271
, https://doi.org/10.3389/fmicb.2019.00271
224.
Marshall
,
C.P.
,
Emry
,
J.R.
and
Olcott Marshall
,
A
.
2011
.
Haematite pseudomicrofossils present in the 3.5-billion-year-old Apex Chert
.
Nature Geoscience
 ,
4
,
240
243
, https://doi.org/10.1038/ngeo1084
225.
Marshall
,
S.M.
,
Murray
,
A.R.G.
and
Cronin
,
L
.
2017
.
A probabilistic framework for identifying biosignatures using Pathway Complexity
.
Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences
 ,
375
,
20160342
, https://doi.org/10.1098/rsta.2016.0342
226.
Martel
,
J.
,
Young
,
D.
,
Peng
,
H.-H.
,
Wu
,
C.-Y.
and
Young
,
J.D
.
2012
.
Biomimetic properties of minerals and the search for life in the Martian meteorite ALH84001
.
Annual Review of Earth and Planetary Sciences
 ,
40
,
167
193
, https://doi.org/10.1146/annurev-earth-042711-105401
227.
Maurice
,
S.
,
Wiens
,
R.C.
et al.
2021
.
The SuperCam instrument suite on the Mars 2020 rover: science objectives and mast-unit description
.
Space Science Reviews
 ,
217
,
47
, https://doi.org/10.1007/s11214-021-00807-w
228.
McAdam
,
A.C.
,
Franz
,
H.B.
et al.
2014
.
Sulfur-bearing phases detected by evolved gas analysis of the Rocknest aeolian deposit, Gale Crater, Mars
.
Journal of Geophysical Research: Planets
 ,
119
,
373
393
, https://doi.org/10.1002/2013JE004518
229.
McCollom
,
T.M
.
2013
.
Laboratory simulations of abiotic hydrocarbon formation in Earth's deep subsurface
.
Reviews in Mineralogy and Geochemistry
 ,
75
,
467
494
, https://doi.org/10.2138/rmg.2013.75.15
230.
McCollom
,
T.M.
and
Donaldson
,
C
.
2019
.
Experimental constraints on abiotic formation of tubules and other proposed biological structures in subsurface volcanic glass
.
Astrobiology
 ,
19
,
53
63
, https://doi.org/10.1089/ast.2017.1811
231.
McCollom
,
T.M.
and
Seewald
,
J.S.
2006
.
Carbon isotope composition of organic compounds produced by abiotic synthesis under hydrothermal conditions
.
Earth and Planetary Science Letters
 ,
243
,
74
84
, https://doi.org/10.1016/j.epsl.2006.01.027
232.
McCollom
,
T.M.
,
Lollar
,
B.S.
,
Lacrampe-Couloume
,
G.
and
Seewald
,
J.S
.
2010
.
The influence of carbon source on abiotic organic synthesis and carbon isotope fractionation under hydrothermal conditions
.
Geochimica et Cosmochimica Acta
 ,
74
,
2717
2740
, https://doi.org/10.1016/j.gca.2010.02.008
233.
McKay
,
D.S.
,
Gibson
,
E.K.
et al.
1996
.
Search for past life on Mars: possible relic biogenic activity in martian meteorite ALH84001
.
Science
 ,
273
,
924
930
, https://doi.org/10.1126/science.273.5277.924
234.
McLennan
,
S.M.
,
Bell
,
J.F.
et al.
2005
.
Provenance and diagenesis of the evaporite-bearing Burns Formation, Meridiani Planum, Mars
.
Earth and Planetary Science Letters
 ,
240
,
95
121
, https://doi.org/10.1016/j.epsl.2005.09.041
235.
McLoughlin
,
N.
and
Grosch
,
E.G
.
2015
.
A hierarchical system for evaluating the biogenicity of metavolcanic- and ultramafic-hosted microalteration textures in the search for extraterrestrial life
.
Astrobiology
 ,
15
,
901
921
, https://doi.org/10.1089/ast.2014.1259
236.
McLoughlin
,
N.
,
Brasier
,
M.D.
,
Wacey
,
D.
,
Green
,
O.R.
and
Perry
,
R.S
.
2007
.
On biogenicity criteria for endolithic microborings on early Earth and beyond
.
Astrobiology
 ,
7
,
10
26
, https://doi.org/10.1089/ast.2006.0122
237.
McLoughlin
,
N.
,
Wilson
,
L.A.
and
Brasier
,
M.D
.
2008
.
Growth of synthetic stromatolites and wrinkle structures in the absence of microbes – implications for the early fossil record
.
Geobiology
 ,
6
,
95
105
, https://doi.org/10.1111/j.1472-4669.2007.00141.x
238.
McLoughlin
,
N.
,
Grosch
,
E.G.
,
Vullum
,
P.E.
,
Guagliardo
,
P.
,
Saunders
,
M.
and
Wacey
,
D.
2019
.
Critically testing olivine-hosted putative martian biosignatures in the Yamato 000593 meteorite—geobiological implications
.
Geobiology
 ,
17
,
691
707
, https://doi.org/10.1111/gbi.12361
239.
McLoughlin
,
N.
,
Wacey
,
D.
,
Phunguphungu
,
S.
,
Saunders
,
M.
and
Grosch
,
E.G
.
2020
.
Deconstructing Earth's oldest ichnofossil record from the Pilbara Craton, West Australia: implications for seeking life in the Archean subseafloor
.
Geobiology
 ,
18
,
525
543
, https://doi.org/10.1111/gbi.12399
240.
McMahon
,
S
.
2019
.
Earth's earliest and deepest purported fossils may be iron-mineralized chemical gardens
.
Proceedings of the Royal Society B: Biological Sciences
 ,
286
, https://doi.org/10.1098/rspb.2019.2410
241.
McMahon
,
S.
and
Ivarsson
,
M
.
2019
.
A new frontier for palaeobiology: Earth's vast deep biosphere
.
BioEssays
 ,
41
,
1900052
, https://doi.org/10.1002/bies.201900052
242.
McMahon
,
S.
,
van Smeerdijk Hood
,
A.
and
McIlroy
,
D.
2017
.
The origin and occurrence of subaqueous sedimentary cracks
.
Geological Society, London, Special Publications
 ,
448
,
285
309
, https://doi.org/10.1144/SP448.15
243.
McMahon
,
S.
,
Bosak
,
T.
et al.
2018
.
A field guide to finding fossils on Mars
.
Journal of Geophysical Research: Planets
 ,
123
,
1012
1040
, https://doi.org/10.1029/2017JE005478
244.
McMahon
,
S.
,
Ivarsson
,
M.
et al.
2021
.
Dubiofossils from a Mars-analogue subsurface palaeoenvironment: the limits of biogenicity criteria
.
Geobiology
 ,
19
,
473
488
, https://doi.org/10.1111/gbi.12445
245.
Meldrum
,
F.C.
and
Cölfen
,
H
.
2008
.
Controlling mineral morphologies and structures in biological and synthetic systems
.
Chemical Reviews
 ,
108
,
4332
4432
, https://doi.org/10.1021/cr8002856
246.
Michalski
,
J.R.
,
Onstott
,
T.C.
,
Mojzsis
,
S.J.
,
Mustard
,
J.
,
Chan
,
Q.H.S.
,
Niles
,
P.B.
and
Johnson
,
S.S
.
2018
.
The Martian subsurface as a potential window into the origin of life
.
Nature Geoscience
 ,
11
,
21
26
, https://doi.org/10.1038/s41561-017-0015-2
247.
Milesi
,
V.
,
Guyot
,
F.
et al.
2015
.
Formation of CO2, H2 and condensed carbon from siderite dissolution in the 200–300°C range and at 50 MPa
.
Geochimica et Cosmochimica Acta
 ,
154
,
201
211
, https://doi.org/10.1016/j.gca.2015.01.015
248.
Ming
,
D.W.
,
Archer
,
P.D.
et al.
2014
.
Volatile and organic compositions of sedimentary rocks in Yellowknife Bay, Gale Crater, Mars
.
Science
 ,
343
,
1245267
, https://doi.org/10.1126/science.1245267
249.
Mojarro
,
A.
,
Jin
,
L.
,
Szostak
,
J.W.
,
Head
,
J.W.
and
Zuber
,
M.T.
2021
.
In search of the RNA world on Mars
.
Geobiology
 ,
19
,
307
321
, https://doi.org/10.1111/gbi.12433
250.
Moreras-Marti
,
A.
,
Fox-Powell
,
M.
et al.
2021
.
Quadruple sulfur isotope biosignatures from terrestrial Mars analogue systems
.
Geochimica et Cosmochimica Acta
 ,
308
,
157
172
, https://doi.org/10.1016/j.gca.2021.06.007
251.
Muirhead
,
B.K.
,
Nicholas
,
A.K.
,
Umland
,
J.
,
Sutherland
,
O.
and
Vijendran
,
S
.
2020
.
Mars sample return campaign concept status
.
Acta Astronautica
 ,
176
,
131
138
, https://doi.org/10.1016/j.actaastro.2020.06.026
252.
Mukkamala
,
S.B.
and
Powell
,
A.K
.
2004
.
Biomimetic assembly of calcite microtrumpets: crystal tectonics in action
.
Chemical Communications
 ,
918
919
, https://doi.org/10.1039/B401754D
253.
Muscente
,
A.D.
,
Czaja
,
A.D.
,
Tuggle
,
J.
,
Winkler
,
C.
and
Xiao
,
S
.
2018
.
Manganese oxides resembling microbial fabrics and their implications for recognizing inorganically preserved microfossils
.
Astrobiology
 ,
18
,
249
258
, https://doi.org/10.1089/ast.2017.1699
254.
Myrgorodska
,
I.
,
Meinert
,
C.
,
Martins
,
Z.
,
d'Hendecourt
,
L.L.S.
and
Meierhenrich
,
U.J.
2015
.
Molecular chirality in meteorites and interstellar ices, and the chirality experiment on board the ESA cometary Rosetta mission
.
Angewandte Chemie International Edition
 ,
54
,
1402
1412
, https://doi.org/10.1002/anie.201409354
255.
Nakamura-Messenger
,
K.
,
Messenger
,
S.
,
Keller
,
L.P.
,
Clemett
,
S.J.
and
Zolensky
,
M.E
.
2006
.
Organic globules in the Tagish Lake meteorite: remnants of the protosolar disk
.
Science
 ,
314
,
1439
1442
, https://doi.org/10.1126/science.1132175
256.
Neveu
,
M.
,
Hays
,
L.E.
,
Voytek
,
M.A.
,
New
,
M.H.
and
Schulte
,
M.D
.
2018
.
The ladder of life detection
.
Astrobiology
 ,
18
,
1375
1402
, https://doi.org/10.1089/ast.2017.1773
257.
Niles
,
P.B.
,
Michalski
,
J.
,
Ming
,
D.W.
and
Golden
,
D.C
.
2017
.
Elevated olivine weathering rates and sulfate formation at cryogenic temperatures on Mars
.
Nature Communications
 ,
8
,
998
, https://doi.org/10.1038/s41467-017-01227-7
258.
Nilson
,
F.P.R
.
2002
.
Possible impact of a primordial oil slick on atmospheric and chemical evolution
.
Origins of Life and Evolution of the Biosphere
 ,
32
,
247
253
, https://doi.org/10.1023/A:1016577923630
259.
Nims
,
C.
,
Lafond
,
J.
,
Alleon
,
J.
,
Templeton
,
A.S.
and
Cosmidis
,
J
.
2021
.
Organic biomorphs may be better preserved than microorganisms in early Earth sediments
.
Geology
 ,
49
,
629
634
, https://doi.org/10.1130/G48152.1
260.
Noffke
,
N
.
2015
.
Ancient sedimentary structures in the <3.7 Ga Gillespie Lake Member, Mars, that resemble macroscopic morphology, spatial associations, and temporal succession in terrestrial microbialites
.
Astrobiology
 ,
15
,
169
192
, https://doi.org/10.1089/ast.2014.1218
261.
Noffke
,
N
.
2018
.
Comment on the paper by Davies
et al.
. ‘Resolving MISS conceptions and misconceptions: a geological approach to sedimentary surface textures generated by microbial and abiotic processes’ (Earth Science Reviews, 154 (2016), 210–246)
.
Earth-Science Reviews
 ,
176
,
373
383
, https://doi.org/10.1016/j.earscirev.2017.11.021
262.
Noffke
,
N.
,
Gerdes
,
G.
,
Klenke
,
T.
and
Krumbein
,
W.E
.
2001
.
Microbially induced sedimentary structures: a new category within the classification of primary sedimentary structures
.
Journal of Sedimentary Research
 ,
71
,
649
656
, https://doi.org/10.1306/2DC4095D-0E47-11D7-8643000102C1865D
263.
Noffke
,
N.
,
Eriksson
,
K.A.
,
Hazen
,
R.M.
and
Simpson
,
E.L
.
2006
.
A new window into Early Archean life: microbial mats in Earth's oldest siliciclastic tidal deposits (3.2 Ga Moodies Group, South Africa)
.
Geology
 ,
34
,
253
256
, https://doi.org/10.1130/G22246.1
264.
Nouet
,
J.
,
Baronnet
,
A.
and
Howard
,
L
.
2012
.
Crystallization in organo-mineral micro-domains in the crossed-lamellar layer of Nerita undata (Gastropoda, Neritopsina)
.
Micron
 ,
43
,
456
462
, https://doi.org/10.1016/j.micron.2011.10.027
265.
Nutman
,
A.P.
,
Bennett
,
V.C.
,
Friend
,
C.R.L.
,
Van Kranendonk
,
M.J.
and
Chivas
,
A.R
.
2016
.
Rapid emergence of life shown by discovery of 3700-million-year-old microbial structures
.
Nature
 ,
537
,
535
538
, https://doi.org/10.1038/nature19355
266.
Nutman
,
A.P.
,
Bennett
,
V.C.
,
Friend
,
C.R.L.
and
Van Kranendonk
,
M.J.
2021
.
In support of rare relict c. 3700 Ma stromatolites from Isua (Greenland)
.
Earth and Planetary Science Letters
 ,
562
,
116850
, https://doi.org/10.1016/j.epsl.2021.116850
267.
Oaki
,
Y.
and
Imai
,
H
.
2003
.
Experimental demonstration for the morphological evolution of crystals grown in gel media
.
Crystal Growth & Design
 ,
3
,
711
716
, https://doi.org/10.1021/cg034053e
268.
Oaki
,
Y.
,
Kotachi
,
A.
,
Miura
,
T.
and
Imai
,
H
.
2006
.
Bridged nanocrystals in biominerals and their biomimetics: classical yet modern crystal growth on the nanoscale
.
Advanced Functional Materials
 ,
16
,
1633
1639
, https://doi.org/10.1002/adfm.200600262
269.
Ohmoto
,
H.
and
Lasaga
,
A.C
.
1982
.
Kinetics of reactions between aqueous sulfates and sulfides in hydrothermal systems
.
Geochimica et Cosmochimica Acta
 ,
46
,
1727
1745
, https://doi.org/10.1016/0016-7037(82)90113-2
270.
Ojha
,
L.
,
Wilhelm
,
M.B.
et al.
2015
.
Spectral evidence for hydrated salts in recurring slope lineae on Mars
.
Nature Geoscience
 ,
8
,
829
832
, https://doi.org/10.1038/ngeo2546
271.
Olszta
,
M.J.
,
Cheng
,
X.
et al.
2007
.
Bone structure and formation: a new perspective
.
Materials Science and Engineering R: Reports
 ,
58
,
77
116
, https://doi.org/10.1016/j.mser.2007.05.001
272.
Ono
,
S.
2008
. Multiple-sulphur isotope biosignatures.
Space Sciences Series of ISSI
 ,
25
,
203
220
, https://doi.org/10.1007/978-0-387-77516-6_14
273.
Oparin
,
A.I.
1924
.
Proiskhozhdenie Zhizny (The Origin of Life)
 , Moskovski Rabochii (Moscow).
274.
Opel
,
J.
,
Wimmer
,
F.P.
,
Kellermeier
,
M.
and
Cölfen
,
H
.
2016
.
Functionalisation of silica–carbonate biomorphs
.
Nanoscale Horizons
 ,
1
,
144
149
, https://doi.org/10.1039/C5NH00094G
275.
Opel
,
J.
,
Kellermeier
,
M.
,
Sickinger
,
A.
,
Morales
,
J.
,
Cölfen
,
H.
and
García-Ruiz
,
J.-M
.
2018
.
Structural transition of inorganic silica–carbonate composites towards curved lifelike morphologies
.
Minerals
 ,
8
,
75
, https://doi.org/10.3390/min8020075
276.
Orgel
,
L.E
.
2004
.
Prebiotic chemistry and the origin of the RNA world
.
Critical Reviews in Biochemistry and Molecular Biology
 ,
39
,
99
123
, https://doi.org/10.1080/10409230490460765
277.
Ortoleva
,
P
.
1994
. Geochemical Self-Organization.
Oxford Monographs on Geology and Geophysics. Oxford University Press
 .
278.
Pan
,
L.
,
Carter
,
J.
et al.
2021
.
Voluminous silica precipitated from martian waters during late-stage aqueous alteration
.
The Planetary Science Journal
 ,
2
,
65
, https://doi.org/10.3847/PSJ/abe541
279.
Pavlov
,
A.A.
and
Kasting
,
J.F
.
2002
.
Mass-independent fractionation of sulfur isotopes in Archean sediments: strong evidence for an anoxic Archean atmosphere
.
Astrobiology
 ,
2
,
27
41
, https://doi.org/10.1089/153110702753621321
280.
Pedersen
,
L.-E.R.
,
McLoughlin
,
N.
,
Vullum
,
P.E.
and
Thorseth
,
I.H
.
2015
.
Abiotic and candidate biotic micro-alteration textures in subseafloor basaltic glass: a high-resolution in-situ textural and geochemical investigation
.
Chemical Geology
 ,
410
,
124
137
, https://doi.org/10.1016/j.chemgeo.2015.06.005
281.
Pellerin
,
A.
,
Antler
,
G.
et al.
2019
.
Large sulfur isotope fractionation by bacterial sulfide oxidation
.
Science Advances
 ,
5
,
eaaw1480
, https://doi.org/10.1126/sciadv.aaw1480
282.
Peng
,
X.
and
Jones
,
B
.
2013
.
Patterns of biomediated CaCO3 crystal bushes in hot spring deposits
.
Sedimentary Geology
 ,
294
,
105
117
, https://doi.org/10.1016/j.sedgeo.2013.05.009
283.
Percak-Dennett
,
E.M.
,
Beard
,
B.L.
,
Xu
,
H.
,
Konishi
,
H.
,
Johnson
,
C.M.
and
Roden
,
E.E
.
2011
.
Iron isotope fractionation during microbial dissimilatory iron oxide reduction in simulated Archaean seawater
.
Geobiology
 ,
9
,
205
220
, https://doi.org/10.1111/j.1472-4669.2011.00277.x
284.
Petrash
,
D.A.
,
Bialik
,
O.M.
,
Bontognali
,
T.R.R.
,
Vasconcelos
,
C.
,
Roberts
,
J.A.
,
McKenzie
,
J.A.
and
Konhauser
,
K.O
.
2017
.
Microbially catalyzed dolomite formation: from near-surface to burial
.
Earth-Science Reviews
 ,
171
,
558
582
, https://doi.org/10.1016/j.earscirev.2017.06.015
285.
Petroff
,
A.P.
,
Sim
,
M.S.
,
Maslov
,
A.
,
Krupenin
,
M.
,
Rothman
,
D.H.
and
Bosak
,
T
.
2010
.
Biophysical basis for the geometry of conical stromatolites
.
Proceedings of the National Academy of Sciences of the United States of America
 ,
107
,
9956
9961
, https://doi.org/10.1073/pnas.1001973107
286.
Philippot
,
P.
,
Zuilen
,
M.V.
,
Lepot
,
K.
,
Thomazo
,
C.
,
Farquhar
,
J.
and
Kranendonk
,
M.J.V
.
2007
.
Early Archaean microorganisms preferred elemental sulfur, not sulfate
.
Science
 ,
317
,
1534
1537
, https://doi.org/10.1126/science.1145861
287.
Phoenix
,
V.R.
,
Renaut
,
R.W.
,
Jones
,
B.
and
Ferris
,
F.G
.
2005
.
Bacterial S-layer preservation and rare arsenic–antimony–sulphide bioimmobilization in siliceous sediments from Champagne Pool hot spring, Waiotapu, New Zealand
.
Journal of the Geological Society, London
 ,
162
,
323
331
, https://doi.org/10.1144/0016-764903-058
288.
Picard
,
A.
,
Kappler
,
A.
,
Schmid
,
G.
,
Quaroni
,
L.
and
Obst
,
M
.
2015
.
Experimental diagenesis of organo-mineral structures formed by microaerophilic Fe(II)-oxidizing bacteria
.
Nature Communications
 ,
6
,
6277
, https://doi.org/10.1038/ncomms7277
289.
Pickersgill
,
A.E.
,
Sapers
,
H.M.
,
Lee
,
M.R.
,
Wildman
,
M.
,
Lindgren
,
P.
and
Hallis
,
L.
2021
.
Microtubules, trichites, and bioalteration in impact glasses
. Paper presented at the
Lunar and Planetary Science Conference
(virtual conference),
2035
.
290.
Poitrasson
,
F.
2015
. Iron isotopes. In:
Gargaud
,
M.
,
Irvine
,
W.M.
et al.
(eds)
Encyclopedia of Astrobiology
 .
Springer
,
1264
1268
, https://doi.org/10.1007/978-3-662-44185-5_811
291.
Preston
,
L.J.
,
Izawa
,
M.R.M.
and
Banerjee
,
N.R
.
2011
.
Infrared spectroscopic characterization of organic matter associated with microbial bioalteration textures in basaltic glass
.
Astrobiology
 ,
11
,
585
599
, https://doi.org/10.1089/ast.2010.0604
292.
Quantin-Nataf
,
C.
,
Carter
,
J.
et al.
2021
.
Oxia planum: the landing site for the ExoMars ‘Rosalind Franklin’ rover mission: geological context and prelanding interpretation
.
Astrobiology
 ,
21
,
345
366
, https://doi.org/10.1089/ast.2019.2191
293.
Rasmussen
,
B.
,
Muhling
,
J.R.
and
Fischer
,
W.W.
2021
.
Ancient oil as a source of carbonaceous matter in 1.88-billion-year-old Gunflint stromatolites and microfossils
.
Astrobiology
 ,
21
,
655
672
, https://doi.org/10.1089/ast.2020.2376
294.
Reeves
,
E.P.
and
Fiebig
,
J
.
2020
.
Abiotic synthesis of methane and organic compounds in Earth's lithosphere
.
Elements
 ,
16
,
25
31
, https://doi.org/10.2138/gselements.16.1.25
295.
Reitner
,
J.
2004
. Organomineralization: a clue to the understanding of meteorite-related ‘bacteria-shaped’ carbonate particles.
Cellular Origin, Life in Extreme Habitats and Astrobiology
,
6
,
195
212
, https://doi.org/10.1007/1-4020-2522-X_13
296.
Roberts
,
J.A.
,
Kenward
,
P.A.
,
Fowle
,
D.A.
,
Goldstein
,
R.H.
,
González
,
L.A.
and
Moore
,
D.S
.
2013
.
Surface chemistry allows for abiotic precipitation of dolomite at low temperature
.
Proceedings of the National Academy of Sciences of the United States of America
 ,
110
,
14540
14545
, https://doi.org/10.1073/pnas.1305403110
297.
Rodriguez-Navarro
,
C.
,
Jimenez-Lopez
,
C.
,
Rodriguez-Navarro
,
A.
,
Gonzalez-Muñoz
,
M.T.
and
Rodriguez-Gallego
,
M
.
2007
.
Bacterially mediated mineralization of vaterite
.
Geochimica et Cosmochimica Acta
 ,
71
,
1197
1213
, https://doi.org/10.1016/j.gca.2006.11.031
298.
Ross
,
C.S
.
1962
.
Microlites in glassy volcanic rocks
.
American Mineralogist
 ,
47
,
723
740
.
299.
Rouillard
,
J.
,
García-Ruiz
,
J.-M.
,
Gong
,
J.
and
van Zuilen
,
M.A
.
2018
.
A morphogram for silica–witherite biomorphs and its application to microfossil identification in the early Earth rock record
.
Geobiology
 ,
16
,
279
296
, https://doi.org/10.1111/gbi.12278
300.
Rouillard
,
J.
,
van Zuilen
,
M.
,
Pisapia
,
C.
and
Garcia-Ruiz
,
J.-M
.
2021
.
An alternative approach for assessing biogenicity
.
Astrobiology
 ,
21
,
151
164
, https://doi.org/10.1089/ast.2020.2282
301.
Ruff
,
S.W.
and
Farmer
,
J.D
.
2016
.
Silica deposits on Mars with features resembling hot spring biosignatures at El Tatio in Chile
.
Nature Communications
 ,
7
,
13554
, https://doi.org/10.1038/ncomms13554
302.
Ruff
,
S.W.
,
Campbell
,
K.A.
,
Van Kranendonk
,
M.J.
,
Rice
,
M.S.
and
Farmer
,
J.D
.
2020
.
The case for ancient hot springs in Gusev Crater, Mars
.
Astrobiology
 ,
20
,
475
499
, https://doi.org/10.1089/ast.2019.2044