Hostname: page-component-76fb5796d-45l2p Total loading time: 0 Render date: 2024-04-27T20:18:20.784Z Has data issue: false hasContentIssue false

Astrobiological applications of μ-mapping X-ray fluorescence spectrometry

Published online by Cambridge University Press:  04 March 2024

David Frederick Blake*
Affiliation:
Space Science Division, NASA Ames Research Center, Moffett Field, CA 94035, USA
Richard C. Walroth
Affiliation:
Space Science Division, NASA Ames Research Center, Moffett Field, CA 94035, USA
Thomas F. Bristow
Affiliation:
Space Science Division, NASA Ames Research Center, Moffett Field, CA 94035, USA
Philippe Sarrazin
Affiliation:
eXaminArt, LLC, Mountain View, CA, USA
Marc Gailhanou
Affiliation:
CNRS, IM2NP, Aix Marseille Université Institut Universitaire de Technologie de Marseille, Marseille, France
Kathleen Thompson
Affiliation:
SETI Institute, Mountain View, CA, USA
Robert Terrance Downs
Affiliation:
Department of Geological Sciences, University of Arizona, Tucson, AZ, USA
Albert Shi-Yuen Yen
Affiliation:
Jet Propulsion Laboratory, Earth and Space Sciences, NASA-JPL, Pasadena, CA, USA
Franck Marchis
Affiliation:
SETI Institute, Mountain View, CA, USA
Samuel Webb
Affiliation:
Stanford Synchrotron Radiation Facility, Stanford University, Stanford, CA, USA
Clement Chalumeau
Affiliation:
eXaminArt, LLC, Mountain View, CA, USA
Vincente Armando Solé
Affiliation:
European Synchrotron Radiation Facility, Grenoble, France
Philippe Walter
Affiliation:
Faculty of Science and Engineering, CNRS – Sorbonne Université, Paris, France
Jianxin Chen
Affiliation:
Baja Technology, LLC, Tucson, AZ, USA
Roger Henderson
Affiliation:
Radiochemistry, Lawrence Livermore National Laboratory, Livermore, CA, USA
Barbara Lafuente
Affiliation:
SETI Institute, Mountain View, CA, USA
*
Corresponding author: David Frederick Blake; Email: david.blake@nasa.gov
Rights & Permissions [Opens in a new window]

Abstract

In situ elemental imaging of planetary surface regolith at a spatial resolution of 100s to 1000s of microns can provide evidence of the provenance of rocks or sediments and their habitability, and can identify post-depositional diagenetic alteration affecting preservation. We use high-resolution elemental maps and XRF spectra from MapX, a flight prototype in situ X-ray imaging instrument, to demonstrate this technology in rock types relevant to astrobiology. Examples are given for various petrologies and depositional/diagenetic environments, including ultramafic/mafic rocks, serpentinites, hydrothermal carbonates, evaporites, stromatolitic cherts and diagenetic concretions.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
Copyright © The Author(s), 2024. Published by Cambridge University Press

Introduction

The value of microscale in situ observations of planetary surfaces

Many low-temperature planetary surface processes leave traces of their actions as features in the size range from 0.01 mm to several mm. Syn-sedimentary, authigenic and diagenetic features at the higher end of this size range include varves, vesicles, precipitates and concretions that can be used to decipher depositional and diagenetic processes. On the Earth, microbial life is known to exploit microscale disequilibria at boundaries where valence, chemical potential, pH, Eh, etc., vary on a length scale commensurate with the organisms themselves – tens to thousands of micrometres. These disequilibria can exist within cracks or veins in rocks and ice, at inter- or intra-crystalline boundaries, at sediment/water or sediment/atmosphere interfaces, or within fluid inclusions trapped inside minerals. Larger organized sedimentary structures such as stromatolites can provide macroscale evidence of biogenic influence, but abiotic examples are known as well and detailed study at the microscale is required for confirmation of biogenicity. Detection of accumulations of the biogenic elements C, N, O, P and S at appropriate concentrations on or in a mineral substrate could constitute permissive evidence of life, but context is also necessary. Does or did the putative biosignature exist in a habitable environment? Under what conditions of P, T and chemical potential was the host mineralogy formed? Most importantly, post-emplacement alteration (taphonomy) resulting from variations in temperature, pressure or fluid chemistry after deposition has the capacity to preserve evidence of biogenicity or its processes, or to erase such evidence completely.

In the characterization of solids (rocks, soil or planetary regolith), three fundamental properties can be used to elucidate conditions of formation and post-formational processing history: crystal structure, elemental composition and morphology. Based on these measurable properties, it is possible to determine provenance (useful in evaluating habitability or biogenicity) and taphonomy (useful in evaluating preservation potential).

Micro-mapping X-ray fluorescence spectroscopy (μ-mapping XRF)

The Mars Exploration Rovers Spirit and Opportunity each had instruments capable of elemental analysis (Alpha Particle X-ray Spectrometer (APXS); Rieder et al., Reference Rieder, Gellert, Bruckner, Klingelhofer, Dreibus, Yen and Squyres2003; Gellert et al., Reference Gellert, Rieder, Bruckner, Clark, Dreibus, Klinglehofer, Lugmair, Ming, Wanke, Yen, Zipfel and Squyres2006), iron mineralogy (Mössbauer Spectrometer; Morris et al., Reference Morris, Klingelhofer, Schroder, Rodionov, Yen, Ming, de Souza, Fleischer, Wdowiak, Gellert, Bernhardt, Evlanov, Zubkov, Foh, Bonnes, Kankeleit, Gutlich, Renz, Squyres and Arvidson2006), IR imagery (Miniature Thermal Emission Spectrometer (Mini-TES); Christensen et al., Reference Christensen, Mehall, Silverman, Anwar, Cannon, Gorelick, Kheen, Tourville, Bates, Ferry, Fortuna, Jeffreys, O'Donnell, Peralta, Wolverton, Blaney, Denise, Radermacher, Morris and Squyres2003; Ruff et al., Reference Ruff, Christensen, Blaney, Farrand, Johnson, Michaelski, Moersch, Wright and Squyres2006) and contextual hand-lens imagery (Athena Microscopic Imager (MI); Herkenhoff et al., Reference Herkenhoff, Squyres, Bell III, Maki, Arneson, Bertelsen, Brown, Collins, Dingizian, Elliott, Goetz, Hagerott, Hayes, Johnson, Kirk, McLennan, Morris, Scherr, Schwochert, Shiraishi, Smith, Soderblom, Sohl-Dickstein and Wadsworth2003; Herkenhoff et al., Reference Herkenhoff, Grotzinger, Knoll, McLennan, Weitz, Yingst, Anderson, Archinal, Arvidson, Barrett, Becker, Bell III, Budney, Chapman, Cook, Ehlmann, Franklin, Gadis, Galuszka, Garcia, Geissler, Hare, Howington-Krause, Johnson, Keszthelyi, Kirk, Lanagan, Lee, Leiff, Maki, Mullins, Parker, Redding, Roseik, Sims, Soderblom, Spanovich, Springer, Squyres, Stolper, Sucharski, Tucharski, Sullivan and Torson2008). The spatial resolution of all but the microscopic imager is on the order of centimetres, with the result that imaged features can rarely be related to elemental chemistry or mineralogy. The Mars Science Laboratory rover Curiosity (Grotzinger et al., Reference Grotzinger, Crisp, Vasavada, Anderson, Baker, Barry, Blake, Conrad, Edgett, Ferdowski, Gellert, Gilbert, Golombek, Gomez-Elvira, Hassler, Jandura, Litvac, Mahaffy, Maki, Meyer, Mitrofanov, Simonds, Vaniman, Welch and Wiens2012) is capable of landed optical imagery of rocks or soil with an ultimate spatial resolution of tens of microns (Mars Hand Lens Imager (MAHLI); Edgett et al., Reference Edgett, Yingst, Ravine, Caplinger, Ghaemi, Maki, Schaffner, Bell, Edwards, Herkenhoff and Heydari2012), and quantitative compositional data can be obtained from quite larger surface areas of 2–3 cm diameter using APXS (Campbell et al., Reference Campbell, Perrett, Gellert, Andrushenko, Boyd, Maxwell, King and Schofield2012). Occasionally, smaller features of a few mm have been characterized by comparing data from areas of regolith on and off a feature of interest (VanBommel et al., Reference VanBommel, Gellert, Berger, Campbell, Thompson, Edgett, McBride, Minitti, Prader and Boyd2016). The Laser-Induced Breakdown Spectrometry (LIBS) instrument on Curiosity (Maurice et al., Reference Maurice, Wiens, Saccoccio, Barraclough, Gasnault, Forni, Mangold, Baratoux, Bender, Berger, Bernadin, Berthe, Bridges, Blaney, Bouye, Cais, Clark, Clegg, Cousin, Cremers, Cros, DeFlores, Derycke, Dingler, Dromart, Dubois, Dupieux, Durand, d'Uston, Fabre, Faure, Gaboriaud, Gharsa, Herkenhoff, Kan, Kirkland, Kouach, Lacour, Langevin, Lasue, LeMouelic, Lescure, Lewin, Limonadi, Manhes, Mauchien, McKay, Meslin, Michel, Miller, Newsom, Orttner, Paillet, Pares, Parot, Perez, Pinet, Poitrasson, Quertier, Salle, Sotin, Sautter, Seran, Simmonds, Sirven, Stiglich, Striebig, Thocaven, Toplis and Vaniman2012) is capable of acquiring sub-mm qualitative to semi-quantitative elemental information, but these data are limited to individual spots and 1D or 2D arrays of analyses that are rudimentary at best when compared to companion optical information. Mineralogical data from the Chemistry and Mineralogy (CheMin) instrument (Blake et al., Reference Blake, Vaniman, Achilles, Anderson, Bish, Bristow, Chen, Chipera, Crisp, Des Marais, Downs, Farmer, Feldman, Fonda, Gailhanou, Ma, Ming, Morris, Sarrazin, Stolper, Treiman and Yen2012) and the Surface Analysis at Mars (SAM) instrument (Mahaffy et al., Reference Mahaffy, Webster, Cabane, Conrad, Coll, Atreya, Arvey, Barciniak, Benna, Bleacher, Brinkerhoff, Eigenbrode, Carignan, Cascia, Chalmers, Dworkin, Errigo, Everson, Franz, Farley, Feng, Frazier, Freissinet, Glavin, Harpold, Hawk, Holmes, Johnson, Jones, Jordan, Kellogg, Lewis, Lyness, Malespin, Martin, Maurer, McAdam, McLennan, Nolan, Noriega, Pavlov, Prats, Raaen, Sheinman, Sheppard, Smith, Stern, Tan, Trainer, Ming, Morris, Jones, Gunderson, Steele, Wray, Botta, Leshin, Owen, Battel, Jakosky, Manning, Squyres, Navarro-Gonzalez, McKay, Raulin, Sternberg, Buch, Sorensen, Kline-Schoder, Coscia, Szopa, Teinturier, Baffes, Feldman, Flesch, Forouhar, Garcia, Keymeulen, Woodward, Block, Arnet, Miller, Edmonson, Gorevan and Mumm2012) represent a homogenized volume of soil collected by a scoop, or a powdered drill core 1 cm in diameter and 3–6 cm deep (Sample Acquisition, Sample Processing and Handling (SA/SPaH) instrument; Anderson et al., Reference Anderson, Jandura, Okon, Sunshine, Roumeliotis, Beegle, Hurrowitz, Kennedy, Limonadi, McCloskey, Robinson, Seybold and Brown2012).

The PIXL instrument on the Mars 2020 rover Perseverance (Allwood et al., Reference Allwood, Wade, Foote, Elam, Hurrowitz, Battel, Dawson, Denise, Ek, Gilbert, King, Liebe, Parker, Pedersen, Randall, Sharrow, Sondheim, Allen, Arnett, Au, Basset, Benn, Bousman, Braun, Calvet, Clark, Cinquini, Conaby, Conley, Davidoff, Delaney, Denver, Diaz, Doran, Ervin, Evans, Flannery, Gao, Gross, Grotzinger, Hannah, Harris, Harris, He, Heirwegh, Hernandez, Hertzberg, Hodyss, Holden, Hummel, Jadusingh, Jorgensen, Kawamura, Kitiyakara, Kozaczek, Lambert, Lawson, Liu, Luchik, Macneal, Madsen, McLennan, McNally, Meras, Muller, Napoli, Naylor, Nemere, Ponomarev, Perez, Pootrakul, Romero, Rosas, Sachs, Schaefer, Schein, Setterfield, Singh, Song, Soria, Stek, Tallarida, Thompson, Tice, Timmermann, Treiman, Tsai, Uckert, Villalvazo, Wang, Wilson, Worel, Zamani, Zappe, Zong and Zimmerman2020) for the first time provided a capability to X-ray map and quantitatively analyse a small portion of Mars regolith with 125 μm lateral spatial resolution (see e.g. Liu et al., Reference Liu, Tice, Schmidt, Treiman, Kizovski, Hurowitz, Allwood, Henneke, Pedersen, VanBommel, Jones, Knight, Orenstein, Clark, Elam, Heirwegh, Barber, Beegle, Benzerara, Bernard, Beyssac, Bosak, Brown, Cardarelli, Catling, Christian, Cloutis, Cohen, Davidoff, Fairen, Farley, Flannery, Galvin, Grotzinger, Gupta, Hall, Herd, Hickman-Lewis, Hodyss, Horgan, Johnson, Jorgensen, Kah, Maki, Mandon, Mangold, McCubbin, McLennan, Moore, Nachon, Nemere, Nothdurft, Nunez, O'Neil, Quantin-Nataf, Sautier, Shuster, Siebach, Simon, Sinclair, Stack, Steele, Tarnas, Tosca, Uckert, Udry, Wade, Weiss, Wiens, Williford and Zorzano2022; Tice et al., Reference Tice, Hurowitz, Allwood, Jones, Orenstein, Davidoff, Wright, Pedersen, Henneke, Tosca, Moore, Clark, McLennan, Flannery, Steele, Brown, Zorzano, Hickman-Lewis, Liu, VanBommel, Schmidt, Kizovski, Treiman, O'Neil, Fairen, Shuster and Gupta2022; Scheller et al., Reference Scheller, Hollis, Cardarelli, Steele, Beegle, Bharita, Conrad, Uckert, Sharma, Ehlmann, Abbey, Asher, Benison, Berger, Beyssac, Bleefeld, Bosak and Brown2023). PIXL is a scanning instrument in which a focused 125 μm diameter X-ray beam is moved in X and Y across a regolith surface. Two silicon drift detectors (SDD) collect the fluoresced X-rays as the beam is moved, to produce point analyses, transects or X, Y elemental maps. The ultimate resolution of the analysis is defined by the distance between the X, Y positions.

A second method of X-ray mapping utilizes a flood of X-rays/γ-rays/α-particles to irradiate the entire sample at once, and an X-ray lens to focus the fluoresced photons from the sample onto a charge coupled device (CCD) detector. This method is called full-frame X-ray fluorescence imaging (FF-XRF). FF-XRF instruments have been demonstrated in space (Mercury Imaging X-ray Spectrometer (MIXS); Fraser et al., Reference Fraser, Carpenter, Rothery, Pearson, Martindale, Huovelin, Treis, Anand, Anttila, Ashcroft, Benkoff, Bland, Bowyer, Bradley, Bridges, Brown, Bullock, Bunce, Christensen, Evans and Whitehead2010) in the laboratory (Price et al., Reference Price, Fraser, Pearson, Nussey, Hutchinson, Holland, Turner and Pullan2004), as planetary instrument prototypes (Mapping X-ray Spectrometer (MapX); Blake et al., Reference Blake, Sarrazin and Bristow2015; Sarrazin et al., Reference Sarrazin, Blake, Thompson, Gailhanou, Chen and Bristow2016, Reference Sarrazin, Blake, Gailhanou, Marchis, Chalumeau, Webb, Walter, Schyns, Thompson and Bristow2018; Thompson et al., Reference Thompson, Blake and Sarrazin2017; Walroth et al., Reference Walroth, Sarrazin, Blake, Thompson, Gailhanou, Bristow, Marchis, Downs, Webb and Chalumeau2019) and as portable field instruments (MapX-II (Cartix); Walter et al., Reference Walter, Sarrazin, Gailhanou, Herrouard, Verney and Blake2018; Sepúlveda et al., Reference Sepúlveda, Pozzi-Escot, Falcon, Bermeo, Lebon, Moulherat, Sarrazin and Walter2020).

Irrespective of how X-rays are detected and imaged, with a full characterization of the excitation source and source–sample–detector geometry, it is possible to calculate spatially resolved quantitative compositional information using the fundamental parameters method (Solé et al., Reference Solé, Papillon, Cotte, Walter and Susini2007; Pia et al., Reference Pia, Weidenspointer, Augelli, Quintieri, Saracco, Sudhakar and Zoglauer2009; Schoonjans et al., Reference Schoonjans, Brunetti, Golosio, del Rio, Sole, Ferrero and Vincze2012).

Here we present elemental imaging results from a variety of sample types relevant to astrobiology, collected at ~125 μm lateral resolution in MapX-III, a FF/XRF planetary instrument prototype. These data are compared with imaging results collected at 50 μm lateral spatial resolution in a commercial scanned beam X-ray imaging instrument (EDAX Orbis).

Materials and methods

Hand samples, thin sections and thin section stubs were selected that represent a range of habitable environments, diagenetic fabrics and biologically mediated structures. Bulk samples were trimmed to petrologic thin-section size (19 × 26 mm) and thin sections were prepared. The prepared slides are intended to represent the smooth and dust-free abrasion targets typically studied by the Mars 2020 Perseverance rover's X-ray fluorescence (PIXL) and Raman (SHERLOC) microscale mapping instruments (Hickman-Lewis et al., Reference Hickman-Lewis, Moore, Razzell-Hollis, Tuite, Beegle, Bhartia, Grotzinger, Brown, Shkolyar, Cavalazzi and Smith2022; Razzell et al., Reference Razzell, Moore, Sharma, Beegle, Grotzinger, Allwood, Abbey, Bharita, Brown, Clark, Cloutis, Corpolongo, Henneke, Hickman-Lewis, Hurowitz, Jones, Liu, Martinez-Frias, Murphy, Pedersen, Shkolyar, Siljestrom, Steele, Tice, Treiman, Uckert, VanBommel and Yanchilina2022). In addition, thin sections allow for a petrographic characterization of the samples, useful for relating underlying microstructures with X-ray fluorescence maps. Representative portions of each sample were powdered and analysed by X-ray diffraction to determine the quantitative mineralogy of the samples.

Samples were first imaged and analysed in a commercial EDAX Orbis X-ray spectrometer in vacuum under a standard set of conditions: accelerating voltage 30 keV, beam current 350 μA (10.5 W), rhodium anode, 30 μm spot size polycapillary optic, 80 mm2 SDD detector. Samples were scanned with a step size of 50 μm, and a dwell time of 0.2 s. In a typical analysis of a 19 × 26 mm thin section, a matrix of 380 × 520 spots is collected, requiring ~11 h and 115 Watt hours (Wh) total beam energy.

The geometry of the MapX instrument used in this work is illustrated in Fig. 1. Sources bombard the sample with X-rays or α-particles/γ-rays, resulting in sample X-ray fluorescence (data for all figures in this paper were collected in MapX-III, a prototype having X-ray tube sources). X-ray photons emitted in the direction of an X-ray sensitive CCD imager pass through a 1:1 lens (X-ray Micro-Pore Optic (MPO)) that focuses a spatially resolved image of the X-rays onto the CCD. The CCD is operated in single photon counting mode (meaning that the imager is read often enough so that in the vast majority of cases, each pixel in the array contains either the charge generated by a single photon, or background). Under these conditions, both the energies and X, Y positions of individual X-ray photons are recorded. In a single analysis, several thousand frames are collected and processed in real time.

Figure 1. Schematic diagram of the MapX instrument. X-ray or radioisotope sources fluoresce the sample. Fluoresced X-ray photons characteristic of elements present in the sample are emitted over 2π steradians from their point of origin on the surface. Photons that enter the MPO within its acceptance angle are focused onto an energy-discriminating X-ray sensitive CCD in a 1:1 configuration with the sample. Hundreds to thousands of short exposures, each one a complete 2D X-ray image of the sample, are summed to produce an hdf5 data file.

MapX-III (Fig. 2) is comprised of custom flight-qualifiable camera electronics that drive a flight-qualified e2v frame-transfer front illuminated CCD-224. The camera is interfaced to a high vacuum enclosure that contains the CCD-224 imager, a Photonis MPO lens and sample stage. Two Newton Scientific Au target transmission window X-ray tubes, operated at 15 keV and 25 μA (0.750 W total) illuminate the sample through ports in the chamber.

Figure 2. (a). Cross-section of MapX-III prototype, showing sample position, X-ray tubes, MPO and CCD224 imager in an evacuated chamber. (b). MapX-III prototype. With the exception of the X-ray tubes, MapX-III is assembled from flight qualified or qualifiable components: two Newton Scientific M54 Au target transmission X-ray tubes; an e2v front illuminated, deep depleted frame transfer CCD-224 (legacy from CheMin, 600 × 600 40 μm pixels); and an Ir coated 1.5 mm thick MPO with 20 μm square channels and 6 μm thick walls. Samples in MapX-III are held in the same vacuum as the CCD, eliminating the need for a Be window and allowing for element detection down to Z = 11 (Na). MapX-III has a lateral spatial resolution of ~125 μm and an energy resolution of ~200 eV.

The overall geometry of the instrument (distances between sample, MPO and CCD, and the characteristics of the MPO) was optimized through the testing of MPO lenses at Stanford Synchrotron Radiation Light Source Beam Line 2–3 (Sarrazin et al., Reference Sarrazin, Blake, Gailhanou, Marchis, Chalumeau, Webb, Walter, Schyns, Thompson and Bristow2018) and by ray-tracing simulations (Gailhanou et al., Reference Gailhanou, Sarrazin and Blake2018). The measured lateral spatial resolution of the instrument is 100 μm with a depth of field of ±5 mm with minimal loss of resolution.

Samples were imaged and analysed in MapX-III under the conditions noted in the text for each analysis. The equivalent of ≤3 h of analysis time is considered as a nominal baseline for a deployed planetary instrument. MapX can be equipped with either an X-ray tube source or a radioisotope source to fluoresce the sample. For an X-ray tube sourced instrument, we modelled 25 KeV accelerating voltage and 200 μA beam current (5 W total power in the beam, 15 Wh total energy per analysis). For a radioisotope sourced instrument, we modelled 30 mCi of 244Cm, equivalent to that used in the APXS instruments (Rieder et al., Reference Rieder, Gellert, Bruckner, Klingelhofer, Dreibus, Yen and Squyres2003; Gellert et al., Reference Gellert, Rieder, Bruckner, Clark, Dreibus, Klinglehofer, Lugmair, Ming, Wanke, Yen, Zipfel and Squyres2006; Economou, Reference Economou and Singh2011). The radioisotope 244Cm produces 14 and 18 KeV γ-rays (suitable for fluorescing Kα lines from calcium through zirconium) and 5.8 MeV α-particles (suitable for fluorescing Kα lines from oxygen through potassium).

Whether from the EDAX instrument or MapX-III, once data are collected, they are processed identically. In MapX-III, the individual images (frames) collected by the CCD are binned by energy and combined into an X × Y × photon energy hdf5 data cube (see Supplementary Information and Fig. S1). The data cube is commonly hundreds of megabytes in size, too large for routine downlink from a spacecraft. Software was developed to generate reduced data records more suitable for downlink. Single-element maps, Mg–Ni are computed from the hdf5 file using custom python code. These maps are qualitative only in that fluorescence efficiency, absorption effects, detector efficiency, etc., are not accounted for. In a 3 h nominal analysis, single-pixel spectra lack sufficient counting statistics to be quantified. However, qualitative single-element maps are the highest resolution and highest sensitivity data products produced by the instrument.

We use an unsupervised machine learning program to identify regions of similar composition (regions of interest or ‘ROI’) in the hdf5 data cube. ROI are identified by finding element–element correlation clusters in N × N dimensional space where N is the number of relevant elements (see Supplementary Information and Fig. S2). These clusters are then applied to the 2D image to generate a map of ROI. The XRF spectra from the individual pixels contained in each ROI are summed to generate high signal-to-noise quantifiable XRF spectra for ground processing. The hdf5 data cube is reduced to a set of single-element maps, a set of ROI (composition maps) having unique element compositions and XRF spectra from each ROI. If necessary, the hdf5 data cube can be reprocessed to generate additional data products on the basis of the initial quick-look elemental images and ROI.

Results

Examples are organized by rock type and depositional/diagenetic environment (ultramafic/mafic rocks, serpentinites, hydrothermal carbonates, evaporites, stromatolitic cherts and diagenetic concretions). All raw hdf5 files, element maps, ROI, ROI spectra, etc., for the samples shown here and many others are archived in and downloadable from the Planetary X-ray Fluorescence Petrography database: https://doi.org/10.48484/R7WV-3T89.

Ultramafic xenoliths from Spitsbergen, Norway; analogues of the ALH84001 meteorite

Xenoliths of mantle material comprise as much as 20% by volume of several quaternary volcanic centres in Spitsbergen, Norway. The xenoliths include spinel lherzolites and pyroxenites having mineralogies broadly similar to that of the Mars meteorite ALH84001, purported to show evidence of microbial life on ancient Mars (McKay et al., Reference McKay, Gibson, Thomas-Keptra, Vali, Romanek, Clemett, Chillier, Maechling and Zare1996).

While the preponderance of available evidence now supports an abiotic origin for the carbonate globules in ALH84001 (e.g. see Treiman et al., Reference Treiman, Amundsen, Blake and Bunch2002), this meteorite nevertheless represents an important rock type for Mars. The mineralogy/petrology of ALH84001 demonstrates that deep-seated volcanism occurred ~4.1 Gy before the present, influenced by a low-temperature near-surface or surface hydrothermal event predating a shock that caused the ejection of ALH84001 from the planet's surface. The hydrothermal event was brief enough, or at low enough temperature (<<200°C) that the ultramafic minerals, unstable in the presence of water, were not altered to serpentine.

Figure 3 illustrates a partial EDAX and MapX-III dataset from AMASE08 UI-3, one of several ultramafic xenoliths from Spitsbergen, Norway that were analysed by X-ray diffraction, electron microprobe and optical point counting (Treiman et al., Reference Treiman, Robinson, Blake and Bish2010). The highest resolution data returned by the instrument are the individual elemental images. ROI representing regions of the sample having common compositions were calculated from the hdf5 data cube using an unsupervised machine learning algorithm (see Supplementary Information and Figs S1 and S2). In this example, ROI are seen to represent three minerals plus basalt: clinopyroxene (cpx, ROI 1), orthopyroxene (opx, ROI 3), chrome spinel (csp, ROI 4) and host basalt (bas, ROI 2). Mineral identifications were made separately on UI-3 thin sections and powder separates using optical point counting, electron microprobe and X-ray diffraction data (enstatite 26.09 wt.%, forsterite 7.79 wt.%, augite 59.22 wt.%, albite 4.05 wt.%, spinel 2.85 wt.%) (Treiman et al., Reference Treiman, Robinson, Blake and Bish2010). XRF spectra returned for each ROI can be quantified using fundamental parameters methods, and in favourable cases, it is possible to relate the compositions of ROI to specific minerals using the RRUFF database (Lafuente et al., Reference Lafuente, Downs, Yang, Stone, Armbruster and Danisi2015). Total beam power: 9 Wh, equivalent to 120 min of analysis with a nominal X-ray tube or 244Cm radioisotope-based spacecraft instrument.

Figure 3. Ultramafic xenolith from Spitsbergen, Norway. (a) Optical image of a thin section stub of ultramafic xenolith AMASE08 UI-3; (b) RGB element map from commercial EDAX-Orbis instrument (50 μm resolution), Red = Fe, Green = Ca, Blue = Cr; (c) RGB element map from MapX-III (~125 μm resolution), same colour scheme as in (b); (d–h) MapX-III elemental maps; (i–l) ROI selected by element correlation cluster analysis from MapX-III hdf5 data cube; (m) ROI map from MapX-III (numbers refer to labelled ROI spectra shown in (o)); (n) ROI map from EDAX (note: colours for ROI-2 and ROI-3 are reversed); (o) summed XRF spectra from individual ROI; ROI 1 = clinopyroxene, ROI 2 = basalt matrix, ROI 3 = orthopyroxene, ROI 4 = chrome spinel (total MapX-III beam power = 9 Wh (120 min analysis)).

It is interesting to note that one of the first rock types imaged in Jezero crater by the PIXL instrument (Farley et al., Reference Farley, Stack, Shuster, Horgan, Hurowitz, Tarnas, Simon, Sun, Scheller, Moore, McLennan, Vasconcelos, Wiens, Treiman and Mayhew2022; Liu et al., Reference Liu, Tice, Schmidt, Treiman, Kizovski, Hurowitz, Allwood, Henneke, Pedersen, VanBommel, Jones, Knight, Orenstein, Clark, Elam, Heirwegh, Barber, Beegle, Benzerara, Bernard, Beyssac, Bosak, Brown, Cardarelli, Catling, Christian, Cloutis, Cohen, Davidoff, Fairen, Farley, Flannery, Galvin, Grotzinger, Gupta, Hall, Herd, Hickman-Lewis, Hodyss, Horgan, Johnson, Jorgensen, Kah, Maki, Mandon, Mangold, McCubbin, McLennan, Moore, Nachon, Nemere, Nothdurft, Nunez, O'Neil, Quantin-Nataf, Sautier, Shuster, Siebach, Simon, Sinclair, Stack, Steele, Tarnas, Tosca, Uckert, Udry, Wade, Weiss, Wiens, Williford and Zorzano2022) has a similar fabric and mineralogy to the ultramafic xenoliths analysed from Spitzbergen (Treiman et al., Reference Treiman, Amundsen, Blake and Bunch2002, Reference Treiman, Robinson, Blake and Bish2010) and illustrated in Fig. 3.

Serpentinization of mafic and ultramafic rocks: a source of energy for chemosynthetic ecosystems

Serpentinization is the process by which ultramafic rocks containing olivine and pyroxene react with water to form magnetite, serpentine and hydroxide minerals. As a consequence of this reaction, molecular hydrogen is released, a potential source of chemical energy for life (McCollom and Shock, Reference McCollom and Shock1997; Sleep et al., Reference Sleep, Meibom, Fridriksson, Coleman and Bird2004). Serpentinizing systems are regarded as important analogues for potential early ecosystems on Earth and Mars, where highly reducing mineralogy was likely widespread in an undifferentiated crust (Kelley et al., Reference Kelley, Karson, Fruh-Green, Yoerger, Shank, Butterfield, Hayes, Schrenk, Olson, Proskurowski, Jakuba, Bradley, Larson, Ludwig, Glickson, Buckman, Bradley, Brazelton, Roe, Elend, Delacour, Bernasconi, Lilley, Baross, Summons and Sylvia2005; Schulte et al., Reference Schulte, Blake, Hoehler and McCollom2006).

The second and third samples collected and analysed by the MSL Curiosity rover during its investigation of Gale crater on Mars were of a ~3.7 G-year-old mudstone interpreted to be a lithified sediment deposited in a lacustrine environment (Grotzinger et al., Reference Grotzinger, Sumner, Kah, Stack, Gupta, Edgar, Rubin, Lewis, Schieber, Mangold, Milliken, Conrad, Des Marais, Farmer, Siebach, Calef III, Hurowitz, McLennan, Ming, Vaniman, Crisp, Vasavada, Edgett, Malin, Blake, Gellert, Mahaffy, Wiens, Maurice and Grant2013; Vaniman et al., Reference Vaniman, Bish, Ming, Bristow, Morris, Blake, Chipera, Morrison, Treiman, Rampe, Rice, Achilles, Grotzinger, McLennan, Williams, Bell III, Newsom, Downs, Maurice, Sarrazin, Yen, Morookian, Farmer, Stack, Milliken, Ehlmann and Sumner2013; Bristow et al., Reference Bristow, Bish, Vaniman, Morris, Blake, Grotzinger, Rampe, Crisp, Achilles, Ming, Ehlmann, King, Bridges, Eigenbrode, Sumner, Chipera, Moorokian, Treiman, Morrison, Downs, Farmer, Des Marais, Sarrazin, Floyd, Mishna and McAdam2015). When compared to the elemental composition of global soil (used as a proxy for average Martian basalt) collected during Curiosity's first analysis on Mars (Bish et al., Reference Bish DL, Blake, Vaniman, Chipera, Morris, Ming, Treiman, Sarrazin, Morrison, Downs, Achilles, Yen, Bristow, Crisp, Rampe, Stolper, Spanovich and Morookian2013; Blake et al., Reference Blake, Morris, Kocurek, Morrison, Downs, Bish, Ming, Edgett, Rubin, Goetz, Madsen, Sullivan, Gellert, Campbell, Treiman, McClennan, Yen, Grotzinger, Vaniman, Chipera, Achilles, Rampe, Sumner, Meslin, Maurice, Forni, Gasnault, Fisk, Schmidt, Mahaffy, Leshin, Glavin, Steele, Freissinet, Navarro-Gonzalez, Yingst, Kah, Bridges, Lewis, Bristow, Farmer, Crisp, Stolper, Des Marais and Sarrazin2013), the elemental composition of the mudstone (now principally comprised of secondary minerals) is a close match, suggesting that the mudstone was similar in its initial mineralogy and experienced a closed-system diagenetic event. On the basis of mass-balance calculations, Bristow et al. (Reference Bristow, Bish, Vaniman, Morris, Blake, Grotzinger, Rampe, Crisp, Achilles, Ming, Ehlmann, King, Bridges, Eigenbrode, Sumner, Chipera, Moorokian, Treiman, Morrison, Downs, Farmer, Des Marais, Sarrazin, Floyd, Mishna and McAdam2015) propose a mechanism akin to serpentinization in which original mafic minerals (e.g. olivine), unstable in the presence of water, transformed into a stable assemblage including clay minerals (ferrian saponite; Treiman et al., Reference Treiman, Morris, Agresti, Graff, Achilles, Rampe, Bristow, Blake, Vaniman, Bish, Chipera, Morrison and Downs2014) and magnetite. Hydrogen would have been released in such a reaction, and on the basis of the mineral assemblage and its sedimentological context, this ancient lakebed was identified as a habitable environment for chemosynthetic organisms (Grotzinger et al., Reference Grotzinger, Sumner, Kah, Stack, Gupta, Edgar, Rubin, Lewis, Schieber, Mangold, Milliken, Conrad, Des Marais, Farmer, Siebach, Calef III, Hurowitz, McLennan, Ming, Vaniman, Crisp, Vasavada, Edgett, Malin, Blake, Gellert, Mahaffy, Wiens, Maurice and Grant2013).

Orbital spectral observations made by the CRISM instrument on Mars Reconnaissance Orbiter have identified serpentine and serpentine-like minerals in many locations on Mars, suggesting the presence of widespread and deep-seated serpentinization reactions in the crust of Mars, creating potential habitable environments early in Mars history (Ehlmann and Edwards, Reference Ehlmann and Edwards2014 and references therein).

The most widespread serpentinizing environment on the Earth is the seafloor, where mafic and ultramafic oceanic crust interacts with seawater (Deming and Baross, Reference Deming and Baross1993; McCollom and Shock, Reference McCollom and Shock1997; Kelley et al., Reference Kelley, Karson, Blackman, Fruh-Green, Butterfield, Lilley, Olson, Schrenk, Roe, Lebon and Rivizzigno2001a, Reference Kelley, Karson, Blackman, Fruh-Green, Butterfield, Lilley, Schrenk, Olson, Roe and Lebon2001b). ‘Ophiolite sequences’, slivers of oceanic crust thrust upward into the shallow continental crust as a result of plate tectonics, are found in northern California and southern Oregon (Coleman, Reference Coleman2000). These rocks, in contact with connate fluids at moderate to low temperatures (~200°C), are presently undergoing serpentinization (Barnes et al., Reference Barnes, LaMarche and Himmelberg1967; Barnes and O'Neil, Reference Barnes and O'Neil1969; Blake and Peacor, Reference Blake and Peacor1985).

Schulte et al. (Reference Schulte, Blake, Hoehler and McCollom2006) analysed hand samples of olivine-rich rocks from actively dewatering outcrops in the Coast Range Ophiolite in N. California. Figure 4 illustrates a partial dataset from a thin section stub of one such sample described in Schulte et al. (Reference Schulte, Blake, Hoehler and McCollom2006). The modal proportions of minerals in the original rock prior to serpentinization were estimated to be 70% olivine, 20% orthopyroxene, 7% clinopyroxene and 3% other minerals. Primary olivine and pyroxene account for about half of the present mineralogy, with a fine-grained groundmass representing the remaining bulk (and comprised of serpentine minerals, fine-grained magnetite and Fe-rich brucite). Despite the fact that most mineral phases are below the resolution of both the EDAX and MapX-III, the mafic-ultramafic elemental composition of ROI and the rock fabric displayed in element images (secondary veining and the absence of large crystals) strongly suggest that this rock is a serpentinite, and represents a potentially habitable environment.

Figure 4. Thin section stub of a partially serpentinized cobble from Complexion Springs, California. (a) Optical image, white rectangle shows area imaged in MapX-III; (b) RGB element map from EDAX-Orbis instrument (50 μm resolution), Red = Fe, Green = Ca, Blue = Cr, white rectangle shows area imaged in MapX-III; (c) RGB element map from MapX-III (~125 μm resolution), same colour scheme as in b; (d–g) element images from MapX-III; (h) ROI selected by elemental correlation cluster analysis from EDAX hdf5 data cube (numbers refer to individual ROI also shown in (i) from the MapX-III instrument; (i) ROI selected by elemental correlation cluster analysis from MapX-III hdf5 data cube; (j) MapX-III XRF spectra from individual ROI, numbers refer to ROI labelled in (i) (total MapX-III beam power = 2.25 Wh (30 min analysis)).

Carbonates on earth and Mars

Carbonates are important rock-forming minerals on early Earth and should have been on early Mars as well. On early Mars, a dense, CO2-rich atmosphere in the presence of surface and near-surface water would promote the dissolution of mafic minerals by carbonic acid and the precipitation of Ca-Fe-Mg carbonates at near-normal pH.

Meteoritic carbonates

Carbonates were described in the Martian meteorites Nakhla (Bridges and Grady, Reference Bridges and Grady2000; Bridges and Schwenzer, Reference Bridges and Schwenzer2012) and in ALH84001 (McKay et al., Reference McKay, Gibson, Thomas-Keptra, Vali, Romanek, Clemett, Chillier, Maechling and Zare1996). While these carbonates comprise a volumetrically small fraction of the overall mineralogy in these meteorites, they are nevertheless important because detailed information about their paragenesis has been obtained with laboratory-based instrumentation. In all cases, carbonate formation can best be explained by surface or near-surface hydrothermal activity in mafic or ultramafic rocks (Bridges et al., Reference Bridges, Hicks, Treiman, Filiberto and Schwenzer2018 and references therein). Treiman et al. (Reference Treiman, Amundsen, Blake and Bunch2002) describe carbonate globules contained within ultramafic xenoliths from the Sverrefjell volcano (Spitsbergen, Norway) that are nearly identical to those found in the ALH84001 meteorite and present evidence that the globules are hydrothermal precipitates formed within a shallow sub-surface volcanic complex. Figure 5 shows data from ultramafic xenolith AMASE08 UI-6. X-ray diffraction analysis shows it to be comprised of 15.7% enstatite, 81.1% olivine, 1.9% augite and 1.2% albite. In this case (unlike AMASE08 UI-3 shown in Fig. 3), only a small amount of Ca-containing clinopyroxene is present, so that carbonate globules are identifiable based on Ca element images. Treiman et al. (Reference Treiman, Robinson, Blake and Bish2010) describe the secondary mineralization shown in (Fig. 5(f)) as partial melts and products of aqueous alteration near and around spinels, the alteration products including zoned spherules of (CaMg,Fe)CO3.

Figure 5. Ultramafic xenolith AMASE UI-6, an orthopyroxenite comprised principally of orthopyroxene and olivine with minor chrome spinel and carbonate. (a) Petrographic thin section, small black rectangle shows area of thin section magnified in (f); (b) RGB element map from EDAX-Orbis instrument, Red = Fe, Green = Ca, Blue = Cr, large white rectangle shows area imaged in MapX-III; (c) RGB element map from MapX-III, colour scheme same as in (b); (d) Ca element image from EDAX instrument; (e) Ca element image from MapX-III instrument; (f) optical micrograph of portion of thin section outlined in figures a–e, g, h, showing carbonates (C) and other secondary mineralization surrounding chrome spinel (S), Opx – orthopyroxene, Ol = olivine; (g) ROI map from EDAX instrument; (h) ROI map from MapX-III instrument; (i) XRF spectra from MapX-III ROI labelled 1 and 3 in (h) (total beam power = 9 Wh (120 min analysis)).

Martian carbonates

Orbital observations of Mars with the Compact Reconnaissance Imaging Spectrometer for Mars (CRISM) instrument have identified volumetrically significant deposits of carbonate in Nili Fossae (Ehlmann et al., Reference Ehlmann, Mustard, Murchie, Poulet, Bishop, Brown, Calvin, Clark, Des Marais, Milliken, Roach, Rousch, Swayze and Wray2008) and elsewhere. Morris et al. (Reference Morris, Ruff, Gellert, Ming, Arvidson, Clark, Golden, Siebach, Klingelhofer, Schroder, Fleischer, Yen and Squyres2010) report the discovery of carbonate-rich basaltic outcrops in Gusev crater, Mars using instruments aboard the Mars Exploration Rover Spirit. Comanche Spur is an apparent volcaniclastic deposit with a fracture filling comprised of Mg-Fe-Ca carbonate. The composition of the carbonate is reported to be Mg0.62Fe0.25Mn0.02Ca0.11CO3 based on Mossbauer, APXS and Mini-TES instrumental data. The closest terrestrial analogue to the Comanche carbonate is from Spitsbergen, Norway in carbonate-cemented, breccia-filled volcanic pipes associated with late-stage hydrothermal activity. Morris et al. (Reference Morris, Blake, Bish, Ming, Agresti, Treiman, Steele and Amundsen2011) describe this occurrence (see also Treiman et al., Reference Treiman, Robinson, Blake and Bish2010; Blake et al., Reference Blake, Treiman, Morris, Bish, Amundsen and Steele2011). The Sverrefjell volcanic province is quaternary in age and manifests itself as volcanic cinder cones which on the basis of pillow basalts exposed on their flanks, must have erupted under water (Treiman et al., Reference Treiman, Amundsen, Blake and Bunch2002). Laboratory and field evidence suggest that the carbonates were deposited by hydrothermal fluids, likely heated by the volcanic rock itself, flowing through pre-existing volcanic pipes in a hydrothermal cell. Figure 6 shows EDAX and MapX results from a thin section of a carbonate-cemented basalt breccia from this locality. At the 50 μm spatial resolution of the EDAX instrument, the prismatic texture of the cement is evident, as are void spaces, demonstrating rapid crystal growth in a surface or shallow sub-surface aqueous environment. At the 125 μm resolution of MapX-III, void spaces are evident, but not the prismatic texture. Elemental maps document changes in carbonate chemistry during cement deposition, and the predominance of the divalent cations Ca-Mg-Mn suggest that the carbonate is rhombohedral. In fact, XRD analyses of these cements (Blake et al., Reference Blake, Treiman, Morris, Bish, Amundsen and Steele2011) reveal a wide variety of rhombohedral carbonates, including Mg-calcite (MgxCa1-xCO3), huntite (Mg3Ca(CO3)4), magnesite (MgCO3), dolomite/ankerite (Ca(Mg,Fe)(CO3)2), siderite (FeCO3) and in some cases late-stage aragonite (orthorhombic CaCO3). From these data, it appears that specific carbonate mineral chemistries (e.g. thin rind of Mn, calcium core, magnesium rim in the cement) are reflective of changing hydrothermal conditions and what divalent cations are present in the solution.

Figure 6. Carbonate-cemented basalt breccia from Spitsbergen, Norway. (a) Petrologic thin section; (b) RGB element map from EDAX-Orbis instrument (50 μm resolution), Red = Fe, Green = Ca, Blue = Mg, white rectangle shows area imaged in MapX-III; (c) RGB element map from MapX-III, same colour scheme as b; (d–g) Element images from EDAX (l) and MapX-III (r); (h) ROI map from EDAX (l) and MapX-III (r) note that ROI 1 from MapX is blue while the same ROI in EDAX is yellow; (i) ROI XRF spectra for MapX-III, ROI labelled in (h) (total beam power = 9 Wh (120 min analysis)).

Ancient cherts: fine-scale early preservation of evidence of life and habitability

On the early Earth, Proterozoic silica deposition (abiotic silica precipitation as either a primary chemical sediment or as an early diagenetic cement in the marine environment) is a predominant mechanism for fine-scale preservation of primary depositional fabrics, early diagenetic features and remnants of the earliest microbial life (cellular processes and organic biosignatures). Original silica was emplaced as amorphous and/or sparingly crystalline opal-A and opal-CT from a silica-saturated ocean (or from silica-rich ground water). These original components have been recrystallized into microcrystalline quartz (i.e. chert) in the rock record. Knoll (Reference Knoll1985) suggests that well-preserved stromatolytic occurrences such as the Gunflint Chert (Tyler and Barghoorn, Reference Tyler and Barghoorn1954; Barghoorn and Tyler, Reference Barghoorn and Tyler1965; Awramik, Reference Awramik and Walter1976) were formed in peritidal environments in which evaporation and heightened salinity played a role in silica deposition. It appears that the presence of organic material is a common feature in early silicification, and petrographic observations of cement fabric suggest that initial opal-A or opal-CT nucleation occurs on individual microbial sheaths and processes. A second type of early silica deposition with fine-scale preservation involves the replacement of pre-existing carbonate completely or partially by silica.

Silica deposition as a chemical sediment or an early diagenetic replacement cement should have been common on early Mars, where water interacted with mafic rocks, dissolving olivine, pyroxene and other minerals with the liberation of aqueous silica (McLennan, Reference McLennan2003). Tosca and Knoll (Reference Tosca and Knoll2009) and references therein document numerous detections of hydrated silica on the surface of Mars from both in situ and remote observations. For example, evaporative sedimentary rocks at Meridiani Planum contain as much as 20 wt.% hydrated silica (McLennan et al., Reference McLennan, Bell III, Calvin, Christensen, Clark, de Souza, Farmer, Farrand, Fike, Gellert, Gosh, Glotch, Grotzinger, Hahn, Herkenhoff, Hurowitz, Johnson, Johnson, Joliff, Klingelhofer, Knoll, Learner, Malin, McSween, Pocock, Ruff, Soderblom, Squyres, Tosca, Watters, Wyatt and Yen2005), and opaline silica was identified in the Columbia Hills of Gusev Crater, interpreted as the product of aqueous activity, possibly related to hydrothermal processes (Squyres et al., Reference Squyres, Arvidson, Ruff, Gellert, Morris, Ming, Crumpler, Farmer, Des Marais, Yen, McClennan, Calvin, Bell III, Clark, Wang, McCoy, Schmidt and DeSouza2008). Remote observations from MRO/CRISM have identified opaline silica (Milliken et al., Reference Milliken, Swayze, Arvidson, Bishop, Clark, Ehlmann, Green, Grotzinger, Morris, Murchie, Mustard and Weitz2008) as well.

The discovery of amorphous and crystalline silica in ~3.5 Ga Murray formation lacustrine mudstone sequences in Gale crater on Mars (Morris et al., Reference Morris, Vaniman, Blake, Gellert, Chipera, Rampe, Ming, Morrison, Downs, Treiman, Yen, Grotzinger, Achilles, Bristow, Crisp, Des Marais, Farmer, Fendich, Frydenvang, Graff, Morookian, Stolper and Schwenzer2016; Rampe et al., Reference Rampe, Ming, Blake, Bristow, Chipera, Grotzinger, Morris, Morrison, Vaniman, Yen, Achilles, Craig, Des Marais, Downs, Farmer, Fendrich, Gellert, Morookian, Peretyazhko, Sarrazin, Treiman, Berger, Eigenbrode, Fairen, Forni, Gupta, Hurowitz, Kah, Lanza, Schmidt, Siebach, Sutter and Thompson2017) raises the possibility that retention of depositional/early diagenetic features in silica (including preservation of organic biosignatures, if present) is possible and even likely on Mars. Silica-rich sequences of the Murray formation (and of other similar sedimentary deposits found elsewhere on Mars) may have the highest probability of retention of evidence of biogenicity (or of abiotically produced endogenous/exogenous carbon) and may offer a taphonomically favourable window into the biologic potential of early Mars.

Figure 7 shows a thin section stub of finely preserved stromatolitic material from the 2.1 Ga Gunflint Chert, Schrieber Beach locality. The chert preserves carbonate features that were likely the original material of the stromatolitic structures. However, abiotically precipitated silica is ubiquitous in the Precambrian, and locations such as Schrieber Beach were only found through extensive fieldwork by paleontologists (and indeed, much of the Gunflint formation was subject to extremes of diagenesis that would have overprinted the features preserved in the Schrieber Beach locality). As shown in the figure, spatially resolved element images can be highly useful in identifying the small fraction of chert samples that may contain evidence of early preservation and of potential biogenicity. It is interesting to note that there are no mineral phases in this sample that are large enough to be resolved by the 125 μm spatial resolution of MapX-III or even the 50 μm spatial resolution of the EDAX instrument (e.g. the Ca-Fe-Mn rich regions in this sample were found by optical microscopy to be comprised of clusters of ~10 μm carbonate rhombs, and the fabric of the microcrystalline chert is only resolvable in a petrographic microscope). The key feature of this sample that would separate it from others is the apparent chemical disequilibrium and heterogeneous distribution of elements in the images and the composition of the ROI at mm- and larger scales.

Figure 7. Gunflint Chert, Schrieber Beach locality stromatolite. (a) Polished thin section stub; (b) RGB element maps from EDAX-Orbis and MapX-III instruments, Red = Fe, Green = Ca, Blue = Mg; (c–f) comparison element images from EDAX (l) and MapX-III (r) for Si, Ca, Fe and Mn; (g–j) ROI maps from EDAX instrument; (k–m) ROI maps from MapX-III; (n) XRF spectra for EDAX ROI; (o) XRF spectra for MapX-III ROI (total beam power = 9 Wh (120 min analysis)).

Evaporites

On early Mars, volcanic sulphur emanations combined with the presence of surface water suggest that sulphates of all kinds should be present on the surface. Indeed, on the basis of extensive observations of sulphates from orbit by the OMEGA instrument (Bibring et al., Reference Bibring, Langevin, Mustard, Poulet, Arvidson, Gendrin, Gondet, Mangold, Pinet, Forget, Berthe, Gendrin, Gomez, Gondet, Jouglet, Poulet, Soufflot, Vincendon, Combes, Drossart, Encrenaz, Fouchet, Merchiorri, Belluci, Alteieri, Formisano, Capaccioni, Cerroni, Coradini, Fonti, Korablev, Kottsov, Ignatiev, Moroz, Titov, Zasova, Loiseau, Pinet, Soute, Schmitt, Sotin, Hauber, Hoffmann, Jaumann, Keller, Arvidson, Mustard, Duxbury and Neukum2006), sulphate deposits are known in the Vallis Marineris (Juventae Chasma), Terra Meridiani (Squyres et al., Reference Squyres, Grotzinger, Arvidson, Bell III, Calvin, Christensen, Farrand, Herkenhoff, Johnson, Klingelhofer, Knoll, McLennan, McSween, Morris, Rice, Rieder and Soderblom2006) and many other places on the Mars surface. In Juventae Chasma, these deposits are over 2 km thick and have an unknown origin (Bishop et al., Reference Bishop, Parente, Weitz, Noe Dobrea, Roach, Murchie, McGuire, McKeown, Rossi, Brown, Calvin, Milliken and Mustard2009). MSL's landing site in Gale crater was chosen in part because the central mound of the crater (Mt. Sharp) is comprised of flat-lying phyllosilicate-rich sediments overlain by a sulphate layer (Milliken et al., Reference Milliken, Grotzinger and Thomson2010). During MSL's traverse, gypsum, bassanite and anhydrite have been detected in secondary diagenetic veins and fillings during its entire traverse of the crater. Most recently, the mineral Starkeyite (MgSO4.4H2O) and an amorphous Mg-sulphate were identified by the CheMin XRD instrument on the Curiosity rover (Chipera et al., Reference Chipera, Vaniman, Rampe, Bristow, Martinez, Tu, Peretyazhko, Yen, Gellert, Berger, Rapin, Morris, Thompson, Simpson, Achilles, Tutolo, Downs, Fraeman, Fischer, Blake, Treiman, Morrison, Thorpe, Gupta, Dietrich, Downs, Castle, Craig, Des Marais, Hazen, Vasavada, Hausrath, Sarrazin and Grotzinger2023). Starkeyite was seen to dehydrate to amorphous Mg-sulphate in the low relative humidity environment of the instrument, and it is likely that much of the detected amorphous Mg-sulphate was originally crystalline. However, the paragenesis of these sulphates is unknown.

The Castile Formation evaporite on Earth is a cm-to-mm scale laminated deposit of gypsum/dolomite (CaSO4.2H2O/CaMg(CO3)2) couplets that is 400–500 m thick that filled portions of the Permian basin of southeast New Mexico (~250 MYa). The deposit is thought to have formed in a shallow basin after reef growth cut off the influx of marine water, causing the basin to fill with gypsum and carbonate (Kirkland et al., Reference Kirkland, Dennison and Dean2000). The carbonate–gypsum couplets are thought to represent annual changes in sedimentation, with gypsum layers forming during high evaporation (summer), and carbonate layers forming during low evaporation (winter). Some variants also include halite. Figure 8 shows X-ray mapping results from a petrologic thin section of Castile Formation evaporite, illustrating the sulphate and carbonate diurnal varves. Varves are plainly visible at both 50 μm resolution (EDAX) and 125 μm resolution (MapX-III) corresponding to alternating sulphate and carbonate layers. In the MapX ROI generated by machine learning, the varves could not be discriminated. However, the hdf5 data cube stored on the instrument is available for reprocessing and new algorithms could be devised (e.g. ‘identify the locus of X, Y pixels containing Ca and Mg’), the data cube reprocessed and quantifiable XRF spectra generated from the newly calculated ROI.

Figure 8. Castile Formation evaporite, Permian basin, New Mexico. (a) Petrographic thin section; (b) RGB element map from EDAX-Orbis instrument, Red = S, Green = Ca, Blue = Mg; (c) RGB element map from MapX-III instrument, same colour scheme as b; (d–i) comparison element images from EDAX (l) and MapX-III (r) for S, Ca and Mg (total beam power = 9 Wh (120 min analysis)).

Hematite/goethite diagenetic crystallization

Clastic, sand-sized particles deposited in aeolian, lacustrine, fluvial and sabkha type environments quite commonly retain their porosity and permeability after deposition. Syn-sedimentary or early diagenetic events caused by fluid migration and oxidation/reduction of soluble species can be characterized by the nature of infilling cement, precipitates, secondary crystallizations, etc. These secondary products provide a window into the pH/Eh, etc., conditions that were present when diagenetic processes were active. Figure 9, for example, shows a quartz sandstone decorated by iron oxide crystals (identified by XRD to be goethite, in all likelihood a pseudomorphic replacement of primary hematite) from Spitzbergen, Norway. The identification of secondary crystallization of Fe-oxide phases points to the early influx of soluble Fe++ and its oxidation and precipitation in place. Taphonomic alteration in this case would have likely destroyed evidence of the provenance of the sediment or of original habitability.

Figure 9. Goethite crystals on quartz sandstone (Spitzbergen, Norway). (a) Hand specimen, scale bar = 1 cm; (b) RGB element map from EDAX-Orbis instrument, Red = Fe, Green = Si, Blue = Ca; (c) RGB element map from MapX-III instrument, Red = Fe, Green = Si; (d) Fe map, MapX-III; (e) Si map, MapX-III; (f–h) ROI maps from MapX-III; (i) XRF spectra from ROI-1 and ROI-2, MapX-III (total beam power = 9 Wh (120 min analysis)).

Over the course of a 10-Earth year, 30 km transect across the central mound of Gale crater (Aeolus Mons, informally known as Mt. Sharp), the MSL Curiosity rover has identified hundreds of examples of secondary concretions, crystals and crystal pseudomorphs in the ancient sediments of Gale lake. Knowledge of the compositions of these features will allow an evaluation of the nature of the taphonomic changes experienced by the sediments and ultimately the potential for preservation of evidence of habitability. Bristow et al. (Reference Bristow, Grotzinger, Rampe, Cuadros, Chipera, Downs, Fedo, Frydenvang, McAdam, Morris, Achilles, Blake, Castle, Craig, Des Marais, Downs, Hazen, Ming, Morrison, Thorpe, Treiman, Tu, Vaniman, Yen, Gellert, Mahaffy, Wiens, Byrk, Bennett, Fox, Milliken, Fraeman and Vasavada2021), for example, describe Vera Rubin Ridge, an Fe-oxide rich prominent feature on lower Mt. Sharp, Gale crater, as depositionally equivalent to the immediately adjacent and clay-rich Glen Torridon region. Brine-driven destruction of clay minerals caused by silica-poor diagenetic fluids is proposed as the mechanism for destruction of the original mineralogy, and under these taphonomic conditions, any evidence of habitability present in the original mudstone would likely have been erased.

Discussion

Much of what we interpret is based on images. In the case of Mars, orbital imagery prior to the arrival of the Mars Express orbiter provided information that was largely morphological in nature. Fluvial, deltaic and lacustrine features and processes were identified based on their shapes and context, analogous to similar features and processes on Earth. With orbital remote data returned from the Observatoire pour la Minéralogie, l'Eau, les Glaces et l'Activité (OMEGA) instrument on the Mars Express orbiter (Bibring et al., Reference Bibring, Langevin, Gendrin, Gondet, Poulet, Berthe, Soufflot, Arvidson, Mangold, Mustard, Drossart, Erard, Forni, Combes, Encrenaz, Fouchet, Merchiorri, Belluci, Altieri, Formisano, Bonello, Capaccioni, Cerroni, Coradini, Fonti, Kottsov, Ignatiev, Moroz, Titov, Zasova, Pinet, Doutes, Schmitt, Sotin, Hauber, Hoffmann, Jaumann, Keller, Arvidson, Mustard, Duxbury and Forget2005, Reference Bibring, Langevin, Mustard, Poulet, Arvidson, Gendrin, Gondet, Mangold, Pinet, Forget, Berthe, Gendrin, Gomez, Gondet, Jouglet, Poulet, Soufflot, Vincendon, Combes, Drossart, Encrenaz, Fouchet, Merchiorri, Belluci, Alteieri, Formisano, Capaccioni, Cerroni, Coradini, Fonti, Korablev, Kottsov, Ignatiev, Moroz, Titov, Zasova, Loiseau, Pinet, Soute, Schmitt, Sotin, Hauber, Hoffmann, Jaumann, Keller, Arvidson, Mustard, Duxbury and Neukum2006, Reference Bibring, Arvidson, Gendrin, Gondet, Langevin, Lemouelic, Mangold, Morris, Mustard, Poulet, Quantin and Sotin2007), these features could be overlain with mineralogical/compositional information, resulting in the identification of clay minerals, sulphate minerals, iron oxide minerals, etc., in spatially resolved chemical features that in many cases validated and refined earlier morphology-based interpretations. The juxtaposition of clay mineral, sulphate mineral and oxide mineral layers in sedimentary sequences on a global scale was seen to correlate with a notional understanding of the history of the planet (‘Phyllosian’, loosely correlated with the Noachian, 4.5–3.6 Ga; ‘Theiikian’, loosely correlated with the Hesperian, 3.6–2.6 Ga; and ‘Siderikian’, loosely correlated with the Amazonian, 2.6 Ga–present) (Bibring et al., Reference Bibring, Langevin, Mustard, Poulet, Arvidson, Gendrin, Gondet, Mangold, Pinet, Forget, Berthe, Gendrin, Gomez, Gondet, Jouglet, Poulet, Soufflot, Vincendon, Combes, Drossart, Encrenaz, Fouchet, Merchiorri, Belluci, Alteieri, Formisano, Capaccioni, Cerroni, Coradini, Fonti, Korablev, Kottsov, Ignatiev, Moroz, Titov, Zasova, Loiseau, Pinet, Soute, Schmitt, Sotin, Hauber, Hoffmann, Jaumann, Keller, Arvidson, Mustard, Duxbury and Neukum2006; Grotzinger and Milliken, Reference Grotzinger, Milliken, Grotzinger and Milliken2012).

Optical in situ high-resolution images of planetary regolith (e.g. from the MAHLI imagers on MER Spirit and Opportunity and MSL Curiosity rovers) far exceed the lateral spatial resolution of companion elemental analysis instruments such as APXS (lateral spatial resolution about 2 cm), resulting in a similar quandary: depositional and diagenetic events can be inferred from optical images, but spatially resolved elemental images are needed to fully characterize them. This has been remedied to the ~125 micron level with the development of in situ X-ray fluorescence imaging and the arrival of the PIXL instrument at Mars (Allwood et al., Reference Allwood, Wade, Foote, Elam, Hurrowitz, Battel, Dawson, Denise, Ek, Gilbert, King, Liebe, Parker, Pedersen, Randall, Sharrow, Sondheim, Allen, Arnett, Au, Basset, Benn, Bousman, Braun, Calvet, Clark, Cinquini, Conaby, Conley, Davidoff, Delaney, Denver, Diaz, Doran, Ervin, Evans, Flannery, Gao, Gross, Grotzinger, Hannah, Harris, Harris, He, Heirwegh, Hernandez, Hertzberg, Hodyss, Holden, Hummel, Jadusingh, Jorgensen, Kawamura, Kitiyakara, Kozaczek, Lambert, Lawson, Liu, Luchik, Macneal, Madsen, McLennan, McNally, Meras, Muller, Napoli, Naylor, Nemere, Ponomarev, Perez, Pootrakul, Romero, Rosas, Sachs, Schaefer, Schein, Setterfield, Singh, Song, Soria, Stek, Tallarida, Thompson, Tice, Timmermann, Treiman, Tsai, Uckert, Villalvazo, Wang, Wilson, Worel, Zamani, Zappe, Zong and Zimmerman2020).

Microscale morphological information combined with elemental composition provides for a much improved interpretation of small-scale features. Principal igneous textures (flow structures, crystal fabrics) and primary sedimentary depositional fabrics (varves, ripples, etc.) can be more confidently interpreted, providing insight into modes of formation and the provenance of the source materials. Diagenetic features (concretions, secondary crystals, crystal pseudomorphs, vein fillings, alteration haloes, etc., can be used to interpret the timing and chemistry of secondary fluids that altered rocks and primary fabrics. In returned sample science as is being initiated by Mars 2020 Perseverance rover, the spatially resolved elemental chemistry of features in candidate rocks is useful in evaluating samples prior to caching. Microscale chemistry informs two important questions for sample triage prior to sample return: (1) Does the original environment, as evaluated on the basis of morphology and chemistry have the potential to harbour life, and (2) Does post-depositional diagenetic change (taphonomy) preserve or destroy such features (or, did the diagenetic environment itself support habitability)?

Not surprisingly, nature doesn't obey arbitrary size constraints or specific chemistries. There is nothing unique about 125 μm resolution or even 5 μm resolution in returned images, although improvements in spatial resolution will always allow for more confident interpretations. Nearly all of the sedimentary rocks imaged during MSL Curiosity's traverse in Gale crater had grain sizes smaller than fine sand (<125 μm) and were in many cases below MAHLI resolution (~15 μm). Fortunately, in many cases, compositionally distinct features (element images, ROI) can be compelling if not diagnostic at spatial scales larger than the dimensions of individual mineral grains. Images of rock fabric combined with elemental chemistry tell a compelling story.

The Mars Science Laboratory rover Curiosity identified the first habitable environment on ancient Mars in 2013 based on the mineralogy of the ancient lake sediment found in the Sheepbed member of the Yellowknife Bay formation (Grotzinger et al., Reference Grotzinger, Sumner, Kah, Stack, Gupta, Edgar, Rubin, Lewis, Schieber, Mangold, Milliken, Conrad, Des Marais, Farmer, Siebach, Calef III, Hurowitz, McLennan, Ming, Vaniman, Crisp, Vasavada, Edgett, Malin, Blake, Gellert, Mahaffy, Wiens, Maurice and Grant2013; Vaniman et al., Reference Vaniman, Bish, Ming, Bristow, Morris, Blake, Chipera, Morrison, Treiman, Rampe, Rice, Achilles, Grotzinger, McLennan, Williams, Bell III, Newsom, Downs, Maurice, Sarrazin, Yen, Morookian, Farmer, Stack, Milliken, Ehlmann and Sumner2013). The presence of significant quantities of clay minerals in the John Klein and Cumberland drill samples, key to habitability, was unexpected because they were not identified in orbital spectra. This has been the case for much of Curiosity's transit through more than 600 vertical meters of flat-lying sedimentary rocks; clay minerals and habitable environments were found through nearly the entire traversed section despite the absence of orbital evidence for clay minerals. There are two notable exceptions: aeolian sandstones that were porous and permeable in which water was apparently never present long enough to induce diagenetic change in the original basaltic sediments, and original basaltic sediments that were diagenetically altered by late-stage fluids to the extent that evidence of habitability (or even of provenance) has been erased (e.g. Bristow et al., Reference Bristow, Grotzinger, Rampe, Cuadros, Chipera, Downs, Fedo, Frydenvang, McAdam, Morris, Achilles, Blake, Castle, Craig, Des Marais, Downs, Hazen, Ming, Morrison, Thorpe, Treiman, Tu, Vaniman, Yen, Gellert, Mahaffy, Wiens, Byrk, Bennett, Fox, Milliken, Fraeman and Vasavada2021). High spatial resolution elemental images can identify and discriminate such occurrences.

Conclusions

In situ X-ray imaging instruments such as PIXL or MapX are capable of analysing the microstructure and elemental composition of rock types relevant to astrobiology rapidly and in many cases with minimal surface preparation. In searching for evidence of life or of habitable environments, the most important consideration aside from the nature of the original depositional environment is the degree of post-depositional taphonomic alteration (i.e. if evidence of life or conditions of habitability were originally present, to what extent were they preserved?). Over the past decade, MSL Curiosity has documented hundreds of post-depositional diagenetic features – veins, concretions, recrystallizations, alteration halos, etc., most in the mm-size range or larger that are useful in evaluating taphonomic change.

Dozens of geologically diverse and scientifically compelling landing sites have been identified and characterized from orbit as a result of the MER, MSL and Mars 2020 site selection workshops (Grant et al., Reference Grant, Golombeck, Parker, Crisp, Squyres and Weitz2004, Reference Grant, Golombek, Grotzinger, Wilson, Watkins, Vasavada, Griffes and Parker2011, Reference Grant, Golombek, Wilson, Farley, Williford and Chen2018). In order to characterize the full geological diversity of Mars, its early habitability and its potential to support in situ resource utilization (ISRU) and eventual human exploration, comprehensive in situ science investigations need to be performed at many diverse sites. A small fleet of MER-class rovers could serve such a purpose (Blake et al., Reference Blake, Walroth, Bristow, Sarrazin, Gailhanou, Downs and Thompson2021a), equipped with smaller, more powerful next-generation analytical instruments (e.g. Blake et al., Reference Blake, Zacny, Bristow, Morrison, Sarrazin, Rampe, Tu, Thorpe, Payre, Smith, Scudder, Bedford and Dehouck2021b; Rampe et al., Reference Rampe, Blake, Sarrazin, Bristow, Gailhanou, Lafuente, Tu, Zacny and Downs2021). If future human exploration of Mars is to be contemplated, a comprehensive understanding of the diversity of its many environments is a necessary prerequisite.

Supplementary material

The supplementary material for this article can be found at https://doi.org/10.1017/S147355042400003X

Data availability statement

The data that support the findings of this study are openly available in the Planetary X-ray Fluorescence Petrography database: https://doi.org/10.48484/R7WV-3T89

Financial support

This work was supported by awards from National Aeronautics and Space Administration (NASA) Research Opportunities in Space and Earth Science (ROSES) Planetary Instrument Concepts for the Advancement of Solar System Observations (PICASSO) program NNH13ZDA001N-PICASSO, award #13-PICASSO13-0111 and Maturation of Instruments for Solar System Exploration (MatISSE) program NNH16ZDA001N-MATISSE, award #16-MATISE16_2-0005.

Competing interest

None.

References

Allwood, AC, Wade, LA, Foote, MC, Elam, WT, Hurrowitz, JA, Battel, S, Dawson, DE, Denise, RW, Ek, EM, Gilbert, MS, King, ME, Liebe, CC, Parker, T, Pedersen, DAK, Randall, DP, Sharrow, RF, Sondheim, ME, Allen, G, Arnett, K, Au, MH, Basset, C, Benn, M, Bousman, JC, Braun, D, Calvet, RJ, Clark, B, Cinquini, L, Conaby, S, Conley, HA, Davidoff, S, Delaney, JH, Denver, T, Diaz, E, Doran, GB, Ervin, J, Evans, M, Flannery, DO, Gao, N, Gross, J, Grotzinger, J, Hannah, B, Harris, JT, Harris, CM, He, Y, Heirwegh, CM, Hernandez, C, Hertzberg, E, Hodyss, RP, Holden, JR, Hummel, C, Jadusingh, C, Jorgensen, JL, Kawamura, JH, Kitiyakara, A, Kozaczek, K, Lambert, JL, Lawson, PR, Liu, Y, Luchik, T, Macneal, KM, Madsen, SN, McLennan, SM, McNally, P, Meras, PL, Muller, RE, Napoli, J, Naylor, BJ, Nemere, P, Ponomarev, I, Perez, RM, Pootrakul, N, Romero, RA, Rosas, R, Sachs, J, Schaefer, RT, Schein, ME, Setterfield, TP, Singh, V, Song, E, Soria, MM, Stek, PC, Tallarida, NR, Thompson, DR, Tice, MM, Timmermann, L, Treiman, A, Tsai, S, Uckert, K, Villalvazo, J, Wang, M, Wilson, DW, Worel, SC, Zamani, P, Zappe, M, Zong, F and Zimmerman, R (2020) PIXL: Planetary Instrument for X-ray Lithochemistry. Space Science Reviews 216, 134268. https://doi.org/10.1007/s11214-020-00767-7.CrossRefGoogle Scholar
Anderson, RC, Jandura, L, Okon, AB, Sunshine, D, Roumeliotis, C, Beegle, LW, Hurrowitz, J, Kennedy, B, Limonadi, D, McCloskey, S, Robinson, M, Seybold, C and Brown, K (2012) Collecting samples in Gale crater, Mars; an overview of the Mars Science Laboratory sample acquisition, sample processing and handling system. Space Science Reviews 170, 5775. http://doi.org/10.1007/s11214-012-9898-9.CrossRefGoogle Scholar
Awramik, SM (1976) Gunflint stromatolites: microfossil distribution in relation to stromatolite morphology. In Walter, MR (ed.), Stromatolites. Amsterdam: Elsevier, pp. 311320.CrossRefGoogle Scholar
Barghoorn, ES and Tyler, SA (1965) Microorganisms from the Gunflint Chert. Science 147, 563575.CrossRefGoogle ScholarPubMed
Barnes, I and O'Neil, JR (1969) The relationship between fluids in some fresh alpine-type ultramafics and possible modern serpentinization, Western United States. GSA Bulletin 80, 19471960.CrossRefGoogle Scholar
Barnes, I, LaMarche, VC Jr and Himmelberg, G (1967) Geochemical evidence of present-day serpentinization. Science 156, 830832.CrossRefGoogle ScholarPubMed
Bibring, J-P, Langevin, L, Gendrin, A, Gondet, B, Poulet, F, Berthe, M, Soufflot, A, Arvidson, R, Mangold, N, Mustard, J, Drossart, P, Erard, S, Forni, O, Combes, M, Encrenaz, T, Fouchet, T, Merchiorri, R, Belluci, GC, Altieri, F, Formisano, F, Bonello, G, Capaccioni, F, Cerroni, P, Coradini, A, Fonti, S, Kottsov, V, Ignatiev, N, Moroz, V, Titov, D, Zasova, L, Pinet, P, Doutes, S, Schmitt, B, Sotin, C, Hauber, E, Hoffmann, H, Jaumann, R, Keller, U, Arvidson, R, Mustard, J, Duxbury, T and Forget, F (2005) Mars surface diversity as revealed by the OMEGA/Mars express observations. Science 307, 15761581. https://doi.org/10.1126/science.1108806.CrossRefGoogle ScholarPubMed
Bibring, J-P, Langevin, Y, Mustard, JF, Poulet, F, Arvidson, R, Gendrin, A, Gondet, B, Mangold, N, Pinet, P, Forget, F, Berthe, M, Gendrin, A, Gomez, C, Gondet, B, Jouglet, D, Poulet, F, Soufflot, A, Vincendon, M, Combes, M, Drossart, P, Encrenaz, T, Fouchet, T, Merchiorri, R, Belluci, G, Alteieri, F, Formisano, V, Capaccioni, F, Cerroni, P, Coradini, A, Fonti, S, Korablev, O, Kottsov, V, Ignatiev, N, Moroz, V, Titov, D, Zasova, L, Loiseau, D, Pinet, P, Soute, S, Schmitt, B, Sotin, C, Hauber, E, Hoffmann, H, Jaumann, R, Keller, U, Arvidson, R, Mustard, JF, Duxbury, T and Neukum, G (2006) Global mineralogical and aqueous Mars history derived from OMEGA/Mars express data. Science 312, 400404. https://doi.org/10.1126/science.1122659.CrossRefGoogle ScholarPubMed
Bibring, J-P, Arvidson, RE, Gendrin, A, Gondet, B, Langevin, Y, Lemouelic, S, Mangold, N, Morris, RV, Mustard, JF, Poulet, F, Quantin, C and Sotin, C (2007) Coupled ferric oxides and sulfates on the Martian surface. Science 317, 12061210. https://doi.org/10.1126/science.1144174.CrossRefGoogle ScholarPubMed
Bish DL, , Blake, DF, Vaniman, DT, Chipera, SJ, Morris, RV, Ming, DW, Treiman, AH, Sarrazin, P, Morrison, SM, Downs, RT, Achilles, CN, Yen, AS, Bristow, TF, Crisp, JA, Rampe, EB, Stolper, EM, Spanovich, N and Morookian, JM (2013) X-ray diffraction results from Mars Science Laboratory: mineralogy of Rocknest aeolian bedform at Gale crater. Science 341, 1238932. https://doi.org/10.1126/science.1238932.CrossRefGoogle ScholarPubMed
Bishop, JL, Parente, M, Weitz, CM, Noe Dobrea, EZ, Roach, LH, Murchie, SL, McGuire, PC, McKeown, NK, Rossi, CM, Brown, AJ, Calvin, WM, Milliken, R and Mustard, JF (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 114, E00D09. https://doi.org/10.1029/2009JE003352.CrossRefGoogle Scholar
Blake, DF and Peacor, DR (1985) TEM/STEM microanalysis of Holocene fresh-water Magnesian carbonate cements from the coast range of California. American Mineralogist 70, 388394.Google Scholar
Blake, DF, Treiman, AH, Morris, R, Bish, D, Amundsen, HEF and Steele, A (2011) ‘Carbonate cements from the Sverrefjell and Sigurdfjell Volcanoes, Svalbard, Norway: analogs for Martian carbonates.’ LPSC 42, The Woodlands, TX, March 7-11, 2011. Abstr. #2167, Available at https://www.lpi.usra.edu/meetings/lpsc2011/pdf/2167.pdfGoogle Scholar
Blake, DF, Sarrazin, P and Bristow, T (2015) ‘Mapping X-ray Fluorescence Spectrometer (Map-X).’ LPSC 46, The Woodlands, TX, March 16-20, 2015. Abstr. #2274. Available at https://www.hou.usra.edu/meetings/lpsc2015/pdf/2274.pdfGoogle Scholar
Blake, DF, Walroth, R, Bristow, T, Sarrazin, P, Gailhanou, M, Downs, R and Thompson, K (2021a) MapX: an in-situ X-ray μ-mapper for habitability and biosignature studies. Bulletin of the American Astronomical Society 53, 15. https://doi.org/10.3847/25c2cfeb.d0806478.Google Scholar
Blake, DF, Zacny, K, Bristow, T, Morrison, S, Sarrazin, P, Rampe, E, Tu, V, Thorpe, MT, Payre, V, Smith, R, Scudder, N, Bedford, CC and Dehouck, E (2021b) MER-class rover investigations of Mars in the coming decades. Bulletin of the American Astronomical Society 53, 15. https://doi.org/10.3847/25c2cfeb.a7226c13.Google Scholar
Blake, DF, Vaniman, D, Achilles, C, Anderson, R, Bish, D, Bristow, T, Chen, C, Chipera, S, Crisp, J, Des Marais, DJ, Downs, RT, Farmer, J, Feldman, S, Fonda, M, Gailhanou, M, Ma, H, Ming, DW, Morris, RV, Sarrazin, P, Stolper, E, Treiman, A and Yen, A (2012) Characterization and calibration of the CheMin mineralogical instrument on Mars Science Laboratory. Space Science Reviews 170, 5775.CrossRefGoogle Scholar
Blake, DF, Morris, RV, Kocurek, G, Morrison, SM, Downs, RT, Bish, DL, Ming, DW, Edgett, KS, Rubin, D, Goetz, W, Madsen, MB, Sullivan, R, Gellert, R, Campbell, I, Treiman, AH, McClennan, SM, Yen, AS, Grotzinger, J, Vaniman, DT, Chipera, SJ, Achilles, CN, Rampe, EB, Sumner, D, Meslin, P-Y, Maurice, S, Forni, O, Gasnault, O, Fisk, M, Schmidt, M, Mahaffy, P, Leshin, LA, Glavin, D, Steele, A, Freissinet, C, Navarro-Gonzalez, R, Yingst, RA, Kah, LC, Bridges, N, Lewis, KW, Bristow, TF, Farmer, JD, Crisp, JA, Stolper, EM, Des Marais, DJ and Sarrazin, P and the Mars Science Laboratory Science Team (2013) Curiosity at Gale crater, Mars: characterization and analysis of the Rocknest sand shadow. Science 341, 1239505.CrossRefGoogle ScholarPubMed
Bridges, JC and Grady, MM (2000) Evaporite mineral assemblages in the Nakhlite (Martian) meteorites. Earth and Planetary Science Letters 176, 267279. https://doi.org/10.1016/S0012-821X(00)00019-4.CrossRefGoogle Scholar
Bridges, JC and Schwenzer, SP (2012) The Nakhlite hydrothermal brine on Mars. Earth and Planetary Science Letters 359–360, 117123. https://doi.org/10.1016/S0012-821X(00)00019-4.CrossRefGoogle Scholar
Bridges, JC, Hicks, LJ and Treiman, AH (2018) Carbonates on Mars. In Filiberto, J and Schwenzer, SP (eds), Volatiles in the Martian Crust. Elsevier, pp. 89118. http://dx.doi.org/10.1016/B978-0-12-804191-8.00005-2.Google Scholar
Bristow, TF, Bish, DL, Vaniman, DT, Morris, RV, Blake, DF, Grotzinger, JP, Rampe, EB, Crisp, JA, Achilles, CN, Ming, DW, Ehlmann, DL, King, PL, Bridges, JC, Eigenbrode, JL, Sumner, DY, Chipera, J, Moorokian, JM, Treiman, AH, Morrison, SM, Downs, RT, Farmer, JD, Des Marais, DJ, Sarrazin, P, Floyd, MM, Mishna, MA and McAdam, AC (2015) The origin and implications of clay minerals from Yellowknife Bay, Gale crater, Mars. American Mineralogist 100, 824836. https://doi.org/10.2138/am-2015-5077CCBYNCND.CrossRefGoogle ScholarPubMed
Bristow, TF, Grotzinger, JP, Rampe, EB, Cuadros, J, Chipera, SJ, Downs, GW, Fedo, CM, Frydenvang, J, McAdam, AC, Morris, RV, Achilles, CN, Blake, DF, Castle, N, Craig, P, Des Marais, DJ, Downs, RT, Hazen, RM, Ming, DW, Morrison, SM, Thorpe, MT, Treiman, AH, Tu, V, Vaniman, DT, Yen, AS, Gellert, R, Mahaffy, PR, Wiens, RC, Byrk, AB, Bennett, KA, Fox, VK, Milliken, RE, Fraeman, AA and Vasavada, AR (2021) Brine-driven destruction of clay minerals in Gale crater, Mars. Science 373, 198204. https://doi.org/10.1126/science.abg5449.CrossRefGoogle ScholarPubMed
Campbell, JL, Perrett, GM, Gellert, R, Andrushenko, SM, Boyd, NI, Maxwell, JA, King, PL and Schofield, CD (2012) Calibration of the Mars Science Laboratory Alpha Particle X-ray Spectrometer. Space Science Reviews 170, 319340. https://doi.org/10.1007/s11214-012-9873-5.CrossRefGoogle Scholar
Chipera, SJ, Vaniman, DT, Rampe, EB, Bristow, TF, Martinez, G, Tu, VM, Peretyazhko, TS, Yen, AS, Gellert, R, Berger, JA, Rapin, W, Morris, RV, , DW, Thompson, LM, Simpson, S, Achilles, CN, Tutolo, B, Downs, RT, Fraeman, AA, Fischer, E, Blake, DF, Treiman, AH, Morrison, SM, Thorpe, MT, Gupta, S, Dietrich, WE, Downs, G, Castle, N, Craig, PI, Des Marais, DJ, Hazen, RM, Vasavada, AR, Hausrath, E, Sarrazin, P and Grotzinger, JP (2023) Mineralogical investigation of Mg-sulfate at the Canaima drill site, Gale crater, Mars. Journal of Geophysical Research: Planets 128, 116. https://doi.org/10.1029/2023JE008041.Google Scholar
Christensen, PR, Mehall, GL, Silverman, SH, Anwar, S, Cannon, G, Gorelick, N, Kheen, R, Tourville, T, Bates, D, Ferry, S, Fortuna, T, Jeffreys, J, O'Donnell, W, Peralta, R, Wolverton, T, Blaney, D, Denise, R, Radermacher, J, Morris, RV and Squyres, S (2003) Miniature Thermal Emission Spectrometer for the Mars Exploration Rovers. Journal of Geophysical Research: Planets 108, 8064. https://doi.org/10.1029/2003JE002117.CrossRefGoogle Scholar
Coleman, RG (2000) ‘Prospecting for ophiolites along the California Continental Margin.’ In Ophiolites and Oceanic Crust: New Insights from Field Studies and the Ocean Drilling Program, Geological Society of America Special Paper 349, edited by Y. Dilek, E. Moores, D. Elthon, and A. Nicolas, Geological Society of America, Boulder, CO, pp. 351364.CrossRefGoogle Scholar
Deming, JW and Baross, JA (1993) Deep-sea smokers: windows to a subsurface biosphere? Geochimica et Cosmochimica Acta 57, 32193230. http://doi.org/10.1016/0016-7073(93)90535-5.CrossRefGoogle ScholarPubMed
Economou, TE (2011) History of applications of radioactive sources in analytical instruments for planetary exploration. In Singh, N (ed.), Radioisotopes: Applications in Physical Sciences. InTech, pp. 155172. http://dx.doi.org/10.5772/23087.Google Scholar
Edgett, KS, Yingst, RA, Ravine, MA, Caplinger, MA, Ghaemi, FT, Maki, JN, Schaffner, JA, Bell, JF, Edwards, LJ, Herkenhoff, KE and Heydari, E (2012) Curiosity's Mars Hand Lens Imager (MAHLI) investigation. Space Science Reviews 170, 259317. https://doi.org/10.1007/s11214-012-9910-4.CrossRefGoogle Scholar
Ehlmann, BL and Edwards, CS (2014) Mineralogy of the Martian surface. Annual Review of Earth and Planetary Sciences 42, 291315. https://doi.org/10.1146/annurev-earth-060313-055024.CrossRefGoogle Scholar
Ehlmann, BL, Mustard, JF, Murchie, SL, Poulet, F, Bishop, JL, Brown, AJ, Calvin, WM, Clark, RN, Des Marais, DJ, Milliken, RE, Roach, LH, Rousch, TL, Swayze, GA and Wray, JJ (2008) Orbital identification of carbonate-bearing rocks on Mars. Science 322, 18281832. http://doi.org/10.1126/science.1164759.CrossRefGoogle ScholarPubMed
Farley, KA, Stack, KM, Shuster, DL, Horgan, BHN, Hurowitz, JA, Tarnas, JD, Simon, JI, Sun, VZ, Scheller, EL, Moore, KR, McLennan, SM, Vasconcelos, PM, Wiens, RC, Treiman, AH, Mayhew, LE and the Mars 2020 Science Team (2022) Aqueously altered igneous rocks sampled on the floor of Jezero crater, Mars. Science 377, 111. https://doi.org/10.1126/science.abo2196.CrossRefGoogle ScholarPubMed
Fraser, GW, Carpenter, JD, Rothery, DA, Pearson, JF, Martindale, A, Huovelin, J, Treis, J, Anand, M, Anttila, M, Ashcroft, M, Benkoff, J, Bland, P, Bowyer, A, Bradley, A, Bridges, J, Brown, C, Bullock, C, Bunce, EJ, Christensen, U, Evans, M and Whitehead, S (2010) The Mercury Imaging X-ray Spectrometer (MIXS) on Bepicolombo. Planetary and Space Science 58, 7995. https://doi.org/10.1016/j.pss.2009.05.004.CrossRefGoogle Scholar
Gailhanou, M, Sarrazin, P and Blake, DF (2018) Modeling of X-ray fluorescence full field imaging using planar square pore micro-channel plate optics. Journal Applied Optics 57, 67956807. https://doi.org/10.1364/AO.57.006795.CrossRefGoogle ScholarPubMed
Gellert, R, Rieder, R, Bruckner, J, Clark, BC, Dreibus, G, Klinglehofer, G, Lugmair, G, Ming, DW, Wanke, H, Yen, A, Zipfel, J and Squyres, SW (2006) Alpha Particle X-ray Spectrometer (APXS): results from Gusev crater and calibration report. Journal of Geophysical Research 111, 132. https://doi.org/10.1029/2005JE002555.CrossRefGoogle Scholar
Grant, JA, Golombeck, MP, Parker, TJ, Crisp, JA, Squyres, SW and Weitz, CM (2004) Selecting landing sites for the 2003 Mars Exploration Rovers. Planetary and Space Science 52, 1121. http://doi.org/10.1016/j.pss.2003.08.011.CrossRefGoogle Scholar
Grant, JA, Golombek, MP, Grotzinger, JP, Wilson, SA, Watkins, MM, Vasavada, AR, Griffes, JL and Parker, TJ (2011) The science process for selecting the landing site for the 2011 Mars Science Laboratory. Planetary and Space Science 59, 11141127. https://doi.org/10.1016/j.pss.2010.06.016.CrossRefGoogle Scholar
Grant, JA, Golombek, MP, Wilson, SA, Farley, KA, Williford, KH and Chen, A (2018) The science process for selecting the landing site for the 2020 Mars rover. Planetary and Space Science 164, 106126. https://doi.org/10.1016/j.pss.2018.07.001.CrossRefGoogle Scholar
Grotzinger, JP and Milliken, RE (2012) The sedimentary rock record of Mars: distribution, origins and global stratigraphy. In Grotzinger, JP and Milliken, RE (eds), Sedimentary Geology of Mars. https://doi.org/10.2110/pec.12.102.0001CrossRefGoogle Scholar
Grotzinger, JP, Crisp, J, Vasavada, AR, Anderson, RC, Baker, CJ, Barry, R, Blake, DF, Conrad, P, Edgett, KS, Ferdowski, B, Gellert, R, Gilbert, JB, Golombek, M, Gomez-Elvira, J, Hassler, DM, Jandura, L, Litvac, M, Mahaffy, P, Maki, J, Meyer, M, Mitrofanov, I, Simonds, JL, Vaniman, D, Welch, RV and Wiens, RC (2012) Mars Science Laboratory mission and science investigation. Space Science Reviews 170, 5775. https://doi.org/10.1007/s11214-012-9892-2.CrossRefGoogle Scholar
Grotzinger, JP, Sumner, DY, Kah, LC , Stack, K, Gupta, S, Edgar, L, Rubin, D, Lewis, K, Schieber, J, Mangold, N, Milliken, R, Conrad, D, Des Marais, D, Farmer, J, Siebach, K, Calef III, F, Hurowitz, J, McLennan, SM, Ming, D, Vaniman, D, Crisp, J, Vasavada, A, Edgett, KS, Malin, M, Blake, D, Gellert, R, Mahaffy, P, Wiens, RC, Maurice, S, Grant, J and the Mars Science Laboratory Science Team (2013) A habitable Fluvio-lacustrine environment at Yellowknife Bay, Gale crater, Mars. Science 343, 1242777. https://doi.org/10.1126/science.1242777Google ScholarPubMed
Herkenhoff, KE, Squyres, SW, Bell III, JF, Maki, JN, Arneson, HM, Bertelsen, P, Brown, DI, Collins, SA, Dingizian, A, Elliott, ST, Goetz, W, Hagerott, EC, Hayes, AG, Johnson, MJ, Kirk, RL, McLennan, S, Morris, RV, Scherr, LM, Schwochert, MA, Shiraishi, LR, Smith, GH, Soderblom, LA, Sohl-Dickstein, JM and Wadsworth, MV (2003) Athena Microscopic Imager investigation. Journal of Geophysical Research: Planets 108, 8065. https://doi.org/10.1029/2003JE002076.CrossRefGoogle Scholar
Herkenhoff, KE, Grotzinger, J, Knoll, AH, McLennan, SM, Weitz, C, Yingst, A, Anderson, R, Archinal, BA, Arvidson, RE, Barrett, JM, Becker, KJ, Bell III, JF, Budney, C, Chapman, MG, Cook, D, Ehlmann, B, Franklin, B, Gadis, LR, Galuszka, DM , Garcia, PA, Geissler, P, Hare, TM, Howington-Krause, E, Johnson, JR, Keszthelyi, L, Kirk, RL, Lanagan, P, Lee, EM, Leiff, C, Maki, JN, Mullins, KF, Parker, TJ, Redding, BL, Roseik, MR, Sims, MH, Soderblom, LA, Spanovich, N, Springer, R, Squyres, SW, Stolper, D, Sucharski, RM, Tucharski, T, Sullivan, R and Torson, JM (2008) Surface processes recorded by rocks and soils on Meridiani Planum, Mars: microscopic imager observations during Opportunity's first three extended missions. Journal of Geophysical Research 113, 139. https://doi.org/10.1029/2008JE003100.CrossRefGoogle Scholar
Hickman-Lewis, K, Moore, KR, Razzell-Hollis, JJ, Tuite, ML, Beegle, LM, Bhartia, R, Grotzinger, JP, Brown, AJ, Shkolyar, S, Cavalazzi, B and Smith, CL (2022) In situ identification of paleoarchean biosignatures using colocated Perseverance rover analyses: perspectives for in situ Mars science and sample return. Astrobiology 22, 11431163. https://doi.org/10.1089/ast.2022.0018.CrossRefGoogle ScholarPubMed
Kelley, DS, Karson, JA, Blackman, DK, Fruh-Green, GL, Butterfield, DA, Lilley, MD, Olson, EJ, Schrenk, MO, Roe, KK, Lebon, GT and Rivizzigno, P, (2001a) An off-axis hydrothermal vent field near the mid-Atlantic ridge at 30°N. Nature 412, 145149. https://doi.org/10.1038/35084000.CrossRefGoogle Scholar
Kelley, D, Karson, J, Blackman, D, Fruh-Green, G, Butterfield, D, Lilley, M, Schrenk, M, Olson, E, Roe, K and Lebon, J (2001b) Discovery of lost city: an off-axis, peridotite hosted, hydrothermal field near 30°N on the mid-Atlantic ridge. Ridge Events 11, 39.Google Scholar
Kelley, DS, Karson, JA, Fruh-Green, GL, Yoerger, DR, Shank, TM, Butterfield, DA, Hayes, JM, Schrenk, MO, Olson, EJ, Proskurowski, G, Jakuba, M, Bradley, A, Larson, B, Ludwig, K, Glickson, D, Buckman, K, Bradley, AS, Brazelton, WJ, Roe, K, Elend, MJ, Delacour, A, Bernasconi, SM, Lilley, MD, Baross, JA, Summons, RE and Sylvia, SP (2005) ‘A serpentinite-hosted ecosystem: the lost city hydrothermal field. Science 307, 14281434. https://doi.org/10.1126/science.1102556.CrossRefGoogle ScholarPubMed
Kirkland, DW, Dennison, RE and Dean, WE (2000) Parent brine of the Castile evaporites (Upper Permian), Texas and New Mexico. Journal of Sedimentary Research 70, 749761. https://doi.org/10.1306/2dc40935-0e47-11d7-8643000102c1865d.CrossRefGoogle Scholar
Knoll, AH (1985) Exceptional preservation of photosynthetic organisms in silicified carbonates and silicified peats. Philosophical Transactions of the Royal Society B 311, 111122.Google Scholar
Lafuente, B, Downs, RT, Yang, H and Stone, N (2015) The power of databases: the RRUFF project. In Armbruster, T and Danisi, RM (eds). Highlights in Mineralogical Crystallography. Berlin, Germany: W. DeGruyter, pp. 1130. https://doi.org/10.1515/9783110417104-003.Google Scholar
Liu, Y, Tice, MM, Schmidt, ME, Treiman, AH, Kizovski, TV, Hurowitz, JA, Allwood, AC, Henneke, J, Pedersen, DAK, VanBommel, SJ, Jones, MWM, Knight, AL, Orenstein, BJ, Clark, BC, Elam, WT, Heirwegh, CM, Barber, T, Beegle, LW, Benzerara, K, Bernard, S, Beyssac, O, Bosak, T, Brown, AJ, Cardarelli, EL, Catling, DC, Christian, JR, Cloutis, EA, Cohen, BA, Davidoff, S, Fairen, AG, Farley, KA, Flannery, DT, Galvin, A, Grotzinger, JP, Gupta, S, Hall, J, Herd, CDK, Hickman-Lewis, K, Hodyss, RP, Horgan, BHN, Johnson, JR, Jorgensen, JL, Kah, LC, Maki, JN, Mandon, L, Mangold, N, McCubbin, FM, McLennan, SM, Moore, K, Nachon, M, Nemere, P, Nothdurft, LD, Nunez, JI, O'Neil, L, Quantin-Nataf, CM, Sautier, V, Shuster, DL, Siebach, KL, Simon, JI, Sinclair, KP, Stack, KM, Steele, A, Tarnas, JD, Tosca, NJ, Uckert, K, Udry, A, Wade, LA, Weiss, BP, Wiens, RC, Williford, KH and Zorzano, M-P (2022) An olivine cumulate outcrop on the floor of Jezero crater, Mars. Science 377, 15131519. https://doi.org/10.1126/science.abo2757.CrossRefGoogle ScholarPubMed
Mahaffy, PR, Webster, CR, Cabane, M, Conrad, PG, Coll, P, Atreya, SK, Arvey, R, Barciniak, M, Benna, M, Bleacher, L, Brinkerhoff, WB, Eigenbrode, JL, Carignan, D, Cascia, M, Chalmers, RA, Dworkin, JP, Errigo, T, Everson, P, Franz, H, Farley, R, Feng, S, Frazier, G, Freissinet, C, Glavin, DP, Harpold, DN, Hawk, D, Holmes, V, Johnson, CS, Jones, A, Jordan, P, Kellogg, J, Lewis, J, Lyness, E, Malespin, CA, Martin, DK, Maurer, J, McAdam, AC, McLennan, D, Nolan, TJ, Noriega, M, Pavlov, AA, Prats, B, Raaen, E, Sheinman, O, Sheppard, D, Smith, J, Stern, JC, Tan, F, Trainer, M, Ming, DW, Morris, RV, Jones, J, Gunderson, C, Steele, A, Wray, J, Botta, O, Leshin, LA, Owen, T, Battel, S, Jakosky, BM, Manning, H, Squyres, S, Navarro-Gonzalez, R, McKay, CP, Raulin, F, Sternberg, R, Buch, A, Sorensen, P, Kline-Schoder, R, Coscia, D, Szopa, C, Teinturier, S, Baffes, C, Feldman, J, Flesch, G, Forouhar, S, Garcia, R, Keymeulen, D, Woodward, S, Block, BP, Arnet, K, Miller, R, Edmonson, C, Gorevan, S and Mumm, E (2012) The sample analysis at Mars investigation and instrument suite. Space Science Reviews 170, 401478. https://doi.org/10.1007/s11214-012-9879-z.CrossRefGoogle Scholar
Maurice, S, Wiens, RC, Saccoccio, M, Barraclough, B, Gasnault, O, Forni, O, Mangold, N, Baratoux, D, Bender, S, Berger, G, Bernadin, J, Berthe, M, Bridges, N, Blaney, D, Bouye, M, Cais, P, Clark, B, Clegg, S, Cousin, A, Cremers, D, Cros, A, DeFlores, L, Derycke, C, Dingler, B, Dromart, G, Dubois, B, Dupieux, M, Durand, E, d'Uston, L, Fabre, C, Faure, B, Gaboriaud, A, Gharsa, T, Herkenhoff, K, Kan, E, Kirkland, L, Kouach, D, Lacour, J-L, Langevin, Y, Lasue, J, LeMouelic, S, Lescure, M, Lewin, E, Limonadi, D, Manhes, G, Mauchien, P, McKay, C, Meslin, P-Y, Michel, Y, Miller, E, Newsom, HE, Orttner, G, Paillet, A, Pares, aL, Parot, Y, Perez, R, Pinet, P, Poitrasson, F, Quertier, B, Salle, B, Sotin, C, Sautter, V, Seran, H, Simmonds, JJ, Sirven, J-B, Stiglich, R, Striebig, N, Thocaven, J-J, Toplis, MJ and Vaniman, D (2012) The ChemCam instrument suite on the Mars Science Laboratory (MSL) rover: science objectives and mast unit description. Space Science Reviews 170, 95166. https://doi.org/10.1007/s11214-012-9912-2.CrossRefGoogle Scholar
McCollom, TM and Shock, EL (1997) Geochemical constraints on chemolithoautotrophic metabolism by microorganisms in seafloor hydrothermal systems. Geochimica et Cosmochimica Acta 61, 43754391. https://doi.org/10.1016/S0016-7037(97)00241-X.CrossRefGoogle ScholarPubMed
McKay, DS, Gibson, EK Jr, Thomas-Keptra, KL, Vali, H, Romanek, CS, Clemett, SJ, Chillier, XD, Maechling, CR and Zare, RN (1996) Search for past life on Mars: possible relict biogenic activity in Martian meteorite ALH84001. Science 273, 924930. https://doi.org/10.1126/science.273.5277.924.CrossRefGoogle Scholar
McLennan, SM (2003) Sedimentary silica on Mars. Geology 31, 15318. https://doi.org/10.1130/0091-7613(2003)031<0315:SSOM>2.0.CO;2.2.0.CO;2>CrossRefGoogle Scholar
McLennan, SM, Bell III, JF, Calvin, WM, Christensen, PR, Clark, BC, de Souza, PA, Farmer, J, Farrand, WH, Fike, DA, Gellert, R, Gosh, A, Glotch, TD, Grotzinger, JP, Hahn, B, Herkenhoff, KE, Hurowitz, JA, Johnson, JR, Johnson, SS, Joliff, B, Klingelhofer, G, Knoll, AH, Learner, Z, Malin, MC, McSween, HY Jr, Pocock, J, Ruff, SW, Soderblom, LA, Squyres, SW, Tosca, NJ, Watters, WA, Wyatt, MB and Yen, A (2005) Provenance and diagenesis of the evaporite bearing burns formation, Meridiani Planum, Mars. Earth and Planetary Science Letters 240, 95121. http://doi.org/10.1016/j.epsl.2005.09.041.CrossRefGoogle Scholar
Milliken, RE, Swayze, GA, Arvidson, RE, Bishop, JL, Clark, RN, Ehlmann, BL, Green, RO, Grotzinger, JP, Morris, RV, Murchie, SL, Mustard, JF and Weitz, C (2008) Opaline silica in young deposits on Mars. Geology 36, 847850. http://doi.org/10.1130/G24967A.1.CrossRefGoogle Scholar
Milliken, RE, Grotzinger, JP and Thomson, BJ (2010) Paleoclimate of Mars as captured by the stratigraphic record in Gale crater. Geophysical Research Letters 37, L04201. https://doi.org/10.1029/2009GL041870.CrossRefGoogle Scholar
Morris, RV, Klingelhofer, G, Schroder, C, Rodionov, DS, Yen, A, Ming, DW, de Souza, PA Jr, Fleischer, I, Wdowiak, T, Gellert, R, Bernhardt, B, Evlanov, EN, Zubkov, B, Foh, J, Bonnes, U, Kankeleit, E, Gutlich, P, Renz, F, Squyres, SW and Arvidson, RE (2006) Mössbauer mineralogy of rock, soil, and dust at Gusev crater, Mars: Spirit's journey through weakly altered olivine basalt on the plains and pervasively altered basalt in the Columbia Hills. Journal of Geophysical Research 111, E02S13. http://doi.org/10.1029/2005JE002584.CrossRefGoogle Scholar
Morris, RV, Ruff, SW, Gellert, R, Ming, DW, Arvidson, RE, Clark, BC, Golden, DC, Siebach, K, Klingelhofer, G, Schroder, C, Fleischer, I, Yen, A and Squyres, SW (2010) Identification of carbonate-rich outcrops on Mars by the Spirit rover. Science 29, 421424. https://doi.org/10.1126/science.1189667.CrossRefGoogle Scholar
Morris, RV, Blake, DF, Bish, D, Ming, DW, Agresti, DG, Treiman, AH, Steele, A and Amundsen, HEF (2011) ‘A terrestrial analogue from Spitsbergen (Svalbard, Norway) for the Comanche Carbonate at Gusev Crater, Mars.’ LPSC 42, Abstract # 1699. Available at https://www.lpi.usra.edu/meetings/lpsc2011/pdf/1699.pdf.Google Scholar
Morris, RV, Vaniman, DT, Blake, DF, Gellert, R, Chipera, SJ, Rampe, EB, Ming, DW, Morrison, SM, Downs, RT, Treiman, AH, Yen, AS, Grotzinger, JP, Achilles, CN, Bristow, TF, Crisp, JA, Des Marais, DJ, Farmer, JD, Fendich, KV, Frydenvang, J, Graff, TG, Morookian, JM, Stolper, EM and Schwenzer, SP (2016) Silicic volcanism on Mars evidenced by Tridymite in high-SiO2 sedimentary rock at Gale crater. Proceedings of the National Academy of Sciences of the USA 113, 70717076. https://doi.org/10.1073/pnas.1607098113.CrossRefGoogle ScholarPubMed
Pia, MG, Weidenspointer, G, Augelli, M, Quintieri, L, Saracco, P, Sudhakar, M and Zoglauer, A (2009) PIXE simulation with Geant4. IEEE Transactions on Nuclear Science 56, 36143649. https://doi.org/10.1109/TNS.2009.2033993.CrossRefGoogle Scholar
Price, GJ, Fraser, GW, Pearson, JF, Nussey, JP, Hutchinson, IB, Holland, AD, Turner, K and Pullan, P (2004) Prototype imaging X-ray fluorescence spectrometer based on microchannel plate optics. Review of Scientific Instruments 75, 23142319. http://doi.org/10.1063/1.764610.CrossRefGoogle Scholar
Rampe, EB, Ming, DW, Blake, DF, Bristow, TF, Chipera, SJ, Grotzinger, JP, Morris, RV, Morrison, SM, Vaniman, DT, Yen, AS, Achilles, CN, Craig, PI, Des Marais, DJ, Downs, RT, Farmer, JD, Fendrich, KV, Gellert, R, Morookian, JM, Peretyazhko, TS, Sarrazin, P, Treiman, AH, Berger, JA, Eigenbrode, JL, Fairen, AG, Forni, O, Gupta, S, Hurowitz, JA, Kah, LC, Lanza, NL, Schmidt, ME, Siebach, K, Sutter, B and Thompson, LC (2017) Mineralogy of an ancient lacustrine mudstone succession from the Murray formation, Gale crater, Mars. Earth and Planetary Science Letters 471, 172185. http://doi.org/10.1016/j.epsl.2017.04.021.CrossRefGoogle Scholar
Rampe, EB, Blake, D, Sarrazin, P, Bristow, T, Gailhanou, M, Lafuente, B, Tu, V, Zacny, K and Downs, R (2021) CheMinX: a next generation XRD/XRF for quantitative mineralogy and geochemistry on Mars. Bulletin of the American Astronomical Society 53. https://doi.org/10.3847/25c2cfeb.a4a55445Google Scholar
Razzell, HJ, Moore, KR, Sharma, S, Beegle, L, Grotzinger, JP, Allwood, A, Abbey, W, Bharita, R, Brown, AJ, Clark, B, Cloutis, E, Corpolongo, A, Henneke, J, Hickman-Lewis, K, Hurowitz, J, Jones, MWM, Liu, Y, Martinez-Frias, J, Murphy, A, Pedersen, DAK, Shkolyar, S, Siljestrom, S, Steele, A, Tice, M, Treiman, A, Uckert, K, VanBommel, S and Yanchilina, A (2022) The power of paired proximity science observations: co-located data from SHERLOC and PIXL on Mars. Icarus 387, 115179. http://doi.org/10.1016/j.icarus.2022.115179.CrossRefGoogle Scholar
Rieder, R, Gellert, R, Bruckner, J, Klingelhofer, G, Dreibus, G, Yen, A and Squyres, SW (2003) The new Athena Alpha Particle X-ray Spectrometer for the Mars Exploration Rovers. Journal of Geophysical Research: Planets 108, 8066. https://doi.org/10.1029/2003JE002150.CrossRefGoogle Scholar
Ruff, SW, Christensen, PR, Blaney, DL, Farrand, WH, Johnson, JR, Michaelski, JR, Moersch, JE, Wright, SP and Squyres, SW (2006) The rocks of Gusev crater as viewed by the Mini-TES instrument. Journal of Geophysical Research 111, E12S18. https://doi.org/10.1029/2006JE002747.CrossRefGoogle Scholar
Sarrazin, P, Blake, D, Thompson, K, Gailhanou, M, Chen, J and Bristow, T (2016) ‘The Map-X μ-imaging XRF spectrometer.’ LPSC 47, The Woodlands, TX, March 21-25, 2016. Abstr. #2883, Available at https://www.hou.usra.edu/meetings/lpsc2016/pdf/2883.pdfGoogle Scholar
Sarrazin, P, Blake, D, Gailhanou, M, Marchis, F, Chalumeau, C, Webb, S, Walter, P, Schyns, E, Thompson, K and Bristow, T (2018) MapX: 2D XRF for planetary exploration – image formation and optic characterization. Journal of Instrumentation 13, C04023. https://doi.org/10.1088/1748-0221/13/04/C04023.CrossRefGoogle Scholar
Scheller, EL, Hollis, JR, Cardarelli, EL, Steele, A, Beegle, L, Bharita, R, Conrad, P, Uckert, K, Sharma, S, Ehlmann, BL, Abbey, WJ, Asher, SA, Benison, KC, Berger, EL, Beyssac, O, Bleefeld, BL, Bosak, T, Brown, AJ (2023) Aqueous alteration processes in Jezero crater, Mars – implications for organic geochemistry. Science 378, 11051110. https://doi.org/10.1126/science.abo5204.CrossRefGoogle Scholar
Schoonjans, T, Brunetti, A, Golosio, B, del Rio, MS, Sole, VA, Ferrero, C and Vincze, L (2012) The Xraylib library for X-ray – matter interactions. Recent developments. Spectrochimica Acta Part B 70, 1023. https://doi.org/10.1016/j.sab.2011.09.011.CrossRefGoogle Scholar
Schulte, M, Blake, D, Hoehler, T and McCollom, T (2006) Serpentinization and its implications for life on the early Earth and Mars. Journal of Astrobiology 6, 364376. https://doi.org/10.1089/ast.2006.6.364.CrossRefGoogle ScholarPubMed
Sepúlveda, M, Pozzi-Escot, D, Falcon, RA, Bermeo, N, Lebon, M, Moulherat, C, Sarrazin, P and Walter, P (2020) Unraveling the polychromy and antiquity of the Pachacamac Idol, Pacific Coast, Peru. PLoS ONE 15, e0226244. https://doi.org/10.1371/journal.pone.0226244.CrossRefGoogle ScholarPubMed
Sleep, NH, Meibom, A, Fridriksson, T, Coleman, RG and Bird, DK (2004) H2-rich fluids from serpentinization: geochemical and biotic implications. Proceedings of the National Academy of Sciences of the USA 101, 1281812823. https://doi.org/10.1073/pnas.0405289101.CrossRefGoogle ScholarPubMed
Solé, VA, Papillon, E, Cotte, M, Walter, P and Susini, J (2007) A multiplatform code for the analysis of energy-dispersive X-ray fluorescence spectra. Spectrochimica Acta Part B 62, 6368. https://doi.org/10.1016/j.sab.2006.12.002.CrossRefGoogle Scholar
Squyres, SW, Grotzinger, JP, Arvidson, RE, Bell III, JF, Calvin, W, Christensen, PR, Farrand, WH, Herkenhoff, KE, Johnson, JR, Klingelhofer, G, Knoll, A, McLennan, SM, McSween, HY, Morris, RV, Rice, JW Jr, Rieder, R and Soderblom, LA (2006) In situ evidence for an ancient aqueous environment at Meridiani Planum, Mars. Science 06, 17091714. https://doi.org/10.1126/science.1104559.Google Scholar
Squyres, SW, Arvidson, RE, Ruff, S, Gellert, R, Morris, RV, Ming, DW, Crumpler, L, Farmer, JD, Des Marais, DJ, Yen, A, McClennan, SM, Calvin, W, Bell III, JF, Clark, BC, Wang, A, McCoy, TJ, Schmidt, ME and DeSouza, PA Jr (2008) Detection of silica-rich deposits on Mars. Science 320, 10631067. https://doi.org/10.1126/science.1155429.CrossRefGoogle ScholarPubMed
Thompson, KA, Blake, DF and Sarrazin, P (2017) ‘X-ray fluorescence source modeling for the MapX imaging spectrometer: rocky planets and ocean worlds.’ LPSC 48, The Woodlands, TX, March 20-24, 2017. Abstr. #1602, Available at https://www.hou.usra.edu/meetings/lpsc2017/pdf/1602.pdfGoogle Scholar
Tice, MM, Hurowitz, JA, Allwood, AC, Jones, MWM, Orenstein, BJ, Davidoff, S, Wright, AP, Pedersen, DAK, Henneke, J, Tosca, NJ, Moore, KR, Clark, BC, McLennan, SM, Flannery, DT, Steele, A, Brown, AJ, Zorzano, M-P, Hickman-Lewis, K, Liu, Y, VanBommel, SJ, Schmidt, ME, Kizovski, TV, Treiman, AH, O'Neil, L, Fairen, AG, Shuster, DL and Gupta, S (2022) Alteration history of Seitah formation rocks inferred by PIXL X-ray fluorescence, X-ray diffraction and multispectral imaging on Mars. Science Advances 8, 47. https://doi.org/10.1126/sciadv.abp9084.CrossRefGoogle ScholarPubMed
Tosca, JA and Knoll, AH (2009) Juvenile chemical sediments and the long term persistence of water at the surface of Mars. Earth and Planetary Science Letters 286, 379386. https://doi.org/10.1016/j.epsl.2009.07.004.CrossRefGoogle Scholar
Treiman, AH, Amundsen, HEF, Blake, DF and Bunch, T (2002) Hydrothermal origin for carbonate globules in Martian meteorite ALH84001: a terrestrial analogue from Spitsbergen (Norway). Earth and Planetary Science Letters 204, 323332. http://doi.org/10.1016/S0012-821X(02)00998-6.CrossRefGoogle Scholar
Treiman, AH, Robinson, KL, Blake, DF and Bish, D (2010) ‘Mineralogy determinations by CheMin XRD, tested on ultramafic rocks.’ LPSC 41, The Woodlands, TX, March 1–5, 2010. Abstr. #1472, Available at https://www.lpi.usra.edu/meetings/lpsc2010/pdf/1472.pdfGoogle Scholar
Treiman, AH, Morris, RV, Agresti, DG, Graff, TG, Achilles, CN, Rampe, EB, Bristow, TF, Blake, DF, Vaniman, DT, Bish, DL, Chipera, SJ, Morrison, SM and Downs, RT (2014) Ferrian saponite from the Santa Monica Mountains (California, USA, Earth): characterization as an analog for clay minerals on Mars with application to Yellowknife Bay in Gale crater. American Mineralogist 99, 22342250. https://doi.org/10.2138/am-2014-4763.CrossRefGoogle Scholar
Tyler, SA and Barghoorn, ES (1954) Occurrence of structurally preserved plants in Pre-Cambrian rocks of the Canadian Shield. Science 119, 606608.CrossRefGoogle ScholarPubMed
VanBommel, SJ, Gellert, R, Berger, JA, Campbell, JL, Thompson, LM, Edgett, KS, McBride, MJ, Minitti, ME, Prader, I and Boyd, NI (2016) Deconvolution of distinct lithology chemistry through oversampling with the Mars Science Laboratory Alpha Particle X-Ray Spectrometer. X-Ray Spectrometry 45, 155161. https://doi.org/10.1002/xrs.2681.CrossRefGoogle Scholar
Vaniman, DT, Bish, DL, Ming, DW, Bristow, TF, Morris, RV, Blake, DF, Chipera, SJ, Morrison, SM, Treiman, AH, Rampe, EB, Rice, M, Achilles, CN, Grotzinger, JP, McLennan, SM, Williams, J, Bell III, JF, Newsom, HE, Downs, RT, Maurice, S, Sarrazin, P, Yen, AS, Morookian, JM, Farmer, JD, Stack, K, Milliken, RE, Ehlmann, BL, Sumner, D and the Mars Science Laboratory Science Team (2013) Mineralogy of a mudstone at Yellowknife Bay, Gale crater, Mars. Science 343, 1243480. https://doi.org/10.1126/science.1243480Google ScholarPubMed
Walroth, RC, Sarrazin, P, Blake, DF, Thompson, K, Gailhanou, M, Bristow, T, Marchis, F, Downs, R, Webb, S and Chalumeau, C (2019) ‘MapX: full-field X-ray fluorescence imager for in-situ habitability and biosignature investigations.’ Ninth Intl. Conf. on Mars, Pasadena, CA, July 22–26, 2019. LPI Contribution No. 2089, id.6329. Available at https://ui.adsabs.harvard.edu/abs/2019LPICo2089.6329WGoogle Scholar
Walter, P, Sarrazin, P, Gailhanou, M, Herrouard, D, Verney, A and Blake, DF (2018) Full-field XRF instrument for cultural heritage: application to the study of a Caillebotte painting. X-Ray Spectrometry 48, 274281. https://doi.org/10.1002/xrs.2841.CrossRefGoogle Scholar
Figure 0

Figure 1. Schematic diagram of the MapX instrument. X-ray or radioisotope sources fluoresce the sample. Fluoresced X-ray photons characteristic of elements present in the sample are emitted over 2π steradians from their point of origin on the surface. Photons that enter the MPO within its acceptance angle are focused onto an energy-discriminating X-ray sensitive CCD in a 1:1 configuration with the sample. Hundreds to thousands of short exposures, each one a complete 2D X-ray image of the sample, are summed to produce an hdf5 data file.

Figure 1

Figure 2. (a). Cross-section of MapX-III prototype, showing sample position, X-ray tubes, MPO and CCD224 imager in an evacuated chamber. (b). MapX-III prototype. With the exception of the X-ray tubes, MapX-III is assembled from flight qualified or qualifiable components: two Newton Scientific M54 Au target transmission X-ray tubes; an e2v front illuminated, deep depleted frame transfer CCD-224 (legacy from CheMin, 600 × 600 40 μm pixels); and an Ir coated 1.5 mm thick MPO with 20 μm square channels and 6 μm thick walls. Samples in MapX-III are held in the same vacuum as the CCD, eliminating the need for a Be window and allowing for element detection down to Z = 11 (Na). MapX-III has a lateral spatial resolution of ~125 μm and an energy resolution of ~200 eV.

Figure 2

Figure 3. Ultramafic xenolith from Spitsbergen, Norway. (a) Optical image of a thin section stub of ultramafic xenolith AMASE08 UI-3; (b) RGB element map from commercial EDAX-Orbis instrument (50 μm resolution), Red = Fe, Green = Ca, Blue = Cr; (c) RGB element map from MapX-III (~125 μm resolution), same colour scheme as in (b); (d–h) MapX-III elemental maps; (i–l) ROI selected by element correlation cluster analysis from MapX-III hdf5 data cube; (m) ROI map from MapX-III (numbers refer to labelled ROI spectra shown in (o)); (n) ROI map from EDAX (note: colours for ROI-2 and ROI-3 are reversed); (o) summed XRF spectra from individual ROI; ROI 1 = clinopyroxene, ROI 2 = basalt matrix, ROI 3 = orthopyroxene, ROI 4 = chrome spinel (total MapX-III beam power = 9 Wh (120 min analysis)).

Figure 3

Figure 4. Thin section stub of a partially serpentinized cobble from Complexion Springs, California. (a) Optical image, white rectangle shows area imaged in MapX-III; (b) RGB element map from EDAX-Orbis instrument (50 μm resolution), Red = Fe, Green = Ca, Blue = Cr, white rectangle shows area imaged in MapX-III; (c) RGB element map from MapX-III (~125 μm resolution), same colour scheme as in b; (d–g) element images from MapX-III; (h) ROI selected by elemental correlation cluster analysis from EDAX hdf5 data cube (numbers refer to individual ROI also shown in (i) from the MapX-III instrument; (i) ROI selected by elemental correlation cluster analysis from MapX-III hdf5 data cube; (j) MapX-III XRF spectra from individual ROI, numbers refer to ROI labelled in (i) (total MapX-III beam power = 2.25 Wh (30 min analysis)).

Figure 4

Figure 5. Ultramafic xenolith AMASE UI-6, an orthopyroxenite comprised principally of orthopyroxene and olivine with minor chrome spinel and carbonate. (a) Petrographic thin section, small black rectangle shows area of thin section magnified in (f); (b) RGB element map from EDAX-Orbis instrument, Red = Fe, Green = Ca, Blue = Cr, large white rectangle shows area imaged in MapX-III; (c) RGB element map from MapX-III, colour scheme same as in (b); (d) Ca element image from EDAX instrument; (e) Ca element image from MapX-III instrument; (f) optical micrograph of portion of thin section outlined in figures a–e, g, h, showing carbonates (C) and other secondary mineralization surrounding chrome spinel (S), Opx – orthopyroxene, Ol = olivine; (g) ROI map from EDAX instrument; (h) ROI map from MapX-III instrument; (i) XRF spectra from MapX-III ROI labelled 1 and 3 in (h) (total beam power = 9 Wh (120 min analysis)).

Figure 5

Figure 6. Carbonate-cemented basalt breccia from Spitsbergen, Norway. (a) Petrologic thin section; (b) RGB element map from EDAX-Orbis instrument (50 μm resolution), Red = Fe, Green = Ca, Blue = Mg, white rectangle shows area imaged in MapX-III; (c) RGB element map from MapX-III, same colour scheme as b; (d–g) Element images from EDAX (l) and MapX-III (r); (h) ROI map from EDAX (l) and MapX-III (r) note that ROI 1 from MapX is blue while the same ROI in EDAX is yellow; (i) ROI XRF spectra for MapX-III, ROI labelled in (h) (total beam power = 9 Wh (120 min analysis)).

Figure 6

Figure 7. Gunflint Chert, Schrieber Beach locality stromatolite. (a) Polished thin section stub; (b) RGB element maps from EDAX-Orbis and MapX-III instruments, Red = Fe, Green = Ca, Blue = Mg; (c–f) comparison element images from EDAX (l) and MapX-III (r) for Si, Ca, Fe and Mn; (g–j) ROI maps from EDAX instrument; (k–m) ROI maps from MapX-III; (n) XRF spectra for EDAX ROI; (o) XRF spectra for MapX-III ROI (total beam power = 9 Wh (120 min analysis)).

Figure 7

Figure 8. Castile Formation evaporite, Permian basin, New Mexico. (a) Petrographic thin section; (b) RGB element map from EDAX-Orbis instrument, Red = S, Green = Ca, Blue = Mg; (c) RGB element map from MapX-III instrument, same colour scheme as b; (d–i) comparison element images from EDAX (l) and MapX-III (r) for S, Ca and Mg (total beam power = 9 Wh (120 min analysis)).

Figure 8

Figure 9. Goethite crystals on quartz sandstone (Spitzbergen, Norway). (a) Hand specimen, scale bar = 1 cm; (b) RGB element map from EDAX-Orbis instrument, Red = Fe, Green = Si, Blue = Ca; (c) RGB element map from MapX-III instrument, Red = Fe, Green = Si; (d) Fe map, MapX-III; (e) Si map, MapX-III; (f–h) ROI maps from MapX-III; (i) XRF spectra from ROI-1 and ROI-2, MapX-III (total beam power = 9 Wh (120 min analysis)).

Supplementary material: File

Blake et al. supplementary material

Blake et al. supplementary material
Download Blake et al. supplementary material(File)
File 1.4 MB