Download as pdf or txt
Download as pdf or txt
You are on page 1of 68

Thompson Lucy, M (Orcid ID: 0000-0002-5444-952X)

Berger Jeff, A (Orcid ID: 0000-0002-0380-4683)


Spray John, G. (Orcid ID: 0000-0003-3418-6862)
Fraeman Abigail, A. (Orcid ID: 0000-0003-4017-5158)
O'Connell-Cooper Catherine, D (Orcid ID: 0000-0003-4561-3663)
Schmidt Mariek, E. (Orcid ID: 0000-0003-4793-7899)
VanBommel Scott, J (Orcid ID: 0000-0002-6565-0827)
Gellert Ralf (Orcid ID: 0000-0001-7928-834X)
Yen Albert, S. (Orcid ID: 0000-0003-2410-0412)

APXS-derived compositional characteristics of Vera Rubin Ridge and


Murray formation, Gale crater, Mars: Geochemical implications for the
origin of the ridge.
L. M. Thompson1, J. A. Berger2, J. G. Spray1, A. A. Fraeman3, M. A. McCraig4, C. D.
O’Connell-Cooper1, M. E. Schmidt5, S. VanBommel6, R. Gellert4, A. Yen3, and N. I. Boyd4
1
Planetary and Space Science Centre, University of New Brunswick, Canada

2
NASA Johnson Space Center, Houston, USA

3
Jet Propulsion Laboratory, California Institute of Technology, Pasadena, USA

4
University of Guelph, Ontario, Canada

5
Brock University, Ontario, Canada

6
Washington University, St Louis, USA

Corresponding author: Lucy Thompson (lthompso@unb.ca)

This article has been accepted for publication and undergone full peer review but has not been
through the copyediting, typesetting, pagination and proofreading process which may lead to
differences between this version and the Version of Record. Please cite this article as doi:
10.1029/2019JE006319
©2020 American Geophysical Union. All rights reserved.
Key Points:

 Vera Rubin ridge is compositionally a continuation of the Murray formation

 Compositional trends cut across stratigraphy, are post-depositional, diagenetic and/or

products of later alteration

 APXS data indicate enhanced fluid flow, and Si, Al and Mn mobilization within VRR

Abstract

The resistant ~50 m thick Vera Rubin ridge (VRR) situated near the base of Mount

Sharp, Gale crater, Mars has been deemed a high priority science target for the Mars Science

Laboratory mission. This is because of 1) its position at the base of the 5 km layered strata of

Mount Sharp, and 2) the detection of hematite from orbit, indicating that it could be the site

of enhanced oxidation. The compositional data acquired by the Alpha Particle X-ray

spectrometer (APXS) during Curiosity’s exploration of VRR helps to elucidate questions

pertaining to the formation of the ridge. APXS analyses indicate that VRR falls within the

compositional range of underlying lacustrine mudstones, consistent with a continuation of

that depositional environment and derivation from a similar provenance. Lower Fe

concentrations for VRR compared to the underlying strata discounts the addition of large

amounts of hematite to the strata, either as cement or as detrital input. Compositional trends

associated with VRR cross-cut stratigraphy, indicating post-depositional processes. Higher Si

and Al, and lower Ti, Fe and Mn than the underlying mudstone, particularly within distinct

patches of gray/blue bedrock are consistent with the addition of Si and Al. Lateral and

vertical compositional variations, suggest enhanced element mobility and fluid flow (possibly

via multiple events) through VRR, increasing towards the top of the ridge, consistent with the

action of warm (~50-100 °C), locally acidic saline fluids as inferred from the mineralogy of

drilled samples.

©2020 American Geophysical Union. All rights reserved.


Plain Language Summary

Curiosity has explored the resistant Vera Rubin ridge (VRR) at the base of Mount

Sharp, Gale crater, Mars owing to 1) its position within the 5 km layered rocks of Mount

Sharp, which record changes in Mars environment through time, and 2) the detection of

hematite from orbit. The Alpha Particle X-ray spectrometer (APXS), measures the elemental

composition of rocks. APXS analyses indicate that VRR has a similar composition to

underlying mudstones, consistent with continued deposition in a lake. Lower iron discounts

the addition of large amounts of hematite, either holding together mineral grains as cement,

or as detrital grains. Other elemental trends cut across layering, indicating post-depositional

processes. Lateral and vertical compositional variations, suggest enhanced element mobility

and fluid flow (possibly via multiple events) through VRR, particularly at the top of the ridge

and within grey/blue patches of bedrock, consistent with the action of warm (~50-100 °C),

acidic saline fluids inferred from the mineralogy of drilled samples

1 Introduction

Vera Rubin ridge (VRR) is a ~6.5 km long, 200 m wide, ~50 m thick resistant feature

at the base of Mount Sharp, located in the center of Gale crater, Mars (Anderson & Bell,

2010; Figure 1). Spectral data acquired from Mars orbiters detect different mineral species

within the layered sequence of Mount Sharp. Hydrated phyllosilicate-bearing assemblages

are detected towards the base, through hydrated sulfates to anhydrous sulfates at the top,

indicating changing geochemical and environmental conditions (Anderson & Bell, 2010;

Bibring et al., 2006; Grotzinger et al., 2014; Milliken et al., 2010). VRR has been a high

priority science target for the Mars Science Laboratory (MSL) mission since landing because

of its association with a relatively strong hematite spectral signature from orbit (Anderson &

Bell, 2010; Fraeman et al., 2013), its distinct topography, and its position within the

lowermost strata of Mount Sharp. It is situated adjacent to the main phyllosilicate-bearing

©2020 American Geophysical Union. All rights reserved.


trough within Gale crater (the clay-bearing unit, informally named Glen Torridon; Figure 1c),

also previously identified from orbit (Milliken et al., 2010). The detection of hematite within

VRR from orbit led Fraeman et al. (2013) to hypothesize that the ridge may preserve a

localized redox interface. The hematite would have formed when Fe2+ dissolved in anoxic

groundwaters was then exposed to an oxidizing environment (e.g., atmosphere or

groundwater), thereby precipitating insoluble Fe3+. This could have occurred in a lake

environment, syn-depositionally within the water column, or during later diagenesis or

alteration. The hematite could also be the result of in situ weathering of precursor Fe2+-

bearing silicates under oxidizing conditions (Fraeman et al., 2013). On Earth, iron oxidation

at chemical interfaces is favorable for chemolithotrophic microorganisms. Therefore, the

ridge is an excellent site to explore for signs of past habitability (Fraeman et al., 2013). The

hematite could also comprise a clastic component of the sedimentary strata forming VRR,

perhaps indicating a provenance distinct from the underlying Murray formation. The Murray

formation is the lower-most unit of the Mount Sharp group, first encountered at Pahrump

Hills (Figure 1c), and is interpreted to represent a primarily lacustrine and marginal lacustrine

depositional environment (Fedo et al., 2018, 2019; Grotzinger et al., 2015; Stack et al., 2019).

A dedicated campaign to study VRR using the full complement of Curiosity’s

instrument payload was planned to address the overarching questions: (1) What was the

primary depositional environment and what is the stratigraphic relationship of the ridge to

Mount Sharp? (2) Was the hematite primary or secondary? (3) What is its relationship to

hematite previously detected in the Murray formation? (4) What were the primary and

secondary geochemical environments? (5) What are the implications for habitability and the

preservation of organic molecules?

The compositional data acquired by the Alpha Particle X-ray Spectrometer (APXS)

during the >1 year-long campaign to study the ridge helps to address some of these questions.

©2020 American Geophysical Union. All rights reserved.


How does the composition of VRR compare with the underlying Murray formation and other

rocks encountered within Gale crater? Do the strata comprising the ridge constitute a

continuation of the Murray formation and how do their compositions compare to the rest of

the Murray formation? How does the chemistry vary within VRR, both laterally and with

elevation? What might this reveal about the formation of the ridge? Are there obvious

diagenetic features and, if so, how does their composition compare with VRR host rock and

diagenetic features encountered within the Murray formation below the ridge? This work is

also the first comprehensive description of the Murray formation compositional

characteristics up to and including VRR, thus providing important context for future studies.

2 Materials and Methods

All ten of the instruments on Curiosity’s payload were utilized during the campaign to

investigate VRR (see Fraeman et al., this issue a). These included the navigation and science

cameras, as well as internal and external geochemical and mineralogical instruments. The

internal analytical laboratory instruments are the Sample Analysis at Mars (SAM; Mahaffy et

al., 2012), which among other things, measures the organic molecule and light element

content of a sample, and the Chemistry and Mineralogy, X-ray diffraction instrument

(CheMin; XRD; Blake et al., 2012), which determines the mineralogy. The Chemistry and

Camera (ChemCam) instrument on the rover’s remote sensing mast uses Laser-Induced

Breakdown Spectroscopy (LIBS) to measure the chemistry of small, ~300 μm size spots, on

rocks and in soils, as well as a Remote Micro-Imager (RMI) to take close-up, monochromatic

images (Maurice et al., 2012). The Mast Camera (Mastcam; Malin et al., 2017) and Mars

Hand Lens Imager (MAHLI; Edgett et al., 2012) are utilized extensively to provide important

context imagery for APXS targets. Grotzinger at al. (2012) provide further details on

Curiosity’s instrument payload.

©2020 American Geophysical Union. All rights reserved.


2.1. Alpha Particle X-ray Spectrometer

The Mars Science Laboratory, Curiosity rover is equipped with a Canadian-built

Alpha Particle X-ray Spectrometer (APXS, Gellert et al., 2015) positioned on the end of the

robotic arm (Figure S2). It is the third generation APXS, and fourth instrument to land on the

surface of Mars (following the Sojourner and the Mars Exploration Rovers). The APXS

obtains quantitative major, minor and some trace element compositional data for rocks and

unconsolidated materials. It combines Particle Induced X-ray Emission (PIXE) and X-ray

Fluorescence (XRF) techniques, inducing characteristic X-rays from atoms within the target

of interest (Gellert et al., 2015). The APXS signal is derived from a 1.5 cm diameter field of

view (FOV) when in contact with a solid target. The FOV increases with higher standoff

distances from the target (VanBommel et al., 2017). APXS acquires acceptable data up to a

~2 cm standoff distance from a target (~3 cm diameter FOV), provided sufficient acquisition

time is allocated (e.g., VanBommel et al., 2019). Because the APXS results are derived from

the sum of X-ray signals within the FOV, it acquires a bulk composition from that FOV. The

1.5 – 2 cm diameter APXS FOV is comparable to the 1.6 cm diameter of the holes drilled by

Curiosity; thus the APXS provides compositional data on a similar areal scale as drilled

samples.

APXS acquires the highest quality spectral data with overnight (cold temperature),

nominally ~8 hour, in contact integrations, on preferably dust-free surfaces. However,

acceptable quality data can be obtained from integration times as short as 10 minutes during

the early morning or evening, provided the sensor head temperatures are sufficiently cool

(Gellert et al., 2015; VanBommel et al, 2019). Curiosity’s robotic arm is also equipped with a

Dust Removal Tool (DRT) that brushes dust from rock surfaces prior to MAHLI and APXS

analysis. The DRT is not used on every target; it’s use is often precluded by surface

roughness and other operational constraints. Dust on a surface will have some effect on the

©2020 American Geophysical Union. All rights reserved.


APXS data, particularly for the lighter, lower atomic number elements (Na through to Si),

whose signals are derived mostly from the uppermost 2 – 5 μm of a sample (Reider et al.,

2003). However, elemental trends for dusty targets, even for the lighter elements, are still

detectable (Schmidt et al., 2018). Further details of the APXS methodology and calibration

are presented in Campbell et al. (2012, 2014) and Gellert et al. (2006, 2015).

The chemistry of features of interest smaller than the APXS FOV can be investigated

using a “raster” technique. A number of overlapping APXS measurements, covering both the

feature(s) of interest (e.g., a nodule) and the substrate, are acquired. The chemistry of distinct

chemical endmembers for the feature(s) of interest are then derived at a sub-cm scale

(VanBommel et al., 2016, 2017).

The APXS has successfully recorded a diversity of compositions along the >20 km

rover traverse (Figure 1c). Bradbury group and Siccar Point APXS analyses are available at

the Planetary Data System (see link in acknowledgements). See Table S1 for the

compositions of all Murray formation rock targets analyzed by APXS up to and including sol

2301 (sol – martian solar day = 24.6 hours) (431 analyses). 159 APXS analyses were

acquired on VRR (Figure 2; Table 1). The APXS cannot distinguish the oxidation states of

polyvalent elements. Hence, total iron and sufur are reported as FeO and SO3 respectively.

2.2. APXS data treatment

Unless otherwise stated, analyses used to examine trends and compute statistical

relationships etc., exclude those containing large (>0.5 cm) veins and diagenetic features.

Diagenetic and alteration features include resistant concretions, nodules, dendritic features,

ridges and fins; different colored (to the bedrock) halos adjacent to fractures or veins; and

different colored (to the bedrock) patches (Tables S1, 1)

©2020 American Geophysical Union. All rights reserved.


APXS data are plotted on a total alkali versus silica plot (TAS plot; Le Maitre et al.,

2006), with the caveat that the majority of rocks sampled by the APXS are sedimentary and

not igneous. The TAS plot has traditionally been used to display martian compositional data

(e.g., McSween, 2015; Mc Sween et al., 2009) and therefore facilitates comparison across

different datasets. The compositional fields on the TAS plot, do not imply that any of the

rocks are igneous, rather that their bulk compositions fall within these ranges. The Gale data

plotted are renormalized to S and Cl free (i.e., volatile-free) as has customarily been done for

comparing martian compositions (e.g., McSween, 2015; Mc Sween et al., 2009).

In order to investigate the correlation of Si, Al and Fe throughout the Murray, plots of

these element oxides ratioed to TiO2 are shown (see §S1) to account for closure associated

with the CaSO4 present in almost all targets (see §S2).

2.2.1. Statistical treatment of the Murray formation data

Statistical treatments were performed on the data after exclusion of the same targets

outlined above:

F-tests (Fisher-tests) were performed on the Bradbury group – Murray formation

below VRR (excluding high Si targets) pairing to determine whether they show equal or

unequal variances (§S1.1.1.). Variance is the square of the standard deviation from the mean.

F-tests and t-tests were also performed on the Murray formation below VRR – Stimson

formation pairing for SiO2 and K2O to determine variance and whether the two formations

are statistically different for those two oxides (§S1.1.1.).

Box and whisker plots were constructed for each oxide/element of the Murray

members (§S1.1.2.). These plots are used to compare the mean/median compositions of each

of the Murray formation members, to illustrate variance, and to identify outlier targets.

A comparison of the Murray members was performed with increasing elevation from

the base of the Murray formation, at Pahrump Hills, including a comparison with the

©2020 American Geophysical Union. All rights reserved.


gray/blue bedrock patches on VRR (§S1.1.3.). F-tests were first performed on each pairing to

determine whether they show equal or unequal variances. Based on these results, t-tests were

then performed on each pairing to determine whether the compositions of the two members

or groupings are statistically the same or different. F-tests were also performed on the Murray

formation below VRR (excluding high Si targets) - VRR, and the upper and lower Sutton

Island to statistically compare variances between each pair.

Pearson correlation coefficients were computed for each Murray member, as well as

the formation Murray below and on VRR to investigate elemental correlations (§S1.2). 5-

point moving averages were fitted to the elevation versus oxide/element plots to identify

compositional trends with elevation (§S1.3).

2.2.2. Comparison of Murray member compositions to mean Murray formation

A comparison of each of the Murray formation members relative to mean Murray formation

bedrock up to VRR was performed to better quantify specific elemental deviations from the

mean, i.e., Si, Al, Fe, Mn and Ti (§S1.4).

2.2.3. Comparison of Jura and gray/blue bedrock to mean Pettegrove Point

Calculations (Brimhall & Dietrich, 1987) were performed to compare the Jura and

gray/blue bedrock patches to the mean Pettegrove Point (§S1.5). These calculations were

performed based on the premise that all VRR strata were deposited in a similar environment

(Edgar et al., this issue). See §5.1 for further discussion.

3 Geological and Geochemical Context of rocks encountered by APXS prior to VRR

The sedimentary strata encountered within Gale crater have been subdivided into

three stratigraphic groups: the Bradbury, Mount Sharp and Siccar Point groups, based on

their sedimentological characteristics and to a lesser extent, orbital geomorphic attributes

(Figures 1c, 2, 3) (Grotzinger et al., 2015; Banham et al., 2018; Fraeman et al., 2016). The

Bradbury group includes all strata encountered from landing to arriving at the base of Mount

©2020 American Geophysical Union. All rights reserved.


Sharp (up to sol 722) and is interpreted as a series of fluvio-deltaic deposits (Edgar et al., this

issue; Grotzinger et al., 2014; Williams et al., 2013). The Mount Sharp group is dominated by

lacustrine mudstones of the Murray formation and is unconformably overlain by the eolian

Stimson formation, Siccar Point group, (Fraeman et al., 2016; Banham et al., 2018). The

following is a description of the geochemical characteristics of the Murray formation

encountered by Curiosity and analyzed by APXS prior to arrival at VRR. This provides

context and enables a comparison of the composition of VRR strata with the underlying

Murray formation and other rocks within Gale crater. As this is the first publication of the

majority of this data (Table S1), we include brief discussions of the implications of these

results.

3.1. Mount Sharp group, Murray formation

The Murray formation, the lowermost unit of the Mount Sharp group, comprises

mainly mudstone, with lesser siltstone and fine sandstone, as well as less common lenses and

discontinuous beds of coarser sandstone. Based on sedimentological observations made along

Curiosity’s traverse, the Murray formation has been further divided into a series of members

that with increasing elevation are: Pahrump Hills, Hartmann’s Valley, Karasburg, Sutton

Island and Blunts Point (see the stratigraphic column in Figure 3) (Edgar et al., this issue;

Fedo et al., 2018, 2019). Curiosity has traversed >250 m in elevation through the Murray

formation, from its base at Pahrump Hills through to the base of VRR. In contrast, Curiosity’s

traverse through the Bradbury group includes only ~70 m of elevation. The Bradbury group

include K-rich sandstones (including the Kimberley formation), basaltic mudstones and fine

sandstones (Yellowknife Bay formation), alkali-rich sandstones and conglomerates, and

alkalic float rocks, many of which are interpreted to be igneous (Cousin et al., 2017; Edwards

et al., 2017; Sautter, et al., 2015; Schmidt et al., 2014; Stolper et al., 2013; Thompson et al.,

2016) (Figure 4). Compositional and mineralogical differences within the Bradbury group are

©2020 American Geophysical Union. All rights reserved.


attributed to provenance (e.g., Thompson et al., 2016; Treiman et al., 2016), and/or physical

sedimentary processes (Siebach et al., 2017).

The chemistry of the Murray formation below VRR is relatively uniform over a

significant elevation range, in contrast to lithologies encountered within the Bradbury group

(Figure 4; Table S3) (McLennan et al., 2013; Schmidt et al., 2014; Thompson et al., 2016).

This, and the similar basaltic, igneous mineralogy (Achilles et al., 2020; Bristow et al., 2018;

Rampe et al., 2020, this issue) of the Murray formation below VRR (excluding the Pahrump

Hills targets), indicates a similar provenance for most of the Murray sediment.

In general, the Murray formation rocks have lower Ca, Mg, Mn, K and Na, and higher

Si and Al, with higher FeO/MnO than Bradbury group strata (Figure 4). Specifically, the

Murray formation mudstones and siltstones have a composition distinct from the Bradbury

group, basaltic Yellowknife Bay formation mudstones, trending to higher Si and Al, and to

lower Ca and to a lesser extent, Mg concentrations (Figures 4, S3). The Murray formation is

also characterized by higher K and Si concentrations (mean 0.84±0.13 wt% K2O and

48.39±3.23 wt% SiO2) than the basaltic composition Stimson formation sandstones (mean

0.42±0.60 wt% K2O, 43.72±1.19 wt% SiO2; Figure 4, Table S3) (Thompson et al., 2016). Mn

is low (majority of targets, <0.4 wt% MnO), and FeO/MnO is high (~100) relative to average

basaltic martian soil/sand (~50, O’Connell-Cooper et al., 2017, Figure 5), and other Gale

bedrock. The ~110 ppm Ge consistently detected in the Murray formation strata, is

significantly elevated compared to martian meteorites (1-3 ppm) and terrestrial basalts (~1.4

ppm) (Berger et al., 2017 and references therein).

The contrasting composition of the Sheepbed member, Yellowknife Bay formation

mudstones (Bradbury group), and the Murray formation mudstones, both interpreted to have

been deposited in primarily lacustrine environments, has variously been attributed to reflect,

(a) differences in provenance for the two mudstones (e.g., Bedford et al., 2019); (b) enhanced

©2020 American Geophysical Union. All rights reserved.


weathering at source and during transport of the Murray sediment, combined with differences

in lake water chemistry during deposition, e.g., a redox stratified lake (Hurowitz et al., 2017);

or (c) different (pH and redox) post-depositional diagenetic processes (Rampe et al., 2017,

Bristow et al 2018).

Despite the relative homogeneity of the Murray formation compared to the Bradbury

group, variations in composition with elevation are observed (Figures 5, 6, 7). In general, Si,

Al, Ti, Mn, Na, Cr and Zn concentrations decrease, and Fe, Ni, Ca, S and Cl contents

increase with elevation from Pahrump Hills through to the Karasburg member (Figures 5, 6).

From the Karasburg through to the Blunts Point member, below the VRR, mean Si, Al and

Na concentrations are statistically the same, while mean Fe, Mn and Zn concentrations

increase (Figure 6; Table S4a). Up to the base of VRR, Si and Al positively correlate, and Al

and Fe show a moderate negative correlation with one another (Figure 8; Table S7i). A

number of Pahrump Hills targets lie outside these trends.

Pahrump Hills member

The lowest elevation lithostratigraphic member of the Murray formation, encountered

at Pahrump Hills (~20 m thick), is chemically distinct from the stratigraphically higher

sections for the majority of the elements analyzed by APXS, e.g., higher Si, Al, Ti, Cr, Mn,

Na, Zn and P, and lower Fe, Mg, Ni and Cl (Figures 5, 6, 7; Table S4a, c). In particular, Al

concentrations deviate from the Al-Si trend line, characteristic of the rest of the Murray,

trending to higher Al contents (Figure 8a). Ratios of Al/Ti also trend higher than for most of

the Murray formation (Figure 8d). The elevated Al is also evident from Figure 7, showing

the markedly higher Al versus Si concentrations for the majority of Pahrump Hills targets

relative to mean Murray, as well as lower Fe.

The Pahrump Hills member also reveals significant compositional variance for a

number of elements including Ti, Al, Fe, Na, P, Ni and Zn (Figure 6, Table S4b). Mn, Ni and

©2020 American Geophysical Union. All rights reserved.


Zn abundances decrease and K increases with increasing elevation (Figures 5, 6). The

mineralogy determined by CheMin from holes drilled within the Pahrump Hills member is

also distinct from the overlying Murray formation to VRR (Bristow et al., 2018), dominated

by magnetite versus hematite, and containing trace olivine, quartz and fluorapatite.

Texturally similar, but lighter toned mudstones encountered slightly higher up in the

Pahrump Hills member (~7 m), have the highest SiO2 contents of any bedrock encountered

by APXS at Gale (e.g., Buckskin DRT and drilled fines with up to 74.43 wt%; Figures 4, 6,

7). The detection of tridymite in the drilled Buckskin sample has been proposed as evidence

of a clastic, silicic volcanic component to the mudstone at this location (Morris et al., 2016).

More recently Yen et al. (2017a) have proposed that the tridymite could have formed in situ,

precipitating from acidic hydrothermal solutions, and that the high Si within the Murray

formation might be generated by similar processes as those responsible for high-Si zones in

the Stimson formation (Yen et al., 2017b). The accompanying elevated Ti and depletions in

nearly every other element (also observed within the high-Si Stimson zones), support this

hypothesis (Figure 7). The elevated Al (as well as Mn, Ni, Zn) observed down section

(Figures 5, 6, 7) maybe the result of downward migration of these fluids and precipitation due

to changing pH (Rampe et al., 2017). Other models invoked to explain the high Si zones

within the Murray and Stimson formations are Si addition during sedimentation and

diagenesis (Hurowitz et al., 2017), or late stage mobilization of Si and subsequent deposition

(Frydenvang et al., 2017).

Isolated, raised, resistant nodular concretions (typically 1-4 cm across) are ubiquitous

throughout the basal Pahrump Hills member, and periodically higher up in the section, within

the Hartmann’s Valley member. They exhibit characteristically high Mg, S and Ni contents,

as well as elevated Cl and Br relative to nearby bedrock and are consistent with the addition

©2020 American Geophysical Union. All rights reserved.


of up to 12 wt% of a Mg-Ni-SO4 phase ((Table S1; Thompson et al., 2015a; Van Bommel et

al., 2016).

A complex vein system at Garden City (Kronyak et al., 2018), within the Pahrump Hills

member, reveals at least four compositional phases, distinct from the Mg-Ni-SO4 features

(Thompson et al., 2015b; Table S1): (1) white, CaSO4-bearing veins, (2) dark, high-Ca

material (Figure S1), with high Mn, Zn, Ni, and Ge (~650 ppm Ge; Berger et al., 2017),

intimately associated with the white CaSO4 vein material (VanBommel et al., 2017), (3) dark,

Mg-rich material (with elevated P and Ge), also associated with white CaSO4 veins, and (4)

resistant dark “fins”/plates with characteristically high Na, K, Fe and P, as well as elevated

Ge (~250 ppm Ge; Berger et al., 2017). The CaSO4 vein networks at Garden City, appear to

have disrupted and incorporated the high-Ca, and high-Mg material as inclusions (Kronyak et

al., 2018). This, accompanied by the association with distinct chemistry resistant “fins” at

Garden City, suggest multiple fluid flow events at this location (Berger, et al., 2017; Kronyak

et al., 2018; Thompson et al., 2015 b).

3.1.2. Hartmann’s Valley

Immediately overlying the Pahrump Hills member, the Hartmann’s Valley member (~

20 m thick) exhibits variability in Si, Mg, K and Cl concentrations (Figure 6, Table S4b). In

particular, Hartmann’s Valley has the most variable, and highest mean K2O concentration of

the Murray below VRR (0.92 wt%). Hartmann’s Valley has lower Si, Ti, Al, Mn, Na, P, Cr

and Zn and higher Fe, Mg, K, Ni, Ca, S and Cl than Pahrump Hills (Figures 5, 6; Table S4a,

c), with the highest mean MgO content of the Murray below VRR (6.06 wt%). Hartmann’s

Valley is also characterized by higher Si and Al, and lower Ti, P and Fe concentrations than

mean Murray below VRR (Figures 5, 6, 7). Within the Hartmann’s Valley member, Si, Al,

Ti, Mn and Na concentrations generally decrease, and Fe, Ca and S concentrations increase

©2020 American Geophysical Union. All rights reserved.


with increasing elevation from the contact with the Pahrump Hills member (Figure 5). Si and

Al, and Ti and Cr positively correlate with one another (Table S7b).

3.1.3. Karasburg

The overlying Karasburg (~40 m thick) is the most compositionally homogenous of

the Murray members below VRR (Figure 6, Table S4b) and reveals positive correlations of

Si, Al and Na, of Si and Ti, of Na and Cl, and Ti, Cr and Fe (Table S7c).

3.1.4. Sutton Island

The lowermost Sutton Island member (~40 m thick) has a similar composition to the

underlying Karasburg (Figures 5, 6, 7; Table S4a, c). This is reflected in the similar

mineralogy determined by CheMin from the drilled samples taken from the Karasburg and

lower Sutton Island members (Quela, Marimba and Sebina) (Achilles et al., 2020; Bristow et

al., 2018). Towards the middle of the Sutton Island member, at -4330 m elevation, Si, Ti, Fe

and Cr concentrations increase, and Ca and S decrease (Figure 5). Sutton Island member

bedrock targets at -4330 m have higher Si, Al, Ti, Fe, K, Cr and FeO/MnO, and lower Mn,

Ca and S than the underlying Karasburg and lower Sutton Island (Figure 5; Table S5).

Two occurrences of elevated K2O, FeO and SO3 concentrations (up to 2.3 wt%, 24.8

wt%, and 13.3 wt% respectively) are detected within the lower Sutton Island member

associated with later diagenetic/alteration features (targets Thrumcap and Fresh Meadow)

(Thompson et al., 2017; Yen et al., 2018). Both reveal evidence for the addition of a K, Fe

sulfate phase or phases.

Through the rest of the Sutton Island member (above -4330 m elevation) to the

contact with the Blunts Point member (~50 m), there is a general trend to lower Si, Ti, Al, K

and Ca, and higher Mg and Zn contents, and to lower FeO/MnO (Figure 5). Sutton Island

(~90 m thick) is consistently depleted in Mn relative to mean Murray, except at the contact

©2020 American Geophysical Union. All rights reserved.


with the overlying Blunts Point member (Figure 7). The Sutton Island/Blunts Point member

contact is marked by low Si, K and Cr; and high Fe, Mn, Mg and Zn concentrations as well as

low FeO/MnO (Figures 5, 7). Mn shows up to 250% higher concentrations than mean Murray

(Figure 7). Five of the Sutton Island bedrock targets at this upper contact also exhibit elevated

P2O5 (1.06-1.27 wt%) compared to the mean for Sutton Island (0.98 wt%) (Figure 5). The

Sutton Island member (not including coarser sandstones) is the most compositionally varied

of all the Murray members below VRR (Figures 5, 6; Table S4b).

Less common, distinctive, thicker (up to ~30 cm thick), coarser grained and resistant

sandstone layers and lenses encountered above -4330 m elevation within the Sutton Island

member have different bulk compositions (consistently >Si, Al, Ti, Cr and K) than the more

typical Murray (e.g., Newport Ledge, Heron Island; Table S1). This is interpreted to reflect

differing provenance and/or sediment processing during transport. However, some of the

minor element trends observed for the coarser sandstones with elevation, are the same as

those within the mudstones (e.g., Mn, Mg and Zn, FeO/MnO; Figure S4).

Numerous diagenetic/alteration features encountered in the vicinity of the Sutton

Island/Blunts Point contact include resistant nodules and concretions, dark gray layers and

patches (Sun et al., 2018). These features reveal higher abundances of many of the same

elements (Fe, Mn, Mg, P and Zn) concentrated in the bedrock at this stratigraphic level

(Table S1; Figure S4) (Thompson et al., 2017). Dark, nodular/patchy features (Berry Cove

and Timber Point) within the Heron Island, coarse sandstone lens, exhibit higher

concentrations in the same elements (3.20-3.32 wt% P2O5, 0.8 wt% MnO, 20.2-23.5 wt%

FeO). This indicates that the coarser sandstones were subject to the action of the same

composition diagenetic fluids as the finer grained strata at this elevation. The abundance of

diagenetic features at the Sutton Island/Blunts Point contact indicates that it was the site of

enhanced post-depositional alteration. The varied chemistry associated with

©2020 American Geophysical Union. All rights reserved.


diagenetic/alteration features at different stratigraphic levels within the Murray formation

point to changing diagenetic/alteration fluid chemistry and the possibility of multiple

diagenetic/alteration episodes.

3.1.5. Blunts Point

Above the Sutton Island/Blunts Point member contact, the Blunts Point strata (~75-85

m thick) are statistically, compositionally the same as the underlying Karasburg and Sutton

Island members for Si, Al, Mg, and Na (Figures 5, 6; Table S4a, c). Additionally, Blunts

Point has the same mean Cr, Ni, Ca and S concentrations as Sutton Island (Figure 6; Table

S4a, c). Mean Ti, Fe, Mn, K, P and Zn concentrations are higher, and Cl concentrations lower

than for Sutton Island (Figure 6). Blunts Point has the highest mean FeO and Zn

concentrations of all the Murray members (21.73 wt%and 1642 ppm respectively). There is a

general trend to higher Si, Al and Ti, and lower Mn and FeO/MnO with increasing elevation

(Figure 5). Fe and K concentrations increase to approximately -4230 m elevation, decreasing

thereafter towards VRR (Figure 5). Zn is consistently high (1500-2400 ppm) to -4230 m, with

lower (mostly <1500 ppm), but variable concentrations (up to 2238 ppm) towards VRR

(Figure 5). Both Sutton Island and Blunts Point strata exhibit positive correlations of Si, Al

and Ti, and of Na and Cl (Table 7d, e). Sutton Island also reveals positive correlations of Mn,

Mg and Zn (Table 7d).

3.1.6. Calcium and Sulfur

Relatively late (post-lithification), white, cross-cutting veins with characteristically

high Ca and S contents are observed throughout the Murray formation and indeed the whole

stratigraphic section encountered by Curiosity (Minitti et al., 2017; Nachon et al., 2014,

2016). Layer-parallel or -subparallel CaSO4 veins are abundant within the Blunts Point

member, particularly in the vicinity of VRR, and are not typically observed below -4330 m

elevation. All vein targets plot on a trend consistent with the addition of a CaSO4 phase on a

©2020 American Geophysical Union. All rights reserved.


CaO versus SO3 plot (Figure S1). See §S2 in Supplementary Information for a detailed

discussion.

Plotting CaO versus SO3 for all Murray bedrock targets lacking evidence of veins or

diagenetic features reveals that the majority also fall on a CaSO4 addition trend (Figure S1).

A general trend to increasing Ca and S is observed, from the base of the Murray at Pahrump

Hills, through the Hartmann’s Valley, Karasburg and basal Sutton Island members (Figures

5, 6). Ca and S concentrations then decrease sharply to their lowest concentrations in the

Murray at -4330 m. Ca and S concentrations remain low compared to mean Murray through

the rest of the Sutton Island member and Blunts Point member.

The detection of Ca and S concentrations consistent with the addition of various

amounts of CaSO4 to a Precipice-like bedrock (the lowest Ca- and S-content typical Murray

bedrock, Table S1), and the lack of observable veins can be explained by the addition of a

CaSO4 cementing phase throughout much of the Murray formation. Changing Ca and S

concentrations with elevation are, therefore, probably reflecting varying abundances of

CaSO4 cement. The implied increase in CaSO4 cement within the Hartmann’s Valley through

to the lower Sutton Island member is supported by higher abundances of CaSO4 phases

detected by CheMin in the drilled samples from these members (Vaniman et al., 2018). We

acknowledge that some of the Ca and S may be present within fine veins or as fine grains, not

observed in MAHLI images. However, a number of diagenetic features (lacking visible

veins) within the Murray show elevated Ca and S relative to the substrate bedrock (Figure S1;

Table S1), indicating that Ca- and S-rich fluids played a role during diagenesis of the

mudstones, e.g., Camera Hill, Ash Island. This supports the interpretation that most of the

CaSO4 (not associated with visible veins) is present as cement.

A few targets have significant excess S that is not coupled with Ca (e.g. Jwaneng

DRT and offset; Tobane; Table S1), consistent with the presence of other, minor S-bearing

©2020 American Geophysical Union. All rights reserved.


phases. Based on SAM analyses of drilled samples, targets with excess S are interpreted to be

dominated by Mg-sulfates and Fe-sulfates, and in the case of Oudam, possibly Fe-sulfide

phases (Franz et al., 2017; McAdam et al., this issue; Sutter et al., 2017; Wong et al., this

issue). Such sulfates are inferred to be present in abundances either below, or at the detection

limits, of the CheMin XRD instrument and/or concentrated in the amorphous component of

drilled samples. Rapin et al. (2019) also infer the presence of Mg- and Ca-sulfate cements

from ChemCam analyses of Blunts Point strata.

4 Composition of VRR

Table 1 lists all bedrock compositional analyses obtained by APXS on VRR. The

VRR falls within the compositional range exhibited by the Murray formation encountered

below the ridge (Figures 6, 9). This supports the sedimentological interpretation that the ridge

strata are a continuation of the Murray formation and its associated depositional environment

(Edgar et al., this issue). Based on sedimentological and stratigraphic relationships (Edgar et

al., this issue) VRR strata have been divided into two additional members of the Murray

formation, namely the Pettegrove Point and Jura members (Figures 3, 5). Both members

exhibit color variations, which are more prevalent in the Jura member. A further distinction

has been made between patches of bedrock that appear bright gray/blue in color-stretched

HiRISE imagery, and the more typical red/tan strata of VRR (Fraeman et al., this issue a;

Figures 2, S6). These bright gray/blue patches are concentrated within the upper VRR Jura

member, but are also observed within the Pettegrove Point member, albeit less commonly

(Figure 2). They do not follow stratigraphy, but instead cut across primary bedding features.

Further details of the stratigraphy, sedimentology, textures and spectral properties of the ridge

strata are provided by Bennett, et al., (this issue), Edgar et al. (this issue), Fraeman et al. (this

issue a,b), Horgan et al., (this issue) and Jacob et al., (this issue).

©2020 American Geophysical Union. All rights reserved.


Although VRR falls within the range of Murray formation compositions encountered

below the ridge, it exhibits almost as much chemical variation as the whole of the underlying

Murray formation (Figures 5, 6, 8, 9; Table S6). A few VRR samples have lower Ti, Mn and

Fe, and higher Ni contents than all underlying Murray formation bedrock (Figures 6, 9). A

key observation for interpreting the ridge is that Fe concentrations on the ridge are typically

within the range of those observed for the underlying Murray formation, which, unlike VRR,

does not exhibit deep hematite absorptions in the orbital CRISM data (Fraeman et al., this

issue b).

4.1. Pettegrove Point member

No abrupt changes in bulk composition were detected as Curiosity drove from the

underlying Blunts Point member onto VRR (Figure 5). However, targets within ~4 m of the

Blunts Point/Pettegrove Point contact reveal low MnO (~0.1 wt%; Figures 5, 7), MgO

(mostly 4-6 wt%) and CaO (mostly 2-5 wt%) concentrations relative to the underlying Blunts

Point strata (Figure 5). MnO then trends to high concentrations (up to 0.42 wt%) towards the

contact with the overlying Jura member (Figures 5, 7). CaO and SO3 are also concentrated in

the vicinity of this contact (up to 7.18 wt% and 9.54 wt% respectively; Figure 5). Pettegrove

Point strata have higher Al, Na and Cl, and lower Ti, Cr, Fe, P, Zn and Br mean

concentrations than the underlying Blunts Point formation (Figure 6; Table S4a, c), with Si

and Al positively correlating with one another (Figure 8a; Table S7f).

Targets analyzed by APXS in areas exhibiting the deepest hematite absorptions in

CRISM and in situ data (ChemCam passive spectra and Mastcam multi-spectral; Fraeman et

al., this issue-a, b; Horgan et al., this issue; Jacob et al., this issue) along the traverse, within

the Pettegrove Point member, have typical Murray formation Fe concentrations (e.g.,

Stranraer DRT, Murchison and Voyageurs; Table 1; Figures 2, 5). Several targets with higher

than mean Murray Fe concentrations (e.g., Waboomberg, 24.37 wt% FeO) were analyzed on

©2020 American Geophysical Union. All rights reserved.


the ridge within the Pettegrove Point member, but they are not associated with the deepest

hematite spectral absorptions.

The Stoer drilled fines and brushed DRT bedrock surfaces have the same

compositional characteristics as the bulk Pettegrove Point member (Figure S7). The drilled

fines reveal an increase in Fe content relative to the corresponding brushed bedrock surface

(~22 wt% versus 18 wt% FeO respectively). Stoer drilled fines have similar compositions to

the underlying Marimba through Duluth drilled fines (Karasburg to Blunts Point; e.g., Figure

S7).

4.2. Typical Jura member

As with the transition from the Blunts Point to Pettegrove Point members, no abrupt

changes in composition were detected as Curiosity drove on to the Jura member (Figure 5).

Typica Jura is characterized by higher mean Si and Al, and lower Fe, Cr, Mn, P and Br

concentrations than the Pettegrove Point member, otherwise it is compositionally similar to

Pettegrove Point (Tables 1, S4a, c; Figure 6). The Si and Al positively correlate with one

another as is typical for the Murray formation in general (Figure 8a; Table S7g).

The Rock Hall drilled fines, and bedrock targets in the same workspace, have

compositions that extend the range of the typical Jura (Figures 5, 6). The unbrushed Rock

Hall bedrock target and two other targets in the same workspace (Corrieshalloch Gorge and

Bothwell DRT), have low Si and Al contents relative to the majority of typical Jura bedrock

targets, even after accounting for the addition of CaSO4. In particular, the Bothwell DRT

target has the lowest SiO2 concentration (40.69 wt%) of all typical Jura bedrock targets. The

Rock Hall drilled fines accentuate this trend, with lower Si and Al than all typical Jura

bedrock (Figure S7). The drilled fines also have more Ca, S and significantly more Cl and Br

than the bedrock (Table 1). The Rock Hall fines exhibit some of the lowest Si and Al and

©2020 American Geophysical Union. All rights reserved.


highest Ca and S concentrations of any Murray bedrock (Figure 6) and are compositionally

distinct from all other Murray drilled holes (Figure S7). The outcrop expression of the typical

Jura at this location was also distinct from other exposures visited by Curiosity within the

Jura. The bedrock was characteristically rough on a cm-scale, with textures suggestive of in

situ brecciation (Figure S5). Hence, the Rock Hall drilled target may not be representative of

typical Jura. This is also supported by the distinct spectral properties of the Rock Hall and

associated targets (Horgan et al., this issue).

4.3. Gray/blue bedrock patches

Targets analyzed by APXS within the gray/blue bedrock patches, the majority of

which occur within the Jura member, accentuate many of the trends observed within the

typical Jura, and are the most compositionally distinct of all VRR strata, compared to the rest

of the Murray formation (Table S4a, c). They have even higher SiO2 (up to 54.26 wt%) and

Al2O3 (up to 10.64 wt%), and lower FeO (down to 13.39 wt%) and TiO2 (down to 0.75 wt%)

contents than the rest of VRR (Figures 5, 6, 7, 8). Si and Al concentrations are the highest,

and Ti, P and Fe concentrations the lowest observed since the base of the Murray formation

(Pahrump Hills and Hartmann’s Valley members, Figure 5). The Si and Al positively

correlate with each other (Figure 8a; Table S7h), and when ratioed to Ti, Al and Fe reveal a

moderate positive correlation, unlike the rest of the Murray formation (Figure 8b). Si and Ti

have a Pearson correlation coefficient of only 0.26, in contrast to the moderate to strong,

positive correlation exhibited by the Karasburg through to Blunts Point members (Table S7c,

d, e, h) and Si/Ti ratios are generally higher than for the rest of the Murray (Figure 8c). Al/Ti

ratios are similarly higher than for the rest of the Murray, excepting a number of Pahrump

Hills targets (Figure 8d). Mn concentrations are lower (mostly <0.15 wt% MnO) than for

mean Murray formation (0.23 wt% MnO) (Tables 1, S4a, S6). The resistant bedrock target,

Aberfoyle, within a gray/blue Jura patch, is an exception with 0.21 wt% MnO. It also has

©2020 American Geophysical Union. All rights reserved.


very low Fe (10 wt% FeO), and elevated Ni (~1900 ppm) and Se (~70 ppm, Table S9). The

Inverness, Calgary and Findon targets, from a different gray/blue Jura patch, also have high

Mn (0.21 – 0.24 wt% MnO).

Highfield drilled fines and brushed DRT bedrock surfaces have the same

compositional characteristics as the bulk gray/blue Jura (Figure S7; Table 1). The drilled

fines are compositionally similar to the Hartmann’s Valley, Oudam drilled fines. Both exhibit

relatively high SiO2 (52.38 wt% - Highfield, 52.87 wt% - Oudam) and Al2O3 (10.02 wt% -

Highfield, 9.40 wt% - Oudam) contents, compared to other post-Pahrump Hills member

drilled samples (Table S1). Oudam does not have as low FeO concentration (18.74 wt%) as

Highfield (16.97 wt%), nor TiO2 (1.05 wt% - Oudam, 0.96 wt% - Highfield).

4.4. Variation in composition with elevation

Lateral variations in composition at the same elevation are observed. For example,

within the Pettegrove Point member, targets encountered at ~-4170 m elevation, but laterally

separated by ~250 m, exhibit a range in compositions e.g., SiO2 concentrations range from

~42-54 wt% and FeO from ~16-24 wt% (Figures 2, 5; Table 1). Within the Jura member,

targets analyzed at ~-4154 m elevation including typical Jura and gray/blue bedrock,

separated by up to 300 m, also reveal compositional heterogeneity, e.g., SiO2 concentrations

range from ~48-53 wt%, FeO from ~16-18.5 wt%, and MnO concentrations range from 0.09-

0.29 wt% (Figures 2, 5; Table 1).

The Si, Al and Fe trends are generally the reverse of those observed as Curiosity

gained in elevation through the Murray formation, from Pahrump Hills to Blunts Point

(Figure 8). Si and Al concentrations generally increase and Fe concentrations decrease from

Pettegrove Point, through typical Jura to the gray/blue bedrock patches (Figures 5, 6, 7, 8). Si

and Al exhibit a strong positive correlation with each other (r = 0.9) and a weak-moderate

negative correlation for Al and Fe (r = -0.5) within the ridge, as is the case for the rest of the

©2020 American Geophysical Union. All rights reserved.


Murray formation, excepting a number of Pahrump Hills targets (Table S7a-i). However,

when Al and Fe are ratioed to Ti, VRR reveals only a very weak negative correlation (r = -

0.19), even when excluding the blue/gray patches (Figure 8b; r = -0.27). Despite increasing

mean Si and Al concentrations for VRR, mean Ti concentrations decrease (Figures 5, 6, 7).

VRR reveals negative r-values for Si and Al with Ti, in contrast with the rest of the Murray

formation, again excepting a number of Pahrump Hills targets (Table S7a-e, i). While APXS

analyses indicate general compositional trends for Si, Al, Fe and Ti associated with each of

VRR members and the gray/blue patches, it is clear that there is significant overlap between

them.

VRR has consistently lower mean Ti, Zn, P, and Zn, and higher Na contents than the

underlying Blunts Point (Figures 5, 6). Potassium shows a broad trend to increasing

concentrations with elevation through the Pettegrove Point member into the base of the Jura

member (Figures 5, 6). K concentrations are variable within the Jura, with the second highest

variance (next to Hartmann’s Valley) of all the Murray members (Table S4b). Cl

concentrations, although variable, trend to high values (>1.5 wt% Cl) for a number of

Pettegrove Point and typical Jura member targets (not within gray/blue patches) (Figure 5).

The ridge exhibits a wide range of Mn concentrations (Figures 5, 6, 7) from relatively high

(~0.40 wt% MnO) compared to mean Murray (~0.23 wt% MnO), to among the lowest

detected at Gale (0.07 wt% MnO, Macleans Nose and Corrieshalloch Gorge). Low Mn

concentrations occur at the contact between the Blunts Point and Pettegrove Point members

and at the top of the ridge, within the Jura member.

Curiosity’s traverse facilitated sampling across the contact between the ridge and the

underlying Blunts Point member, as well as across the contact between the Pettegrove Point

and Jura members of VRR, at two laterally separated locations (Figure 3). The same overall

©2020 American Geophysical Union. All rights reserved.


chemical trends between the different members are observed across both transects, but at

different elevations (e.g., Figures 5, S8).

4.5. Diagenetic and alteration features

The following section documents the chemistry of diagenetic and alteration features,

including CaSO4 veining associated with VRR strata. This is to facilitate comparisons with

features observed throughout the Murray formation below VRR, and to assess their

relationship to compositional changes observed within the bulk bedrock on VRR. Where

observed, resistant nodules and concretions, are significantly finer scale (< 1 cm size) than

those within the underlying Murray formation (Bennett et al., this issue), and do not typically

exhibit different compositions than the surrounding host rock (Figure S4). A few exceptions

show the addition of CaSO4 and other possible minor sulfate phases (e.g., Barberton raster).

Distinctive, dark gray, rectangular raised features (<1 cm long) analyzed within one of

the gray/blue patches of bedrock detected from orbit (Haroldswick raster; Figure S9), do not

show clear compositional trends away from the substrate bedrock. The spot with highest

proportion of the raised features in the APXS FOV (Haroldswick 2) has elevated FeO (22.22

wt%) relative to typical gray/blue bedrock (~16 wt%).

Three resistant targets within another blue/grey Jura bedrock patch show variable Fe

contents (Figure S6). A bulbous raised target, Macleans Nose (~4 cm across), has slightly

elevated Fe (19.59 wt% FeO) relative to typical gray/blue bedrock (~16 wt% FeO). The

resistant bedrock target, Aberfoyle, with no obvious diagenetic features, has the lowest FeO

concentration (10.17 wt%) of any Murray bedrock analyzed by APXS (except for the high-Si

targets in the vicinity of Buckskin). The resistant bedrock target, Funzie has FeO

concentrations typical of other gray/blue bedrock targets (16.90 wt%). All three exhibit

elevated Ni (1300-1800 ppm) and all bedrock targets (not including vein targets) analyzed by

©2020 American Geophysical Union. All rights reserved.


APXS within this gray/blue patch have high Se concentrations (~70-100 ppm, detection limit

~40 ppm) (Table 1, S8).

In contrast to the underlying Blunts Point member, concordant or slightly discordant

veins parallel or subparallel to primary sedimentary layering are rarely observed on the ridge.

The Jura targets, Lyttleton and Balmedie, within a gray/blue patch, may be exceptions.

Strongly discordant white CaSO4 veins occur throughout VRR. As for the rest of the Murray

formation, a large number of targets do not exhibit obvious CaSO4 veins, yet have higher Ca

and S contents than the Precipice DRT target (lowest Ca- and S-content typical Murray

bedrock). Ca and S concentrations were low as Curiosity climbed onto the ridge, within the

basal Pettegrove Point member, but trend to high concentrations throughout most of the

overlying Pettegrove Point (Figures 5, 6). Pettegrove Point, Jura and gray/blue patches have

statistically the same mean Ca and S concentrations (Figure 6; Table S4a, c)

Within the gray/blue bedrock patches, CaSO4 veins are commonly observed to

contain very fine, dark inclusions (<2 mm) and rarer, coarser dark patches (up to 5 mm). The

only targets analyzed by APXS to specifically examine these phenomena were the Rona

raster and the Grange target (Figures S6, S9). However, several other APXS targets;

Newmacher offset, Forties, Portabello and Highfield offset had thin CaSO4 veins with very

fine dark inclusions within the nominal FOVs. All targets reveal elevated Fe associated with

increased proportions of the dark inclusions in the APXS FOVs. For example, the Rona 2

spot (Figure S9), with the greatest proportion of dark material included in the CaSO4 vein,

has the highest Fe content, of the three spots on the vein, as well as ~60 ppm Se (~40 ppm

detection limit; Table S8).

There are also regions of bedrock, adjacent to CaSO4 veins, possessing finer, dark inclusions

(<2 mm) surrounded by lighter gray areas (L’Haridon et al., this issue). APXS did not

specifically analyze any of the areas. However, a number of APXS targets within the

©2020 American Geophysical Union. All rights reserved.


gray/blue patches likely include both the fine, darker gray inclusions and lighter gray areas

within the APXS FOV, albeit not visible because of dust. This results in the APXS analyses

of the gray/blue patches propbably representing a bulk averaging of the fine, dark inclusions

and more areally extensive lighter gray bedrock. An example of an APXS target in which

these features are observable is Highfield DRT (Figure S9), which has a typical gray/blue

bedrock composition (Table 1).

5 Implications

How does the APXS data help to address the questions outlined in the Introduction?

The mineralogy and sedimentology of the ridge are consistent with VRR being a continuation

of the Murray formation depositional sequence (Edgar et al., this issue; Rampe et al., this

issue). Bulk compositions of VRR strata determined by APXS indicate that the ridge lies

within the compositional range of the Murray formation encountered prior to VRR,

supporting the interpretation, based on mineralogy and sedimentology, that there was no

significant change in the primary depositional environment associated with the ridge strata,

nor provenance.

Compositional trends in both major and trace elements associated with the Blunts

Point/Pettegrove Point and Pettegrove Point/Jura contact, cutting across elevation, imply that

the chemical characteristics and morphology of the ridge are not the result of primary

sedimentary or depositional processes, assuming that the strata are relatively flat lying (Stein

et al., this issue). In particular, the contacts between the high Si, Al and low Fe, Ti, Mn

gray/blue patches and more typical bedrock are observed to occur at relatively high angles to

the observed bedding (e.g., Figure 13 in Fraeman et al., this issue, a). Changes in provenance,

lake water chemistry or redox conditions at the sediment/water interface would be expected

to result in compositional changes that follow the flat-lying stratigraphy. Instead, chemical

variations along and across stratigraphy are more consistent with subsequent

©2020 American Geophysical Union. All rights reserved.


diagenetic/alteration episodes. The lateral and vertical compositional variability of VRR

compared to the underlying Murray formation, the presence of the gray/blue patches and

small-scale diagenetic features (mainly documented by ChemCam cf. L’Haridon et al., this

issue) indicate enhanced mobility of elements and, hence probable fluid flow within the

ridge. The fact that VRR is compositionally within the range of the Murray encountered

below the ridge also indicates that any post-depositional diagenesis/alteration did not result in

wholesale changes in composition.

No increase in Fe was observed as Curiosity climbed onto the ridge, nor within any of

the ridge strata. Specifically, the areas showing the deepest hematite spectral absorptions

from orbit and in situ measurements have average Murray Fe concentrations. Therefore, no

significant amount of Fe has been added to the bulk Murray bedrock comprising the ridge,

e.g., as a cementing oxide or as a detrital component. This is consistent with similar, or only

slightly higher, hematite abundances detected in VRR and Murray formation drilled samples

(the Pettegrove Point member, Stoer sample has 15.3 wt% hematite versus 13.9 wt% in

Oudam; Rampe et al., this issue). The lack of significant Fe addition to VRR strata

immediately discounts several of the proposed models for the formation of the ridge (see

Introduction, and Fraeman et al., 2013, 2016). Specifically, the redox interface and top down

oxidative weathering theories, that could explain enhanced hematite formation are not

feasible. Both would be expected to result in significantly increased Fe concentrations (as

well as Al and Ti for the oxidative weathering hypothesis; Nesbitt & Young, 1982) and Fe-

oxide mineral abundances within the ridge. Any model proposed must explain the trends

observed in the compositional data, as well as the mineralogy. Any model needs to account

for enhanced mobility of Mn versus Fe on the ridge, as well as the increased Si and Al, and

low Ti observed within the Jura member and the gray/blue bedrock patches specifically.

©2020 American Geophysical Union. All rights reserved.


Before assessing the merits of the different hypotheses for ridge formation, further

examination of the variation of Mn, Fe, Si and Al is required.

5.1. Elemental mass change calculation

We can investigate the variation of Si, Al, Fe and Mn within the Jura and gray/blue

bedrock patches by performing a mass change calculation relative to Pettegrove Point.

Sedimentological observations indicate no significant change in the depositional environment

between any of the ridge strata (Edgar et al., this issue). This, as well as the preservation of

the same igneous, crystalline mineralogy in all three VRR drilled samples (Rampe et al., this

issue), indicate the same sediment source input for all VRR strata. We attribute differences in

composition to post-depositional processes and propose that Pettegrove Point strata is the

least altered (supported by its similar composition to the underlying Blunts Point) on VRR.

Thus, we investigate possible gains and losses of elements within the Jura and gray/blue

bedrock relative to mean Pettegrove Point to provide insight into post-depositional processes

on VRR. We assume that Ti is behaving conservatively. Ti is typically assumed to behave

conservatively in terrestrial weathering solutions under almost all environmental conditions

(Mukhopadhyay et al., 2014, Nesbitt, 1979; Nesbitt and Wilson, 1992), only showing

significant mobility at pH<2 and pH>8 (Brookins 1988, Knauss et al., 2001). It has also been

shown to behave conservatively in geothermal (~100°C), acid sulfate solutions acting on

basalt (Markússon & Stefánsson, 2011) and within acid sulfate soils (Åström & Deng, 2003),

scenarios likely to resemble some aqueous conditions on Mars. The results of these

calculations are shown in Figure 10. See §S1.5 in Supplementary Information for details and

Table S2. The standard deviations of SiO2/TiO2, Al2O3/TiO2, FeO/TiO2, and MnO/TiO2 for

Pettegrove Point are also shown on the plot.

The results of the calculations, shown in Figure 10, as well as the comparison with

mean Murray formation (Figure 7), highlight the elevated Si and Al within the gray/blue

©2020 American Geophysical Union. All rights reserved.


patches, relative to both mean Murray, typical Jura and Pettegrove Point. Specifically, the

mass change calculations relative to Pettegrove point suggest the addition of Si and Al

(~20%) to the gray/blue patches versus passive enrichment. This is more than two times the

standard deviation of SiO2/TiO2 and Al2O3/TiO2 for Pettegrove Point. Typical Jura may also

be somewhat enriched in Si and Al relative to Pettegrove Point (~5-7%), with the exception

of the Rock Hall targets which show ~15% depletion in Si and Al. The low P concentrations

within the majority of VRR strata (Figure 6) relative to mean Murray, can also be explained

by the addition of Si and Al.

Iron is consistently low for the majority of grey/blue targets and typical Jura bedrock

relative to mean Murray (Figure 7). The mass change comparison with Pettegrove Point

indicates that there may have been some Fe loss from both the typical Jura and gray/blue

patches. However, the majority of gray/blue Jura targets (23 of 39) show <20% Fe loss,

similar to the standard deviation of FeO/TiO2 for Pettegrove Point, and six targets reveal Fe

gains. This indicates that the Fe was probably not significantly mobilized or leached from the

blue/grey patches. There is evidence for small scale Fe mobility within the gray/blue patches

(e.g., §4.5), but we rule out the bulk leaching of Fe. Any leaching of Fe was likely more

localized. Figure 7 reveals that the Pettegrove Point member has somewhat elevated Fe

relative to mean Murray formation, but it is not as marked as within the underlying Blunts

Point, which shows the most consistently elevated Fe concentrations of any Murray member.

The mass change relative to Pettegrove Point also highlights Mn variability within the

Jura, and the consistently depleted Mn within the gray/blue patches (excepting 6 targets,

§4.3). The comparison with mean Murray in Figure 7 reveals the same relationships. The

mass change indicates possible Mn gains and losses for the Jura, with Mn losses defining the

highest elevations on VRR within the Jura, and the gray/blue patches (Figure 10). The Mn

losses are consistently more marked than any loss of Fe, highlighting the increased mobility

©2020 American Geophysical Union. All rights reserved.


of Mn, and fractionation of Mn and Fe. Relative to mean Murray, the majority of Pettegrove

Point targets have high Mn concentrations, except those within 4 m of the contact with Blunts

Point (Figure 7). This implies that Pettegrove Point strata have also experienced Mn mobility.

It is also worth noting the contrast in the pattern of high and low concentrations

associated with the high-Si and -Al Pahrump Hills member relative to mean Murray, versus

those within the Jura member and grey/blue patches on VRR (Figure 7). The high-Si

Pahrump Hills targets also show elevated Ti, and very low Al, Fe and Mn concentrations (as

well as the majority of other elements). This suggests that passive enrichment of Si and Ti

and removal of other elements, versus addition of Si, is the more likely process. Additionally,

typical Pahrump Hills targets reveal more pronounced increases in Al versus Si

concentrations, and significant increases in Mn and decreases in Fe concentrations. This

contrasts with the equally elevated Al and Si, low Mn and only moderately decreasing Fe

concentrations within the Jura and grey/blue patches on VRR. These differences all point to

distinct processes being responsible for the elemental patterns observed within Pahrump Hills

versus VRR.

Hartmann’s Valley targets, however, do exhibit similar elemental increases and

decreases relative to mean Murray as the Jura and gray/blue patches, i.e., higher Si and Al,

and lower Ti, P and Fe concentrations than mean Murray (Figures 5, 6, 7). But they do not

show the same, consistently low Mn concentrations typical of the Jura and gray/blue patches.

The Oudam drilled fines (Hartmann’s Valley member) are also share some of the same

mineralogical characteristics as Highfield. Both have a siliceous amorphous component, opal-

CT and minor quartz, similar hematite abundances and low, phyllosilicate (ferripyrophyllite)

©2020 American Geophysical Union. All rights reserved.


contents (Achilles, 2020; Bristow et al., 2018; Rampe et al., this issue). These similar

compositional and mineralogical characteristics may indicate somewhat analogous processes.

5.2. Manganese and FeO/MnO

As discussed above, high and low Mn concentrations on VRR indicate mobility of Mn

within the ridge sediments (Figures 5, 7, 10). Increases and decreases in Mn concentrations

are more pronounced than for Fe, and variable FeO/MnO on the ridge (and the high ratio for

the Murray formation in general) indicates alteration of the primary igneous minerals and

other ferromagnesian minerals, and the fractionation of Fe and Mn. FeO/MnO is typically

~50 for unaltered igneous rocks on Mars and within martian meteorites (Mittlefehldt et al.,

2017; Papike et al., 2003; Yen et al., 2010) but ranges from ~50 to >300 on the ridge and is

high throughout the Murray formation (~100). Olivine is not detected by CheMin in any of

the Murray drilled samples above the Pahrump Hills member (Achilles, 2020; Bristow et al.,

2018; Rampe et al., this issue). Therefore, we conclude that olivine, phyllosilicates, pre-

existing Fe-oxides/hydroxides, the amorphous component and pyroxene are all potential

sources of mobile Fe and Mn (Tosca et al., 2004).

Divalent Fe and Mn cations have similar ionic radii, such that they have similar

geochemical behaviours. However, there are conditions under which fractionation can occur

by way of alteration in lower temperature fluids due to differences in solubility with

oxidation state and pH (e.g., Stumm & Morgan, 1996, Krauskopf, 1957). Mn2+ is oxidized

more slowly than Fe2+, such that it can be transported further into an oxidizing environment

before precipitating and becoming concentrated (Postma, 1985). Fluids containing Fe2+ can

reduce MnO present in sediments, resulting in Mn2+ enrichment in the fluid. If the Mn2+ is

then reoxidized, it can lead to concentrations of Mn (Postma, 1985; Postma & Appelo, 2000).

The geochemical cycling of Mn in lakes is influenced by acidification, and the solubility of

Mn has an inverse relationship with pH, such that low pH favors increased solubility of Mn,

©2020 American Geophysical Union. All rights reserved.


while high pH lakes act as sinks for Mn (Urban et al., 1990). Slowly neutralizing an acidic

solution acting on a basaltic andesite, in an oxidizing environment was shown to concentrate

Mn in the fluid, by precipitating out the less soluble, oxidized Fe compounds (Krauskopf,

1957). Mn concentrated in a fluid could then be precipitated out where evaporation

concentrates the Mn, for example in a closed basin brine, (e.g., Berger et al., 2019). In the

brine scenario, Mn2+ can also be oxidized by chlorate, or possibly hydrochlorous acid or

perchlorate and precipitated (Berger et al., 2019).

5.3. Mineralogical manifestation of the increased Si and Al on VRR

Increased Si and Al in the Jura, particularly for the gray/blue patches, is not consistent

with increased plagioclase content. The crystalline components of the Jura drilled samples

have similar bulk plagioclase contents to Blunts Point and Pettegrove Point drilled samples

(Rampe et al., this issue). 4 wt% opal-CT and minor quartz were detected in the Highfield

drilled fines, and most of the excess Si and, to a lesser extent Al, is associated with the

amorphous component. The Highfield amorphous material has 61.7 wt% SiO2 and 7.9 wt%

Al2O3 versus 34.3 wt% SiO2 and 0 wt% Al2O3 for Duluth, and 55.0 wt% SiO2 and 5.1 wt%

Al2O3 for the Stoer amorphous components. Additionally, Highfield contains a higher

abundance of amorphous material (49 wt%) than Duluth (37 wt%) and Stoer (38 wt%)

(Rampe et al., 2020b). The excess Si/Al ratio for Highfield relative to Stoer, both with similar

plagioclase contents, is approximately the same as the Si/Al ratio in the amorphous

component for Highfield. This suggests that any addition of Si and Al is manifest

predominantly as amorphous phases and not as clay or other mineral species.

5.4. Source of increased Al and Si in the Jura and gray/blue patches

Lower plagioclase abundances are detected by CheMin from the Karasburg and

Sutton Island drilled fines, than for the rest of the Murray formation (Achilles, 2020; Bristow

et al., 2018, Rampe et al., this issue). However, neither member shows consistently lower Si

©2020 American Geophysical Union. All rights reserved.


or Al concentrations (Figure 7). Hence it is unlikely that additional Si and Al is from the

dissolution of feldspar in the Karasburg and Sutton Island members. The Blunts Point

member shows minor, relative decreases in Si and Al concentrations (and Fe and Ti

increases) compared to mean Murray (Figure 7), but CheMin indicates a higher plagioclase

abundance for the drilled fines (Duluth) than for VRR samples. We therefore discount Si and

Al being sourced from the dissolution of plagioclase from any unit underlying VRR. Higher

Si and Al could be the result of dissolution of plagioclase, originally present in higher

concentrations within the Jura, and precipitation within an amorphous phase or phases,

concentrated in the gray blue patches.

The dissolution of Fe-Mg silicates (e.g., olivine, pyroxene and phyllosilcates) to form

Fe-oxides (fine grained hematite, magnetite, FeO.OH) in a closed system, would leave a Si-

rich (and Al-rich) residue (McLennan et al., 2003), which could then be dissolved and

mobilized.

Duluth has a slightly lower abundance of phyllosilicate and less amorphous material

than the Sutton Island drilled fines (Quela and Sebina, Achilles, 2020; Rampe et al., this

issue). Dissolution of the Blunts Point phyllosilicate and amorphous phases, and removal of

Al and Si from these phases might therefore be the source for at least some of the Si and Al.

Amorphous and poorly crystalline Al minerals are likely more soluble in acid sulfate fluids

than crystalline Al-minerals (Jones, Collins & Waite, 2011) and amorphous phases in general

are predicted to be more soluble than crystalline phases (e.g., Tosca & Knoll, 2009). Rock

Hall is also characterized by a relatively low amorphous content, with a low Si concentration

(Rampe et al., this issue).

5.5. Geochemical Models for VRR

The results of Curiosity’s exploration of Vera Rubin ridge have resulted in several

different models being proposed to explain the presence of the ridge and the observed

©2020 American Geophysical Union. All rights reserved.


mineralogy, composition and spectral properties (see summary in Fraeman et al., this issue,

b). How does the APXS-derived compositional data constrain/fit with the different models?

5.5.1. Action of reducing diagenetic fluids

David et al. (this issue), Frydenvang et al. (this issue) and L’Haridon et al. (this issue)

favor the action of reducing diagenetic fluids, resulting in small-scale mobilization of Fe, and

larger-scale mobilization of Mn. The action of reducing fluids could also explain the likely

presence of reduced sulfur species from SAM Evolved Gas Analyses (EGA) of VRR drilled

samples (McAdam et al., this issue; Wong et al., this issue). This model could be consistent

with the variation in Mn concentrations measured by APXS. It could explain the larger scale

mobilization of Mn versus Fe from the Jura. The presence of the dark, Fe-rich inclusions

within CaSO4 veins points to relatively late-stage diagenetic events being responsible for

these features (e.g., Rona and Grange, §4.5.). At least one separate alteration/fluid flow event

would also be required to explain the other compositional characteristics and the mineralogy

observed. These scenarios do not account for the higher Si and Al concentrations and lower

Ti evident in the APXS data from the Jura and specifically, the gray/blue bedrock patches.

5.5.2. High water to rock ratio alteration

Based on geochemical modelling, Turner et al. (this issue) suggest that much of the

observed mineralogy of VRR can be produced by the action of a dilute brine at temperatures

between 50 °C and 90-100 °C, with a relatively high water to rock ratio compared to at

Yellowknife Bay (~10,000). This model does not account for the presence of jarosite and

akageneite in Stoer and Rock Hall, which they attribute to later, localized acidic alteration.

Such a high water/rock would be expected to mobilize fluid-mobile cations. While variable

concentrations of Mn, Na, K, Mg and Ca are consistent with increased water/rock, relative to

the underlying Murray formation, other fluid-mobile cations such as Zn do not exhibit

significant variation in concentration within VRR strata, although Zn concentrations are

©2020 American Geophysical Union. All rights reserved.


consistently lower than in the underlying Blunts Point strata. Preservation of the igneous

minerals, plagioclase, potassium feldspar and pyroxene in all VRR drilled samples at greater

abundances than in the Karasburg and Sutton Island members, also indicate that, while

water/rock may be higher within VRR relative to most of the Murray formation, it was

probably not significantly enhanced, or may have been more localized. A relatively high

water/rock ratio would be consistent with the mobilization of Si and Al and their addition to

the gray/blue areas on VRR.

5.5.3. Action of warm, saline, acidic fluids

Based primarily on the mineralogy detected by CheMin in VRR drilled samples and

SAM EGA results, Rampe et al. (this issue), propose a model whereby the lacustrine

sediments are altered by the action of saline, acid sulfate fluids, with at least some of the

ridge strata experiencing moderately warm temperatures (~50-100 ºC). The presence of more

coarsely crystalline hematite on the ridge, and specifically gray hematite, and likely Fe-

pyrophyllite within the Highfield drilled sample, are attributed to the action of relatively

warm diagenetic fluids (e.g., Catling & Moore, 2003; Badaut et al., 1992). Opal-CT and

minor quartz present within the Highfield sample, and the likely presence of opal-CT in the

Stoer sample (Rampe et al., this issue), are also consistent with relatively warm fluids (e.g.,

Jones & Renaut, 2007).

The detection of relatively high abundances of well-crystalline akageneite and jarosite

within the Rock Hall drilled sample, are consistent with the sample having interacted with

acid-saline fluids (Rampe et al., this issue). Despite the texture and composition of Rock Hall

indicating that it may not be representative of typical Jura, the detection of minor akageneite

and jarosite in the Stoer drilled fines, indicate that saline, acid-sulfate fluids were likely

present within the ridge strata and not just confined to the Rock Hall area. However, they

may have been relatively localized.

©2020 American Geophysical Union. All rights reserved.


Addition of Si and Al to the Jura and gray/blue bedrock, and mobilization of these

elements, could be consistent with interaction with localized acid-saline fluids with strata

elsewhere within the Jura (e.g. Rock Hall), Pettegrove Point and maybe Blunts Point. The

highest dissolution rates of aluminosilicates, Al-oxide, -oxyhydroxides and -hydroxides occur

under acidic and alkaline conditions (Tobler et al., 2017). Dissolution of silica and silicate

phases is also increased under acidic conditions relative to circumneutral pH (Iler, 1979),

although Si is less easily mobilized by acidic than alkaline fluids, but elevated temperatures

will enhance solubility (Drever, 1997; Tobler et al., 2017). Precipitation of Al-phases

typically occurs through neutralization, at pH greater than ~3 to 4. Si is also less soluble at

pH>3 and in circumneutral fluids (Tobler et al., 2017).

Bowen et al., (2012) report significantly enhanced solubility of Si and Al in acidic

brine lake water and groundwater associated with acid saline lakes in Western Australia,

although the basement rocks are dominated by felsic gneisses, but intermediate and basic

gneisses are also present. Markússon & Stefánsson (2011) indicate that Si and Al are mobile

in geothermal (~100°C), acid sulfate solutions acting on basalt. The Si and Al are more

mobile than Fe below the oxidation front, with Fe mobility increasing above the oxidation

front, such that it is more mobile than Si. The same study also showed that Mn is more

mobile than Fe both below and above the oxidation front, and this is consistent with the

discussion in §5.2. Sánchez-España et al. (2016) report Si and Al in solution associated with

acid mine drainage pit lakes and flooded underground mines at pH<4. Wintsch & Kvale

(1994) note that silicate dissolution reactions take place relatively rapidly in the presence of

excess acid, with acid solutions being capable of transporting Si and Al. We are not implying

that Gale crater lake was an acid saline lake, nor that conditions akin to acid mine drainage

were prevalent within the lake or subsurface. However, these examples serve to illustrate that

the mobility of Si and Al could be enhanced by the presence of acid saline fluids.

©2020 American Geophysical Union. All rights reserved.


The localized action of subsurface acid saline fluids (pH = 2-3) could therefore

mobilize Si and Al (and Mn) from within the Jura (e.g., the Rock Hall area), Pettegrove Point

and possibly the underlying Blunt’s Point (all reveal possible depletion in Si and Al and

enrichment in Fe relative to mean Murray; Figure 7). The Si and Al dissolved in these fluids

would precipitate out upon encountering higher pH conditions (pH = 4-7) focused within the

gray/blue areas and the Jura at the top of VRR. The mineralogy of the Highfield samples

suggests that this precipitation occurred at somewhat elevated temperatures (~100°C). This

scenario requires enhanced fluid flow (relative to most of the rest of the Murray formation) to

transport and concentrate Al and Si, (see Turner et al., this issue, and above). Al and Si could

be further passively concentrated within the gray blue patches by local, small-scale Fe-losses.

The dark, Fe-rich inclusions in CaSO4 veins, and associated small-scale, dark Fe-rich features

surrounded by Fe-poor areas (§4.5; L’Haridon et al., this issue) within the gray/blue patches

are consistent with at least some of diagenetic/alteration processes and Fe-mobilization being

relatively late stage. However, the pervasive nature of the majority of compositional changes

and mineralogy associated with the ridge (i.e., not focused in discrete veins), indicates that

some of the alteration took place earlier. This suggests more than one diagenetic/alteration

event.

5.5.4. Proximity with the Greenheugh Pediment and cap rock

The proponents of the different geochemical models outlined above (Frydenvang et

al., this issue; Rampe et al., this issue; Turner et al., this issue) all suggest that an overlying,

relatively impermeable caprock, could have focused secondary, subsurface fluid flow along

the contact with underlying Murray formation strata (either associated with groundwater flow

from the Southern Highlands to the Northern Lowlands, or from Mount Sharp). The cap rock,

now eroded away, is considered to have been relatively impermeable owing to the fact that its

remnants form more resistant mesas, ledges and caps of hills, and are crater-retaining, all

©2020 American Geophysical Union. All rights reserved.


features consistent with it being more indurated and better cemented than the recessive

Murray formation, which does not retain craters.

The erosional expression of the Greenheugh pediment, and the associated cap rock to

the south of VRR, mimics the geomorphology of the ridge (Fraeman et al., 2016; Bryk et al.,

2019). The dips of the Greenheugh pediment capping unit, and extrapolation towards VRR,

indicate that the cap rock may have extended unconformably over the ridge (Bryk et al.,

2019; Stein et al., this issue). This could explain the relatively more varied chemistry within

the Jura member. The distinct composition of the gray/blue bedrock patches indicate that they

may have acted as a locus for enhanced flow of relatively warm fluids and secondary,

diagenetic/alteration within VRR. Rampe et al. (this issue) and Fraeman et al. (this issue)

discuss potential heat sources and the origin of the acidity and salinity for the fluids.

The compositionally and mineralogically similar Oudam drill hole is situated within

the Hartmann’s Valley member, only 1 m below the contact with the unconformably

overlying, resistant Stimson formation sandstone. Both the Greenheugh pediment capping

unit and Stimson formation are interpreted to belong to the unconformable Siccar Point group

(Bryk et al., 2019). Therefore, the Murray formation at both locations may have been subject

to the same, post- depositional, diagenetic/alteration processes, as a result of their proximity

to the Siccar Point group cap rocks. Thus, the distribution of many diagenetic/alteration

effects might primarily be controlled by proximity to the unconformity rather than the

stratigraphy of the Murray formation.

The high-Si, Murray formation, Buckskin drill hole and associated targets are within

0.2 m of a contact with Stimson formation sandstones, suggesting that the observed

mineralogy and chemistry could also be the result of enhanced fluid flow along the Murray

formation/Siccar Point group contact. The Garden City vein system within the Pahrump Hills

member is also close to the contact with the overlying Siccar Point group. If Curiosity is able

©2020 American Geophysical Union. All rights reserved.


to access the contact between the Murray formation and the overlying Greenheugh pediment

capping unit during the continued ascent of Mount Sharp we may encounter similar rocks,

indicating a relatively widespread diagenetic/alteration event at this contact, albeit focused in

specific areas along the contact.

6 Conclusions

The compositional data obtained by the APXS on VRR, as well as within the

underlying Murray formation, combined with sedimentological and mineralogical

observations, support a similar provenance and depositional history for the whole of the

Murray formation. The discordant nature of elemental trends associated with the ridge-

forming strata, particularly the gray/blue patches, indicate that secondary

diagenetic/alteration processes cutting across the stratigraphy were responsible for the

formation of VRR. Changes in provenance, sedimentary processes or systematic changes in

lake water chemistry during deposition are considered unlikely. The hematite spectral

signature detected from orbit on VRR is not associated with an increase in the bulk iron

content within the APXS data. This rules out the addition of significant amounts of Fe-rich

cement or detrital minerals.

The compositional variability of the ridge over a relatively narrow elevation range

(~50 m), and laterally (over 300-400 m), indicates some enhanced mobility of elements and

fluid flow relative to the underlying Murray formation. In particular, Si and Al were likely

added to the blue/gray patches, and Mn was mobilized and fractionated from Fe. We favor

the action of acid saline fluids at moderate water/rock ratios facilitating mobilization of Si, Al

and Mn. The somewhat enhanced fluid flow allowed transport and addition of Si and Al as

amorphous phases, at high pH and relatively warm temperatures within the gray/blue patches.

A relatively impermeable, overlying capping unit may have contained and focused the fluid

©2020 American Geophysical Union. All rights reserved.


flow, which was likely concentrated within the Jura member, and more specifically within the

gray/blue outcrops identified from orbit.

Different elemental enrichments and depletions within the ridge strata, versus the

underlying Murray formation, support multiple, post-depositional, diagenetic/alteration

processes with varying fluid chemistries, redox conditions, pH, temperatures and timing

occurring throughout the Murray formation.

Acknowledgments, Samples, and Data:

MSL APXS is managed and financed by the Canadian Space Agency (CSA). We
acknowledge the important role of the engineers at JPL, the MSL science team and everyone
involved in operations with respect to obtaining APXS data.
Science team member funding for Thompson, Boyd, Gellert, O’Connell-Cooper, Schmidt and
Spray is provided by the CSA. Fraeman and Yen acknowledge that a portion of this research
was carried out at the Jet Propulsion Laboratory, California Institute of Technology, under a
contract with the National Aeronautics and Space Administration. Berger was funded by a
NASA Postdoctoral Program fellowship administered by USRA. VanBommel was supported
by NASA/Caltech/JPL Contract 1549716 to Washington University in St. Louis for
participation in the Mars Science Laboratory Science Team. Thompson thanks the VRR
working group, and specifically; Kristen Bennett, Briony Horgan and Samantha Jacob for
fruitful discussions. We acknowledge the thorough and constructive feedback from three
anonymous reviewers.
Data Availability
All MSL, raw data are available at the planetary data system:
https://pds-geosciences.wustl.edu/missions/msl/index.htm
https://an.rsl.wustl.edu/msl/mslbrowser/an3.aspx
Bradbury group and Siccar Point group APXS analyses are available at the planetary data
system:
https://pds-geosciences.wustl.edu/msl/msl-m-apxs-45-rdr-v1/mslapx1xxx/data/
All data presented in this paper, including the results of statistical tests and data used to make
plots are available at: Thompson, L. (2020). Alpha Particle X-ray spectrometer geochemistry
of the Murray formation and Vera Rubin ridge, Gale crater, Mars.
https://doi.org/10.25545/ZXDJZ7, UNB, V1,
UNF:6:bL/a2qTZBu6DNJlzkmGAbg==[fileUNF]

©2020 American Geophysical Union. All rights reserved.


Cross References to articles in this issue:

Bennett et al., Extensive diagenesis revealed by fine-scale features at Vera Rubin ridge,
Gale crater, Mars, 2019JE006311
David et al., Iron oxide mineral grains observed by ChemCam across Vera Rubin ridge
sedimentary rocks at Gale crater, Mars, 2019JE006314 (accepted)
Edgar et al., A lacustrine paleoenvironment recorded at Vera Rubin ridge, Gale crater:
Overview of the sedimentology and stratigraphy observed by the Mars Science
Laboratory Curiosity rover, 2019JE006307RR (accepted)
Fraeman et al., Evidence for a diagenetic origin of Vera Rubin ridge, Gale crater, Mars:
Summary and synthesis of Curiosity’s exploration campaign, 2020JE006527 (accepted)
Fraeman et al., Synergistic ground and orbital observations of iron oxides on Mt. Sharp and
Vera Rubin ridge, 2019JE006294 (accepted)
Frydenvang et al., The chemostratigraphy of the Murray formation and role of diagenesis
at Vera Rubin ridge in Gale crater, Mars, as observed by the ChemCam instrument,
2019JE006320R (accepted)
Horgan et al., Diagenesis of Vera Rubin ridge, Gale crater, Mars from Mastcam
multispectral images, 2019JE006322
Jacob et al., Spectral, Compositional, and Physical Properties of the Upper Murray
Formation and Vera Rubin Ridge, Gale Crater, Mars, 2019JE006290 (accepted)
L’Haridon et al., Iron Mobility during Diagenesis as Observed by ChemCam at the Vera
Rubin Ridge, Gale Crater, Mars, 2019JE006299 (accepted)
McAdam et al., Constraints on the Mineralogy and Geochemistry of the Vera Rubin ridge,
Gale crater, Mars, from Mars Science Laboratory Sample Analysis at Mars Evolved Gas
Analyses, 2019JE006309R (accepted)
Rampe et al., Mineralogy of Vera Rubin Ridge from the Mars Science Laboratory CheMin
Instrument, 2019JE006306R (accepted)
Stein et al., Regional structural orientation of the Mt. Sharp group revealed by in-situ dip
measurements and stratigraphic correlations on the Vera Rubin ridge, 2019JE006298R
(accepted)
Turner et al., Fluid-rock Reactions in the Murray Formation, including Vera Rubin ridge,
Gale crater, Mars, 2020JE006447
Wong et al., Detection of reduced sulfur on Vera Rubin ridge by quadratic discriminant
analysis of volatiles observed during evolved gas analysis, 2019JE006304 (accepted)

References:
Achilles, C. N., Rampe, E. B., Downs, R. T., Bristow, T. F., Ming, D. W., Morris, R. V., et
al. (2020). Evidence for multiple diagenetic episodes in ancient fluvial-lacustrine
sedimentary rocks in Gale crater, Mars. Journal of Geophysical Research, Planets
125(8), doi:10.1029/2019JE006295
Anderson, R. B., & Bell, J. F. III (2010). Geologic mapping and characterization of Gale
crater and implications for its potential as a Mars Science Laboratory landing site. Mars
5, 76–128. doi.org/10.1555/mars.2010.0004 Åström, M., & Deng, H. (2003).
Assessment of the mobility of trace elements in acidic soils using soil and stream

©2020 American Geophysical Union. All rights reserved.


geochemical data. Geochemistry: Exploration, Environment, Analysis, 3(2), 197–203.
doi.org/10.1144/1467-787302-034
Badaut, D., Decarreau, A., & Besson, G. (1992). Ferripyrophyllite and related Fe3+-rich
2:1 clays in recent deposits of Atlantis II Deep, Red Sea. Clay Minerals, 27, 227-244.
Banham, S.G., Gupta, S., Rubin, D.M., Watkins, J.A., Sumner, D.Y., Edgett, K.S., et al.
(2018). Ancient martian aeolian processes and palaeomorphology reconstructed from
the Stimson formation on the lower slope of Aeolis Mons, Gale crater, Mars.
Sedimentology, 65, 993–1042. doi.org/10.1111/sed.12469
Bedford, C. C., Bridges J. C., Schwenzer, S. P., Wiens, R. C., Rampe, E. B., Frydenvang,
J. & Gasda, P. J. (2019). Alteration trends and geochemical source region characteristics
preserved in the fluviolacustrine sedimentary record of Gale crater, Mars. Geochimica et
Cosmochimica Acta, 246, 234-266. doi.org/10.1016/j.gca.2018.11.031
Bennett, K. A., Rivera-Hernandez, F., Tinker, C., Horgan, B., Fey, D. M., Edwards, C., et
al. (2020). Extensive diagenesis revealed by fine-scale features at Vera Rubin ridge,
Gale crater, Mars. Submitted to Journal of Geophysical Research
Berger, J. A., Schmidt, M. E., Gellert, R., Boyd,N. I., Desouza, E. D., Flemming, R. L., et
al. (2017). Zinc and germanium in the sedimentary rocks of Gale Crater on Mars
indicate hydrothermal enrichment followed by diagenetic fractionation. Journal of
Geophysical Research, Planets, 122, 1747–1772, doi:10.1002/ 2017JE005290
Berger, J.A., King, P.L., Gellert, R., Clark, B.C., O’Connell-Cooper, C.D., Thompson,
L.M., VanBommel, S.J.V. & Yen A.S. (2019). Manganese enrichment pathways
relevant to Gale crater, Mars: Evaporative concentration and chlorine-induced
precipitation. Paper presented at 50th Lunar and Planetary Science Conference, Abstract
2847. Lunar and Planetary Institute, Houston.
Berger, J.A., Schmidt, M. E., Campbell, J. L., Flannigan, E. L., Gellert, R., Ming, D. W., &
Morris, R. V. (2020) Particle Induced X-ray Emission spectrometry (PIXE) of Hawaiian
volcanics: An analogue study to evaluate the APXS field analysis of geologic materials
on Mars. Icarus, 345, 113708. doi.org/10.1016/j.icarus.2020.113708
Bibring, J-P., Langevin, T., Mustard, J. F., Poulet, F., Arvidson, R., Gendrin, A., et al.
(2006). Global mineralogical and aqueous Mars history derived from OMEGA/Mars
Express data. Science 312(5772), 400-404. doi:10.1126/science.1122659
Blake, D.F., Vaniman, D., Achilles, C., Anderson, R., Bish, D., Bristow, T., et al. (2012).
Characterization and calibration of the CheMin mineralogical instrument on Mars
Science Laboratory. Space Science Reviews 170: 341-399. doi:10.1007/s11214-012-
9905-1
Bookins, D. G. (1988). Eh-pH diagrams for geochemistry, Berlin: Springer-Verlag, p.176.
Bowen, B.B., Benison, K.C., & Story, S. (2012). Early diagenesis by modern acid brines in
Western Australia and implications for the history of sedimentary modification on
Mars, In J. Grotzinger & R. Milliken (Eds.), Sedimentary Geology of Mars: SEPM
Special Publication (102, pp. 229–252).
Brimhall, G. H., & Dietrich, W. E. (1987). Constitutive mass balance relations between
chemical composition, volume, density, porosity, and strain in metasomatic
hydrochemical systems: results on weathering and pedogenesis. Geochimica et
Cosmochimica Acta, 51(3), 567–587.

©2020 American Geophysical Union. All rights reserved.


Bristow, T.F., Rampe, E.B., Achilles, C.N., Blake, D.F., Chipera, S.J., Craig, P. et al.
(2018). Clay mineral diversity and abundance in sedimentary rocks of Gale crater, Mars.
Science Advances, 4(6), article ear3330, doi:10.1126/sciadv.aar3330
Bryk, A. B., Dietrich, W. E., Lamb, M. P., Grotzinger, J. P., Vasavada, A. R., Stack, K. M.,
et al., (2019). In Curiosity’s path; The geomorphology and stratigraphy of the
Greenheugh pediment and Gediz Vallis ridge in Gale crater. Paper presented at 50th
Lunar and Planetary Science Conference, Abstract 2263. Lunar and Planetary Institute,
Houston.
Campbell, J. L., Perrett, G. M., Gellert, R., Andrushenko, S. M., Boyd, N. I., Maxwell, J.
A., et al. (2012). Calibration of the Mars Science Laboratory alpha particle X‐ray
spectrometer. Space Science Reviews, 170, 319– 340. doi.org/10.1007/s11214‐012‐
9873‐5
Campbell, J. L., King, P. L., Burkemper, L., Berger, J. A., Gellert, R., Boyd, N. I., et al.
(2014). The Mars Science Laboratory APXS calibration target: Comparison of Martian
measurements with the terrestrial calibration. Nuclear Instruments and Methods in
Physics Research Section B: Beam Interactions with Materials and Atoms, 323, 49– 58.
doi.org/10.1016/j.nimb.2014.01.011
Catling, D. C., & Moore, J. M. (2003). The nature of coarse-grained crystalline hematite
and its implications for the early environment of Mars. Icarus, 165(2), 277-300.
doi.org/10.1016/S0019-1035(03)00173-8.
Cousin, A., Sautter, V., Payre, V., Forni, O., Mangold, N., Le Diet, L., et al. (2017).
Classification of igneous rocks analyzed by ChemCam at Gale Crater, Mars. Icarus 288,
265-283. doi:10.1016/j.icarus.2017.01.014
David, G., Cousin, A., Forni, O., Meslin, P.-Y., Dehouck, E., Mangold, N., et al. (2020).
Analyses of high‐iron sedimentary bedrock and diagenetic features observed with
ChemCam at Vera Rubin ridge, Gale crater, Mars: calibration and characterization.
Journal of Geophysical Research doi: 10.1029/2019JE006314
Drever, J. I., (1997). The geochemistry of natural waters. Upper Saddle River, New Jersey,
Prentice Hall. P. 1997
Edgar, L.A., Fedo, C. M., Gupta, S., Banham, S.G., Fraeman, A.A. Grotzinger, J.P et al.
(2020). A lacustrine paleoenvironment recorded at Vera Rubin ridge, Gale crater:
Overview of the sedimentology and stratigraphy observed by the Mars Science
Laboratory Curiosity rover. Journal of Geophysical Research, Planets, 125(3) doi:
10.1029/2019JE006307
Edgar, L.A., Gupta, S., Rubin, D.M., Lewis, K.W., Kocurek, G.A., Anderson, R.B., et al.
(2018). Shaler: in situ analysis of a fluvial sedimentary deposit on Mars. Sedimentology,
65(1):96-122, doi:10.1111/sed.12370
Edgett, K. S., Yingst, R. A., Ravine, M. A., Capilinger, M. A., Maki, J. N., Ghaemi, F. T.,
et al. (2012). Curiosity’s Mars Hand Lens Imager (MAHLI) Investigation. Space
Science Review, 170: 259. https://doi.org/10.1007/s11214-012-9910-4
Edwards, P. H., Bridges, J. C., Wiens, R., Anderson, R., Dyar, D., Fisk, M., et al. (2017).
Basalt–trachybasalt samples in Gale crater, Mars. Meteoritics and Planetary Science, 52
2931-2410. doi.org/10.1111/maps.12953
Fedo, C.M., Grotzinger, J.P., Gupta, S., Fraeman, A., Edgar, L., Edgett, K., et al. (2018a).
Sedimentology and stratigraphy of the Murray formation, Gale crater, Mars. Paper

©2020 American Geophysical Union. All rights reserved.


presented at 49th Lunar and Planetary Science Conference, Abstract 2078. Lunar and
Planetary Institute, Houston.
Fedo, C., Grotzinger, J. P., Gupta, S., Banham, S., Bennett, K., Edgar, L. A., et al. (2019).
Evidence for persistent, water-rich lacustrine deposition preserved in the Murray
formation, Gale crater: A depositional system suitable for sustained habitability. Paper
presented at Ninth International Conference on Mars. Abstract 6308. Pasadena,
California
Fraeman, A., Arvidson, R., Catalano, J., Grotzinger, J., Morris, R., Murchie, S., et al.
(2013). A hematite‐bearing layer in Gale Crater, Mars: Mapping and implications for
past aqueous conditions, Geology, 41, 1103–1106. doi:10.1130/G34613.1
Fraeman, A. A., Ehlmann, B. L., Arvidson, R. E., Edwards, C. S., Grotzinger, J. P.,
Milliken, R. E., Quinn, D. P., & Rice, M. S. (2016). The stratigraphy and evolution of
lower Mount Sharp from spectral, morphological, and thermophysical orbital data sets.
Journal of Geophysical Research, Planets, 121, 1713–1736. doi:10.1002/2016JE005095
Fraeman, A.A., Edgar, L.A., Rampe, E. B., Thompson, L. M., Frydenvang, J., Fedo, C.
M., et al. (2020a). Evidence for a diagenetic origin of Vera Rubin ridge, Gale crater,
Mars: Summary and synthesis of Curiosity’s exploration campaign. Journal of
Geophysical Research Planets doi: 10.1029/2020JE006527
Fraeman, A. A., Johnson, J. R., Arvidson, R. E., Rice, M. S., Wellington, D. F., Morris
R. V., et al. (2020b). Synergistic ground and orbital observations of iron oxides on
Mt. Sharp and Vera Rubin ridge. Journal of Geophysical Research Planets doi:
10.1029/2019JE006294
Franz, H. B., McAdam, A.C., Ming, D.W., Freissinet, C., Mahaffy, P.R., Eldridge, D.L., et
al. (2017). Large sulfur isotope fractionations in Martian sediments at Gale crater.
Nature Geoscience, 10, p.658. https://doi.org/10.1038/ngeo3002.
Frydenvang, J., Gasda, P.J., Hurowitz, J.A., Grotzinger, J.P., Wiens, R.C., Newsom, et al.
(2017). Diagenetic silica enrichment and late-stage groundwater activity in Gale crater,
Mars. Geophysical Research Letters, 44, 4716–4724.
https://doi.org/10.1002/2017GL073323
Frydenvang, J., Mangold, N., Wiens, R. C., Fraeman, A. A., Edgar, L. A., Fedo, C. M., et
al. (2020). The chemostratigraphy of the Murray formation and role of diagenesis at
Vera Rubin ridge in Gale crater, Mars, as observed by the ChemCam instrument.
Journal of Geophysical. Research Planets doi:10.1029/2019JE006320
Gellert, R., Rieder, R., Brückner, J., Clark, B.C., Dreibus, G., Klingelhöfer, G., et al.
(2006). Alpha Particle X-Ray Spectrometer (APXS): Results from Gusev crater and
calibration report. Journal of Geophysical. Research, Planets, 111, E02S05.
doi.org/10.1029/2005JE002555
Gellert, R., Clark, B.C. III, & Mars Science Laboratory (MSL) Science Team (2015). In
situ compositional measurements of rocks and soils with the Alpha Particle X-ray
Spectrometer on NASA’s Mars rovers. Elements 11, 39–44.
https://doi.org/10.2113/gselements.11.1.39
Grotzinger, J. P., Crisp, J., Vasavada, A. R., Anderson, R. C., Baker, C. J., Barry, R., et al.
(2012). Mars Science Laboratory mission and science investigation. Space Science
Reviews, 170, 5 – 56. doi:10.1007/s11214‐012‐9892‐2.

©2020 American Geophysical Union. All rights reserved.


Grotzinger, J.P., Sumner, D. Y., Kah, L. C., Stack, K., Gupta, S., Edgar, L., et al. (2014). A
habitable fluvio‐lacustrine environment at Yellowknife Bay, Gale Crater, Mars. Science,
343(6169), 1242777. doi:10.1126/science.1242777.
Grotzinger, J. P., Gupta, S., Malin, M.C., Rubin, D.M., Schieber, J., Siebach, K., et al.
(2015). Deposition, exhumation, and paleoclimate of an ancient lake deposit, Gale
Crater, Mars. Science, 350(6257), aac7575. doi:10.1126/science.aac7575.
Horgan, B., H. N., Johnson, J.R., Fraeman, A. A., Rice, M. S., Seeger, C., Bell, J. F. III, et
al., (2020) Diagenesis of Vera Rubin ridge, Gale crater, Mars from Mastcam
multispectral images. Submitted to Journal of Geophysical Research
Hurowitz, J. A., Grotzinger, J. P., Fischer, W. W., McLennan, S. M., Milliken, R. E., Stein,
N., et al. (2017). Redox stratification of an ancient lake in Gale crater, Mars. Science,
356, eaah6849
Iler, R. K. (1979). The chemistry of silica. New York, NY: John Wiley and Sons, Inc., 866
p.
Jacob, S. R., Wellington, D. F., Bell III, J. F., Achilles, C., Fraeman, A. A., Horgan, B. , et
al. (2020) Spectral, compositional, and physical properties of the upper Murray
formation and Vera Rubin ridge, Gale Crater, Mars. Journal of Geophysical Research
doi.org/10.1029/2019JE006290
Jones, B., & Renaut, R. W. (2007). Microstructural changes accompanying the opal-A to
opal-CT transition: new evidence from the siliceous sinters of Geysir, Haukadalur,
Iceland. Sedimentology, 54(4), doi:10.1111/j.1365-3091.2007.00866.x
Jones, A. M., Collins, R. N., & Waite, T. D. (2011). Mineral species control of aluminum
solubility in sulfate-rich acidic waters. Geochmica et Cosmochimica Acta, 75, 965-977.
Knauss, K. G., Dibley, M. J., Bourcier, W. L., & Shaw, H. F. (2001). Ti (IV) hydrolysis
constants derived from rutile solubility measurements made from 100 - 300°C. Applied
Geochemistry, 16, 1115 – 1128.
Krauskopf, K. B. (1957). Separation of manganese from iron in sedimentary processes.
Geochimica et Cosmochimica Acta, 12, 61-84. https://doi.org/10.1016/0016-
7037(57)90018-2
Kronyak, R.E., Kah, L.C., Edgett, K.S., VanBommel, S.J., Thompson, L.M., Wiens, R.C.,
Sun, V.Z., & Nachon, M. (2018). Mineral-filled fractures as indicators of
multigenerational fluid flow in the Pahrump Hills member of the Murray formation,
Gale crater, Mars. Earth and Space Science, 6, 238–265.
https://doi.org/10.1029/2018EA000482
Le Maitre, R. W. (Editor), Streckeisen, A., Zanettin, B., Le Bas, M. J., Bonin, B., Bateman,
P., et al. (2002). Igneous Rocks: A Classification and Glossary of Terms,
Recommendations of the International Union of Geological Sciences, Subcommission of
the Systematics of Igneous Rocks. Cambridge University Press, ISBN 0-521-66215-X
L’Haridon, J., Mangold, N., Fraeman, A. A., Johnson, J. R., Cousin, A., Rapin, W., et al.
(2020). Iron mobility during diagenesis at the Vera Rubin ridge, Gale Crater, Mars.
Journal of Geophysical Research doi:10.1029/2019JE006299
Mahaffy, P. R., Webster, C. R., Cabane, M., Conrad, P. G., Coll, P., Atreya, S. K., et al.
(2012). The sample analysis at Mars investigation and instrument suite. Space Science
Reviews, 170(1-4), 401-478.

©2020 American Geophysical Union. All rights reserved.


Malin, M.C., Ravine, M.A. Caplinger, M.A. Ghaemi, F.T. Schaffner, J.A. Maki, J.N., et
al. (2017). The Mars Science Laboratory (MSL) Mast cameras and Descent imager:
Investigation and instrument descriptions. Earth and Space Science, 4(8):506-539,
doi:10.1002/2016EA000252
Markússon, S. H., & Stefánsson, A. (2011). Geothermal surface alteration of basalts,
Krýsuvík Iceland - Alteration mineralogy, water chemistry and the effects of acid
supply on the alteration process. Journal of Volcanology and Geothermal Research,
206(1-2), 46-59. doi.org/10.1016/j.jvolgeores.2011.05.007
Maurice, S., Wiens, R.C., Saccoccio, M., Barraclough, B., Gasnault, O., Forni, O., et al.
(2012). The ChemCam instrument suite on the Mars Science Laboratory (MSL) rover:
Science objectives and mast unit description. Space Science Reviews, 170:95-166,
doi:10.1007/s11214-012-9912-2
McAdam, A. C., Sutter, B., Archer, P. D., Franz, H. B., Wong, G. M., Lewis, J. M. T., et
al. (2020). Constraints on the Mineralogy and Geochemistry of the Vera Rubin ridge,
Gale crater, Mars, from Mars Science Laboratory Sample Analysis at Mars Evolved Gas
Analyses. Journal of Geophysical Research Planets, doi:10.1029/2019JE006309
McLennan, S. M., (2003). Sedimentary silica on Mars. Geology, 31, p.315-318.
Doi:10.1130/0091-7613(2003)031<0315:SSOM>2.0.CO;2
McLennan, S. M., Anderson, R. B., Bell III, J. F., Bridges, J. C., Calef III, F., Campbell, J.
L., et al.,(2013). Elemental Geochemistry of Sedimentary Rocks at Yellowknife Bay,
Gale Crater, Mars. Science, 343, Iss. 6169. doi:10.1126/science.1244734
McSween, H.Y. Jr. (2015). Petrology on Mars. American Mineralogist, 100, 2380–2395.
https://doi.org/10.2138/am-2015-5257
McSween, H.Y. Jr., Taylor, G.J. & Wyatt, M.B. (2009). Elemental composition of the
martian crust. Science 324, 736-739. https://doi.org/10.1126/science.1165871
Milliken, R. E., Grotzinger, J. P., & Thomson, B. J. (2010). Paleoclimate of Mars as
captured by the stratigraphic record in Gale Crater. Geophysical Research Letters,
37(4), L04201. doi:10.1029/2009GL041870
Minitti, M. E., Kennedy, M. R., Krezoski, G. M., Rowland, S. K., Schieber, J., Stack, K.
M., & Yingst, R. A. (2017). Using MARDI twilight images to assess variations in the
Murray formation with elevation, Gale crater, Mars. Presented at 48th Lunar and
Planetary Science Conference, Abstract 2622. Lunar and Planetary Institute, Houston.
Mittlefehldt, D. W., Gellert. R., VanBommel, S., Ming, D. W., Yen, A.S., Clark, B. C., et
al. (2018). Diverse lithologies and alteration events on the rim of Naochian-aged
Endeavour crater, Meridiani Planum, Mars: In situ compositional evidence. Journal of
Geophysical Research, Planets, 123, 1255-1306. https://doi.org/10.1002/2017JE005474
Morris, R.V., Vaniman, D.T., Blake, D.F., Gellert, R., Chipera, S.J., Rampe, E.B., et al.
(2016). Silicic volcanism on Mars evidenced by tridymite in high-SiO2 sedimentary rock
at Gale crater. Proceedings of the National Academy of Science, 113, 7071–7076.
https://doi.org/10.1073/pnas.1607098113
Morrison, S. M., Downs, R. T., Blake, D. F., Vaniman, D. T., Ming, D. W., Hazen, R. M.,
et al. (2018). Crystal chemistry of martian minerals from Bradbury Landing through
Naukluft Plateau, Gale crater, Mars. American Mineralogist, 103(6), p.857-871.
Doi.org/10.2138/am-2018-6124

©2020 American Geophysical Union. All rights reserved.


Mukhopadhyay, J., Crowley, Q.G., Ghosh, S., Ghosh, G., Chakrabarti, K., Misra, B.,
Heron, K., & Bose, S. (2014). Oxygenation of the Archean atmosphere: new paleosol
constraints from eastern India. Geology, 42(10), 923–926. doi.org/10.1130/ G36091.1.
Nachon, M., Clegg, S. M., Mangold, N., Schröder, S., Kah, L. C., Dromart, G., et al.
(2014). Calcium sulfate veins characterized by ChemCam/Curiosity at Gale crater,
Mars. Journal of Geophysical Research, Planets, 119, 1991–2016.
https://doi.org/10.1002/ 2013JE004588
Nachon, M., Mangold, N., Forni, O., Kah, L. C., Cousin, A., Wiens, R. C., et al. (2016).
Chemistry of diagenetic features analyzed by ChemCam at Pahrump Hills, Gale crater,
Mars. Icarus, 281, 121–136. doi.org/10.1016/j.icarus.2016.08.026
Nesbitt, H.W. (1979). Mobility and fractionation of rare earth elements during weathering
of a granodiorite. Nature 279, 206–210.
Nesbitt, H.W., & Wilson, R.E. (1992). Recent chemical weathering of basalt. American
Journal of Science, 292, 740–777.
Nesbitt, H. W., & Young, G. M. (1982). Early Proterozoic climates and plate motions
inferred from major element chemistry of lutites. Nature, 299(5885), 715–717.
doi.org/10.1038/299715a0
O’Connell-Cooper, C.D., Spray, J.G., Thompson, L.M., Gellert, R., Berger, J.A., Boyd,
N.I., et al. (2017). APXS-derived chemistry of the Bagnold dune sands: Comparisons
with Gale Crater soils and the global Martian average. Journal of Geophysical Research
Planets, 122(12):2623-2643, doi:10.1002/2017JE005268
Papike, J. J., Karner, J. M. & Shearer C. K. (2003) Determination of planetary basalt
parentage: A simple technique using the electron microprobe. American Mineralogist,
88(2-3), 469-472. doi.org/10.2138/am-2003-2-323
Postma, D., (1985). Concentration of Mn and separation from Fe in sediments – I. Kinetics
and stoichiometry of the reaction between birnessite and dissolved Fe(II) at 10°C.
Geochimica et Cosmochimica Acta, 49(4), 1023-1033. doi.org/10.1016/0016-
7037(85)90316-3
Postma, D., & Appelo, C. A. J. (2000). Reduction of Mn-oxides by ferrous iron in a flow
system: Column experiment and reactive transport modeling. Geochimica et
Cosmochimica Acta, 64(7), 1237-1247. doi.org/10.1016/S0016-7037(99)00356-7
Rampe, E.B., Ming, D.W., Blake, D.F., Bristow, T.F., Chipera, S.J., Grotzinger, J.P., et al.
(2017). Mineralogy of an ancient lacustrine mudstone succession from the Murray
formation, Gale crater, Mars. Earth and Planetary Science Letter,s 471, 172– 158.
doi.org/10.1016/j.epsl.2017.04.021
Rampe, E. B., Bristow, T. F., Morris, R. V., Morrison, S. M., Achilles, C. N., Ming, D. W.,
et al. (2020a). Mineralogy of Vera Rubin Ridge from the Mars Science Laboratory
CheMin Instrument. Journal of Geophysical Research-Planets.
doi:10.1029/2019JE006306
Rampe, E. B., Blake, D.F., Bristow, T. F., Ming, D. W., Vaniman, D. T., Morris, R. V., et
al. (2020b). Mineralogy and geochemistry of sedimentary rocks and eolian sediments in
Gale crater, Mars: A review after six Earth years of exploration with Curiosity.
Geochemistry, 80(2). doi:10.1016/j.chemer.2020.125605

©2020 American Geophysical Union. All rights reserved.


Rapin, W., Ehlmann, B. L., Dromart, G., Schrieber, J., Thomas, N. H., Fischer, W. W., et
al. (2019). An interval of high salinity in ancient Gale crater lake on Mars. Nature
Geoscience, 12, p.889-895. Doi.org//10.1038/s41561-019-0458-8
Rieder, R., Gellert, R., Brückner, J., Klingelhöfer, G., Dreibus, G., Yen, A. & Squyres, S.
W. (2003) The new Athena alpha particle X‐ray spectrometer for the Mars Exploration
Rovers. Journal of Geophysical Research, Planets, 108, E12, 8066.
https://doi.org/10.1002/2015GL066675
Sánchez-España, J., Yusta, I., & Burgos, W. D. (2016). Geochemistry of dissolved
aluminum at low pH: Hydrobasaluminite formation and interaction with trace metals,
silica and microbial cells under anoxic conditions. Chemical Geology, 441, 124-137.
doi.org/10.1016/j.chemgeo.2016.08.004
Sautter, V., Toplis, M. J., Wiens, R. C., Cousin, A., Fabre, C., & Gasnault, O. (2015). In
situ evidence for continental crust on early Mars. Nature Geoscience, 8(8), 605.
Schmidt, M. E., Campbell, J. L., Gellert, R., Perrett, G. M., Treiman, A. H., Blaney, D. L.
et al. (2014) Geochemical diversity in first rocks examined by the Curiosity rover in
Gale crater: Evidence for and significance of an alkali and volatile-rich igneous source,
Journal of Geophysical Research Planets, 119(1):64-81, doi:10.1002/2013JE004481
Schmidt, M. E., Perrett, G. M., Bray, S. L., Bradley, N. J., Lee, R. E., Berger, J. A.,
Campbell, J. L., Ly, C., Squyres, S. W., & Tesslaer, D. (2018). Dusty Rocks in Gale
Crater: Assessing areal coverage and separating dust and rock contributions in APXS
analyses. Journal of Geophysical Research, Planets, 123(7), 1649-1673.
doi.org/10.1029/2018JE005553
Siebach, K. L., Baker, M. B., Grotzinger, J. P., McLennan, S. M., Gellert, R., Thompson,
L. M., & Hurowitz, J. A. (2017). Sorting out Compositional Trends in Sedimentary
Rocks of the Bradbury Group (Aeolus Palus), Gale Crater, Mars. Journal of
Geophysical Research, Planets, 122, 295-328. doi/10.1002/2016JE005195
Stack, K.M., Grotzinger, J.P., Lamb, M.P., Gupta, S., Rubin, D.M., Kah, L.C., et al.
(2019). Evidence for plunging river plume deposits in the Pahrump Hills member of the
Murray formation, Gale crater, Mars. Sedimentology 66, 1768–1802.
https://doi.org/10.1111/sed.12558
Stein, N.T., Quinn, D.P., Grotzinger, J.P., Fedo, C., Ehlmann, B.L., Stack, K. M., et al.
(2020). Regional structural orientation of the Mt. Sharp group revealed by in-situ dip
measurements and stratigraphic correlations on the Vera Rubin ridge. Journal of
Geophysical Research, Planets, 125(5), doi: 10.1029/2019JE006298
Stolper, E. M., Baker, M. B., Newcombe, M. E., Schmidt, M. E., Treiman, A. H., Cousin,
A., et al. (2013). The petrochemistry of Jake M: A martian mugearite. Science 341
(6153), 1239463. doi:10.1126/science.1239463
Stumm, W. & Morgan, J. J. (1996). Aquatic Chemistry: Chemical equilibria and rates in
natural waters. New York, NY: John Wiley, 1022 pp.
Sun, V.Z., Stack, K.M., Kah, L.C., Thompson, L., Fischer, W., Williams, et al. (2019).
Late-stage diagenetic concretions in the Murray formation, Gale crater, Mars. Icarus
321, 866–890. https://doi.org/10.1016/j.icarus.2018.12.030
Sutter, B., McAdam, A.C., Rampe, E.B., Thompson, L.M., Ming, D.W., Mahaffy, P.R. et
al. (2017). Evolved gas analyses of the Murray formation in Gale crater, Mars: Results
of the Curiosity rover’s Sample Anlaysis at Mars (SAM) instrument. Paper presented at

©2020 American Geophysical Union. All rights reserved.


48thLunar and Planetary Science Conference, Abstract 3009, Lunar and Planetary
Institute, Houston
Thompson, L. (2020). Alpha Particle X-ray spectrometer geochemistry of the Murray
formation and Vera Rubin ridge, Gale crater, Mars. https://doi.org/10.25545/ZXDJZ7,
UNB, V1, UNF:6:bL/a2qTZBu6DNJlzkmGAbg==[fileUNF]
Thompson, L. M., Gellert, R., Spray, J. G., Kah, L. C. & the APXS and MSL Science
Teams (2015a). The composition of the basal Murray formation at Pahrump hills, Gale
crater, Mars. Paper presented at 46thLunar and Planetary Science Conference, Abstract
1429, Lunar and Planetary Institute, Houston
Thompson, L. M., Gellert, R., Spray, J. G., Berger, J. A., Boyd, N., Campbell, J. L.,
deSouza, E., Pavri, B., Perrett, G. M., VanBommel, S. & Yen, A. S. (2015b).
Compositions of sedimentary strata, nodular features and veins at the base of Mount
Sharp, Gale crater, Mars: an APXS perspective. Paper presented at European Planetary
Science Congress, Vol. 10, EPSC2015-827. Nantes, France
Thompson, L. M., Schmidt, M. E., Spray, J. G., Berger, J. A., Fairén, A. G., Campbell J.
L., et al. (2016). Potassium-rich sandstones within the Gale impact crater, Mars: The
APXS perspective. Journal of Geophysical Research Planets, 121(10):1981-2003,
doi:10.1002/2016JE005055
Thompson, L. M., Yen, A., Spray, J. G., Johnson, J. R., Fraeman, A. A., Berger, J. A.,
Gellert, R., Boyd, N., Desouza, E., O'Connell-Cooper C. & VanBommel, S. (2017).
Recent compositional trends within the Murray formation, Gale crater, Mars, as seen by
APXS: Implications for sedimentary, diagenetic and alteration history. Paper presented
at American Geophysical Union, Fall Meeting 2017, abstract #P31A-0846.
Tobler, D. J., Stawski, T. M., & Benning L. G. (2017). Silica and Alumina Nanophases:
Natural Processes and Industrial Applications. In A.E.S. Van Driessche et al. (Eds.),
New Perspectives on Mineral Nucleation and Growth (pp.293-316). Springer
International Publishing Switzerland doi:10.1007/978-3-319-45669-015
Tosca, N. J, McLennan, S. M, Hindsley, D. H., & Schoonen, M. A. A. (2004). Acid-sulfate
weathering of synthetic Martian basalt; the acid fog model revisited. Journal of
Geophysical Research, 109(E5), doi:10.1029/2003JE00221
Tosca , N. J., & Knoll, A. H. (2009). Juvenile chemical sediments and the long term
persistence of water at the surface of Mars. Earth and Planetary Science Letters, 286,
379-386.
Treiman, A. H., Bish, D. L., Vaniman, D. T., Chipera, S. J., Blake, D. F., Ming, D. W., et
al. (2016). Mineralogy, provenance and diagenesis of a potassic basaltic sandstone on
Mars: CheMin X-ray diffraction of the Windjana sample (Kimberley area, Gale Crater).
Journal of Geophysical Research, 121(1), p.75 – 106. doi:10.1002/2015JE004932
Turner, S. M. R., Schwenzer, S. P., Bridges, J. C., Rampe, E. B., Bedford, C. C., Achilles,
C. N., et al. (2020). Fluid-Rock Reactions in the Murray Formation, including Vera
Rubin ridge, Gale crater. Submitted to Journal of Geophysical Research
Urban, N. R., Gorham, E., Underwood, J. K., Martin, F. B., & Ogden III, J. G. (1990)
Geochemical processes controlling concentrations of Al, Fe and Mn Nova Scotia lakes.
Limnology and Oceanography 35(7), p. 1516-1534.
VanBommel, S.J., Gellert, R., Berger, J.A., Campbell, J.L., Thompson, L.M., Edgett, K.S.,
et al. (2016). Deconvolution of distinct lithology chemistry through oversampling with

©2020 American Geophysical Union. All rights reserved.


the Mars Science Laboratory Alpha Particle X-ray Spectrometer. X-Ray Spectrometry,
45, 155–161. https://doi.org/10.1002/xrs.2681
VanBommel, S.J., Gellert, R., Berger, J.A., Thompson, L.M., Edgett, K.S., McBride, M.J.,
et al. (2017). Modeling and mitigation of sample relief effects applied to chemistry
measurements by the Mars Science Laboratory Alpha Particle X-ray Spectrometer. X-
Ray Spectrometry, 46, 229–236. https://doi.org/10.1002/xrs.2755
VanBommel, S.J., Gellert, R., Boyd, N.I., & Hanania, J.U. (2019). Empirical simulations
for further characterization of the Mars Science Laboratory Alpha Particle X-ray
Spectrometer: An Introduction to the ACES program. Nuclear Instrumentation and
Methods, B 441, 79–87. https://doi.org/10.1016/j.nimb.2018.12.040
Vaniman, D.T., Martínez, G.M., Rampe, E.B., Bristow, T.F., Blake, D.F., Yen, A.S. et al.
(2018). Gypsum, bassanite, and anhydrite at Gale crater, Mars. American Mineralogist,
103:1011-1020, doi:10.2138/am-2018-6346
Williams, R.M.E., Grotzinger, J.P., Dietrich, W.E., Gupta, S., Sumner, D.Y., Wiens, R.C.,
et al. (2013). Martian fluvial conglomerates at Gale crater. Science 340, 10681072.
https://doi.org/10.1126/science.1237317
Wintsch, R. P., & Kvale, C. M. (1994). Differential mobility of elements in burial
diagenesis of siliciclastic rocks. Journal of Sedimentary Research, A64(2a), 349-361.
doi.org/10.1306/D4267D9D-2B26-11D7-8648000102C1865D
Wong, G. M., Lewis, J. M. T., Knudson, C. A., Millan, McAdam, A. C., Eigenbrode, J. L.,
et al. (2020). Detection of reduced sulfur on Vera Rubin ridge by quadratic discriminant
analysis of volatiles detected during evolved gas analysis. Journal of Geophysical
Research Planets, doi:10.1029/2019JE006304
Yen, A.S., Clark, B. C., Ming, D. W., Mittlefehldt, D. W., Gellert, R. & Morris, R. V.
(2010) Chemical alteration on Mars indicated by the iron-manganese ratio. Paper
presented at 41st Lunar and Planetary Science Conference, Abstract 2546, Lunar and
Planetary Institute, Houston
Yen, A.S., Morris, R.V., Gellert, R., Berger, J. A., Sutter, B., Downs, R.T., et al. (2017a).
Hydrothermal signatures at Gale crater, Mars, and possible in-situ formation of
tridymite. Paper presented at American Geophysical Union, Fall Meeting 2017, abstract
#P24B-04, New Orleans.
Yen, A.S., Ming, D.W., Vaniman, D.T., Gellert, R., Blake, D.F., Morris, R.V., et al.
(2017b). Multiple stages of aqueous alteration along fractures in mudstone and
sandstone strata in Gale crater, Mars. Earth and Planetary Science Letters, 471, 186–
198. https://doi.org/10.1016/j.epsl.2017.04.033
Yen, A.S., Gellert, R., Thompson, L.M., Treiman, A. H., Morris, R.V., Vaniman, et al.,
(2018). Mobility of potassium-rich fluids on Mars: Implications for diagenesis. Paper
presented at 49th Lunar and Planetary Science Conference, Abstract 2690, Lunar and
Planetary Institute, Houston.

©2020 American Geophysical Union. All rights reserved.


Figure 1a. Mars Orbiter Laser Altimeter (MOLA) topographic map of Mars showing the
location of Curiosity in Gale crater. 1b: Oblique, computer-generated view of Gale crater
based on multiple orbital observations (Thermal Emission Imaging System camera on Mars
Odyssey, MOLA on Mars Global Surveyor orbiter., and High Resolution Imaging Science
Experiment (HiRISE) camera on Mars Reconnaissance Orbiter). MSL landing ellipse and
traverse overlain (credit:NASA/JPL-Caltech/ASU/UA). 1c: HiRISE map showing Curiosity’s
traverse, points/areas of interest and the division of the Bradbury and Mount Sharp groups
(credit:NASA/JPL-Caltech/ASU/UA).

©2020 American Geophysical Union. All rights reserved.


Figure 2. Color-stretched HiRISE image with Curiosity’s traverse on the Vera Rubin ridge.
White dots denote end of drive locations where APXS data was not acquired. Colored dots
and stars correspond with locations where APXS data was acquired. Colors correspond to the
respective Murray member or grouping (see legend). Note the numerous patches of gray/blue
material within the upper VRR Jura member, and less commonly in the Pettegrove Point
member (blue dots correspond with locations where APXS was acquired within these areas).
The four drill locations on the ridge and in the underlying Blunt’s Point member are also
labelled (white outlined dots); Duluth, Stoer, Highfield and Rock Hall. Thicker, black dashed
lines delineate VRR. Thinner, black dashed line delineates the Pettegrove Point/Jura contact.
Thick black arrow indicates upsection direction. (credit: NASA/JPL-Caltech/ASU/UA)

©2020 American Geophysical Union. All rights reserved.


Figure 3. Stratigraphic column for Gale crater up to and including VRR. Colored dots
indicate drill holes within the Murray formation, and are shown at the elevation level they
were drilled. Colors correspond with the color of the respective Murray member shown in the
column. Black dots correspond with drill holes sampled from other Gale bedrock (i.e.,
Stimson formation and Bradbury group). (credit: MSL Sedimentology and Stratigraphy
Working Group)

©2020 American Geophysical Union. All rights reserved.


Figure 4. Figure 4: Total alkalis versus silica (TAS) plot (Le Maitre et al., 2006) for all Gale
rock targets analyzed by APXS (not including targets with obvious CaSO4 veins or
diagenetic/alteration features; except the high-Si targets associated with alteration halos
within the Stimson formation to facilitate a comparison with high-Si Murray formation). All
data normalized to S and Cl free. YKB – Yellowknife Bay formation mudstones. Trends
away from basaltic compositions highlighted by arrows (Thompson et al., 2016). Caveat: The
majority of rocks sampled by APXS are sedimentary and not igneous. This plot is used to
represent the data in a way that has commonly be used for martian data, and does not imply
that samples are igneous, only that they have certain compositions.

©2020 American Geophysical Union. All rights reserved.


Figure 5: Stratigraphic column for the Murray formation up to and including the VRR (top
left), and APXS elemental and oxide concentrations, and FeO/MnO (bottom left) plotted with
elevation (m) for the Murray formation and VRR. All concentrations are wt% except, Ni, Zn
and Br, which are ppm. Targets with obvious diagenetic/alteration features and/or veins are
not included. High-Si Pahrump Hills targets are also not shown for most plots, i.e., they are
off the scale. Symbols are colored to match their corresponding member in the stratigraphic
column. Dashed dark red vertical lines represent average martian basaltic soil (O’Connel-
Cooper et al., 2017). Solid blue vertical lines represent mean Murray formation bedrock
below the ridge, and the pale blue shaded boxes the standard deviation. Solid black lines are
5-point moving average trend lines. The white filled, red outlined stars on the FeO plot
represent bedrock targets analyzed by APXS from within the areas exhibiting the deepest
hematite absorptions from orbital CRISM data. The horizontal black line on the FeO/MnO
plot represents the typical error associated with the lower Mn, high (>190) FeO/MnO points.
Grey filled, black outlined points at the highest elevation on each plot are the targets analyzed
as Curiosity descended off the ridge, into the Glen Torridon area.

©2020 American Geophysical Union. All rights reserved.


Figure 6. Box and whisker plots for the Murray formation members (PH – Pahrump Hills,
TP – Telegraph Peak, HV – Hartmann’s Valley, K – Karasburg, SI – Sutton Island, BP –
Blunts Point, PPt – Pettegrove Point, J – Jura, RH – Rock Hall, g/b – gray/blue). X – mean,
horizontal line – median, boxes – interquantile range, whiskers – minimum and maximum
values, circles – ouliers; dotted vertical line marks the start of VRR.

©2020 American Geophysical Union. All rights reserved.


Figure 7. Plot showing the results of a comparison of SiO2, Al2O3, MnO, FeO and TiO2
concentrations for the Murray member targets relative to mean Murray formation. Except for
the high Si and TP (Telegraph Peak and associated) targets, analyses are shown with
increasing elevation from left to right. The blue/gray patches occur over the same elevation
range as the typical Jura. Note that the width of each member/subdivision is based on the
number of analyses and not proportional to the thickness of the unit. RH – Rock Hall. The
bars on the far, right side of the plot show the Murray standard deviations for the oxides
plotted. See §S1.4 for methodology.

©2020 American Geophysical Union. All rights reserved.


Figure 8. Plots showing a) positive correlation of Al2O3/TiO2 and SiO2 /TiO2 for the VRR
compared to the rest of the Murray formation; b) negative correlation ofAl2O3 /TiO2 and FeO
/TiO2 for the Murray formation below VRR, and VRR; c) higher SiO2/TiO2 for the blue/grey
patches than the rest of the Murray formation; d) higher Al2O3/TiO2 for the blue/grey patches
than the rest of the Murray formation except a number of Pahrump Hills targets. Black
crosses represent typical error bars. The outlier targets (low Si, Al and high Fe) within the
typical Jura (red diamonds) are the Rock_Hall drill fines and associated targets. PH –
Pahrump Hills member (not including high Si targets), HV – Hartmann’s Valley member, K
– Karasberg member, SI – Sutton Island member, BP – Blunts Point member, PP –
Pettegrove Pt member, J-T – Jura member (typical), J-G – Jura member (grey/blue includes
one target from PP), CBU – clay-bearing unit

©2020 American Geophysical Union. All rights reserved.


Figure 9. Murray formation bedrock targets below the ridge (black outline) and VRR targets
(shaded grey) ratioed (logarithmic scale) to a typical Murray formation target, Ganda (sol
1444). High Si Murray targets, and targets with prominent veins or diagenetic features not
included. Arrows highlight low Ti, Mn, Fe and high Ni concentrations within some targets on
VRR relative to other Murray targets.

©2020 American Geophysical Union. All rights reserved.


Figure 10. Plot showing the results of a mass change comparison of SiO2, Al2O3, MnO and
FeO for the Jura and gray/blue bedrock targets relative to mean Pettegrove Point. For each
group, elevation increases from right to left. The elevations of the gray/blue targets overlap
with the typical Jura. The bars on the far, right side of the plot show the standard deviations
for SiO2/TiO2, Al2O3/TiO2 and FeO/TiO2 for Pettegrove Point. RH – Rock Hall. See §S1.5
for methodology.

©2020 American Geophysical Union. All rights reserved.


Table 1: APXS compositional data for all targets analyzed on VRR; all concentrations are reported as wt% unless otherwise indicated (see also Table S1)

Target name (corresponds with Target Ni/ Zn/ Br/


Sol Notes/features SiO2 TiO2 Al2O3 Cr2O3 MnO FeO MgO CaO Na2O K2O P2O5 SO3 Cl FeO/MnO
NASA PDS) Type1 ppm ppm ppm

Pettegrove Point member

1811 Sasanoa_DRT RB 51.79 1.16 9.52 0.34 0.12 22.52 4.18 2.28 2.13 0.91 0.99 2.78 0.88 982 2108 176 188

1811 Kemps_Folly R 50.48 1.12 9.59 0.32 0.10 21.33 5.04 2.73 2.31 0.90 0.97 3.80 0.93 1101 1946 85 213

1814 Pumpkin_Nob R 47.00 1.14 8.79 0.32 0.12 20.98 5.94 3.73 2.59 0.79 0.85 5.25 2.10 1034 1320 447 175

1816 Schoppee R 50.06 1.09 9.57 0.32 0.09 22.08 4.69 2.46 2.44 0.81 0.98 3.78 1.33 949 1142 116 245

1818 Christmas_Cove_preDRT R 45.27 1.08 8.65 0.33 0.13 20.83 6.68 4.45 2.69 0.75 0.84 6.08 1.93 924 1044 287 160

1818 Christmas_Cove_DRT RB 47.06 1.07 9.05 0.32 0.11 20.55 4.27 3.76 3.08 0.86 0.82 6.04 2.64 949 1054 272 187

1818 Mitten_Ledge_DRT RB 49.63 1.14 9.30 0.32 0.07 21.50 4.63 2.68 2.65 0.90 0.91 4.00 1.96 926 1159 156 307

1821 Pennessewassee R 46.27 1.04 8.66 0.30 0.13 20.14 5.99 4.38 2.54 0.81 0.96 6.66 1.77 991 1013 555 155

1821 Passadumkeag_DRT RB 47.19 1.07 9.01 0.31 0.12 18.78 4.83 4.40 2.75 0.98 0.86 6.59 2.78 920 941 379 157

1824 Troll_Valley R nodules 45.41 1.02 8.69 0.30 0.30 19.08 6.53 5.57 2.49 0.78 0.93 7.47 1.10 777 916 187 64

1824 Sherwood_Forest_DRT RB 46.65 1.05 8.96 0.30 0.25 18.03 5.69 5.19 2.37 0.96 0.93 8.46 0.88 808 962 323 72

1829 Enon R nodules 45.92 1.16 8.78 0.36 0.23 19.89 6.73 4.90 2.53 0.69 0.89 6.83 0.85 716 1045 68 86

1830 Collingham R 42.09 1.05 8.01 0.33 0.09 21.96 6.48 5.45 2.77 0.62 1.00 8.23 1.69 767 732 452 244

1834 Katberg R 46.64 1.04 8.84 0.27 0.23 18.60 5.64 4.62 2.88 0.96 0.89 7.13 2.01 881 935 223 81

1836 Lucknow_DRT RB 45.34 1.04 8.64 0.30 0.26 20.92 5.47 3.93 2.63 1.09 1.14 6.65 2.31 752 1239 146 80

1836 Ecca_DRT RB 47.27 1.03 9.18 0.29 0.26 20.00 5.48 4.01 2.44 1.02 1.21 6.42 1.10 738 1182 88 77

1838 Cheshire R resistant features 47.89 1.09 8.88 0.34 0.25 19.55 5.79 4.57 2.30 1.01 0.89 6.20 0.92 829 1374 176 78

1838 Duitschland R 46.38 1.09 9.00 0.32 0.37 18.27 6.28 5.09 2.53 0.93 0.83 7.44 1.15 868 1410 83 49

1845 Stormberg R 44.90 1.06 8.85 0.33 0.34 19.96 6.83 4.59 2.72 0.87 0.96 6.38 1.88 805 1331 140 59

1853 Balfour R 49.79 1.07 9.30 0.30 0.30 20.59 5.46 3.26 2.42 1.06 0.89 4.57 0.73 949 968 96 69

©2020 American Geophysical Union. All rights reserved.


1863 Gamka R 45.87 0.95 9.00 0.33 0.21 22.41 6.15 3.89 2.40 0.88 1.03 5.45 1.04 781 722 119 107

1863 Sibasa_DRT RB 48.36 1.06 9.00 0.30 0.30 19.32 5.39 4.33 2.38 1.04 0.85 6.60 0.82 884 896 80 64

1865 Barberton_raster1 R resistant features 44.35 1.00 9.15 0.36 0.30 18.31 6.80 6.92 2.60 0.75 0.86 7.76 0.64 702 668 41 61

1865 Barberton_raster2 R resistant features 44.47 0.97 8.68 0.32 0.27 18.27 6.01 6.06 2.38 0.94 0.90 9.21 1.05 841 819 69 68

1865 Barberton_raster3 R resistant features 43.50 0.93 8.74 0.31 0.20 18.56 6.17 6.86 2.34 0.87 0.81 9.60 0.78 696 852 90 93

1865 Campbellrand R 46.59 1.10 9.13 0.32 0.26 20.42 6.79 4.02 2.55 0.89 1.00 5.72 0.94 834 939 92 79

1868 Volksrust R resistant features 44.63 1.06 8.70 0.33 0.39 19.54 6.65 5.75 2.54 0.91 0.94 7.39 0.87 852 981 353 50

1870 Waboomberg R 41.81 0.92 7.83 0.30 0.37 24.37 6.34 4.66 2.56 0.85 1.34 5.66 2.72 721 688 516 66

1870 Platberg_DRT RB 45.00 1.00 8.49 0.30 0.38 17.18 5.68 6.09 2.82 0.94 0.80 9.21 1.84 739 972 80 45

2000 Sgurr_of_Eigg_DRT RB 48.69 1.01 9.20 0.32 0.33 19.05 6.01 4.02 2.66 1.00 0.89 5.40 1.06 836 1014 40 58

2001 Appin R 46.32 0.96 8.75 0.26 0.32 17.46 5.34 6.05 2.53 0.96 0.90 8.82 1.07 722 887 34 55

2001 Brora_DRT RB 45.64 0.98 8.55 0.30 0.31 17.39 5.37 6.43 2.42 0.91 0.88 9.54 1.03 750 951 78 56

2005 Murchison R Deepest hematite absorption, CRISM 46.99 1.06 8.96 0.30 0.25 20.39 6.01 3.54 2.82 1.03 0.80 5.66 1.90 1003 1043 71 82

2005 Stranraer_DRT RB Deepest hematite absorption, CRISM 47.57 1.05 8.78 0.31 0.25 20.68 4.99 3.40 2.85 1.10 0.79 5.55 2.40 1035 1080 82 83

2008 Dun_Caan R Deepest hematite absorption, CRISM 46.25 1.04 9.49 0.28 0.20 21.76 5.49 2.97 2.80 1.06 1.49 4.67 2.25 865 869 27 109

2008 Lanark_DRT RB Deepest hematite absorption, CRISM 47.31 1.00 8.92 0.30 0.25 20.47 5.59 3.30 2.69 1.05 0.99 5.83 2.05 1011 881 35 82

2013 Kinloch Re pebble and sand 35.54 0.79 7.09 0.27 0.25 30.64 6.08 3.25 2.79 0.70 2.13 7.90 2.12 747 985 503 123

2013 Lingarabay_DRT RB few nodules 51.08 1.03 9.80 0.30 0.27 18.10 5.05 3.87 2.64 1.03 0.87 4.94 0.74 921 952 56 67

2014 Corsehill R nodules 45.18 1.00 8.97 0.36 0.31 17.12 6.26 7.18 2.52 0.73 0.96 8.49 0.70 711 694 61 55

2029 Babbitt R Deepest hematite absorption, CRISM 47.19 1.07 9.07 0.32 0.20 19.72 4.99 3.33 2.98 0.96 0.80 6.65 2.45 992 1093 38 99

2032 Pokegama R Deepest hematite absorption, CRISM 46.95 1.08 9.56 0.31 0.23 19.19 6.02 3.55 2.73 1.01 0.77 6.46 1.86 859 1145 35 83

2038 Nashwauk R 46.81 1.05 9.09 0.32 0.10 21.90 5.38 3.15 2.76 0.82 1.11 5.48 1.70 838 1229 148 219

2042 Pigeon_River_DRT RB 45.13 1.04 8.62 0.31 0.12 20.00 4.74 4.83 2.74 0.82 0.91 7.90 2.50 939 1307 201 167

2042 Bald_Eagle_Lake_DRT RB 43.64 1.01 8.58 0.32 0.12 21.28 4.70 4.65 2.64 0.81 1.25 7.99 2.68 865 1164 160 177

2100 Dumbarton_Rock_DRT RB 47.05 1.07 8.87 0.32 0.29 20.56 5.05 4.05 2.52 1.07 0.75 6.38 1.67 1007 1243 94 71

©2020 American Geophysical Union. All rights reserved.


2101 Duntarvie_Castle R 44.12 1.04 8.55 0.34 0.28 20.73 6.19 4.86 2.48 0.91 0.83 7.83 1.54 976 1277 135 74

2101 Duntelchaig R 42.25 1.04 8.33 0.33 0.32 20.21 6.73 5.65 2.68 0.85 0.84 8.57 1.83 1018 1284 159 63

2104 Oskaig R 46.08 1.05 8.77 0.34 0.35 20.49 6.50 3.92 2.52 0.86 0.95 6.13 1.65 1067 1211 78 59

2107 Chippewa R float with resistant features 43.03 0.91 8.30 0.28 0.42 17.91 5.89 6.23 2.84 1.00 1.07 9.44 2.19 713 1569 140 43

2108 Orr R dark patchy 43.69 0.98 8.57 0.28 0.19 24.31 5.63 3.09 2.72 0.70 2.46 5.07 2.00 903 796 60 128

2109 Voyageurs_DRT RB 43.97 1.00 8.28 0.32 0.25 20.88 5.00 4.86 2.79 0.97 0.75 8.24 2.34 1009 1095 60 84

2113 Voyageurs_DRT_Drill_Tailings DT failed drill attempt 46.58 1.07 8.88 0.32 0.27 22.06 4.76 4.20 2.56 1.06 0.72 5.78 1.33 1081 1189 56 82

2113 Voyageurs_DRT_post_Drill MD failed drill attempt 47.49 1.13 9.01 0.34 0.27 21.72 4.32 4.45 2.44 1.09 0.67 6.07 0.69 1072 1175 57 80

2117 Walsay R 47.65 1.04 8.89 0.30 0.25 19.23 5.44 4.36 2.64 1.03 0.89 6.62 1.38 859 892 46 77

2121 Ailsa_Craig_offset RB 47.92 0.99 9.02 0.29 0.34 19.18 5.79 4.57 2.52 0.98 0.92 6.20 0.94 834 1002 37 56

2122 Ailsa_Craig_DRT RB 45.85 0.97 8.65 0.28 0.31 17.94 5.33 6.28 2.46 0.93 0.92 8.87 0.99 749 961 36 58

2124 Ailsa_Craig_post_Drill MD failed drill attempt 47.90 1.15 9.02 0.35 0.37 21.92 4.43 4.92 2.56 1.05 0.75 4.82 0.43 942 1233 37 59

2124 Ailsa_Craig_Drill_Tailings DT failed drill attempt 48.76 1.09 9.26 0.34 0.36 21.48 4.67 4.67 2.62 1.06 0.81 4.07 0.49 909 1136 36 60

2127 Diabaig R 50.03 1.02 9.46 0.31 0.23 18.04 5.23 3.34 2.88 1.06 0.82 5.13 2.15 959 1121 94 78

2127 Slioch R dark patch 44.76 0.97 8.46 0.28 0.33 18.41 6.57 4.88 2.42 0.87 0.94 8.93 1.90 1045 1127 109 56

2131 Mount_Battock R 50.14 0.99 9.62 0.31 0.20 21.46 6.20 2.74 2.70 0.88 1.38 2.40 0.69 880 886 499 107

2131 Dobbs_Linn_DRT RB 48.71 0.97 9.32 0.31 0.23 17.74 4.98 5.28 2.53 0.94 0.92 7.01 0.81 828 1037 121 77

2131 Scourie_More R grey/blue from orbit 53.98 1.00 10.39 0.30 0.25 16.43 5.73 3.33 2.71 1.02 0.89 3.02 0.66 933 861 350 66

2134 Stoer_DRT RB 47.88 1.02 8.89 0.32 0.09 18.46 5.02 5.27 2.42 0.93 0.81 7.76 0.80 802 857 429 205

2134 Stoer_offset RB 46.39 1.00 8.86 0.32 0.11 17.51 5.29 6.27 2.53 0.87 0.83 8.86 0.88 739 787 254 159

2154 Stoer_dump_centre DBA 44.08 1.03 8.75 0.34 0.17 21.53 4.72 6.44 2.45 0.91 0.84 7.59 0.89 914 847 44 127

2155 Stoer_tailings DT 45.95 1.01 8.49 0.38 0.20 23.15 4.79 5.36 2.44 0.87 0.97 5.49 0.68 835 771 70 116

Jura member

1875 Middleton R 46.16 1.15 9.15 0.32 0.31 20.27 5.79 4.28 2.70 1.16 0.73 5.68 2.04 822 1096 29 65

1875 Fort_Brown_DRT RB 46.04 1.05 8.75 0.30 0.41 19.87 5.54 4.71 2.38 1.02 0.94 7.89 0.84 825 960 117 48

©2020 American Geophysical Union. All rights reserved.


1879 Zululand R 49.47 1.00 9.41 0.28 0.26 17.21 5.25 4.22 2.59 1.04 0.85 6.67 1.56 642 650 41 66

1879 Hexriver_DRT RB 49.33 1.00 9.29 0.28 0.26 17.94 5.56 3.28 2.87 1.15 0.85 6.18 1.76 838 903 54 69

1885 Klippan_DRT RB 50.05 1.01 9.36 0.29 0.32 17.48 5.61 3.73 2.49 1.09 0.84 6.23 1.23 920 987 44 55

1885 Klipfonteinheuwel R 50.37 0.98 9.98 0.28 0.24 16.57 5.58 3.59 2.74 1.00 0.90 5.85 1.69 903 822 46 69

1889 Lyttelton R concordant CaSO4? 44.95 0.92 8.92 0.25 0.15 16.24 5.15 6.50 2.48 0.94 0.98 10.72 1.40 752 756 115 108

1892 Strubenkop R 52.28 1.00 10.21 0.29 0.09 15.52 5.35 3.47 2.72 1.17 0.72 5.75 1.23 867 548 24 172

1893 Drakensberg_DRT RB 50.01 1.04 9.93 0.30 0.10 16.85 5.53 3.70 2.86 1.17 0.81 5.74 1.72 1004 589 56 169

1895 Mzamba R 47.82 1.03 8.94 0.29 0.38 20.52 5.75 2.99 2.47 1.17 0.83 5.62 1.88 993 1304 112 54

1897 Wick R 48.64 1.07 9.59 0.30 0.19 17.04 5.57 4.47 2.58 1.04 0.73 6.86 1.67 886 717 20 90

1897 Muck R 51.52 1.04 10.13 0.29 0.16 16.79 5.19 3.33 2.78 0.95 0.77 5.30 1.49 639 521 64 105

1904 Talisker R 49.40 0.91 9.73 0.29 0.10 16.78 5.42 4.40 2.84 0.79 0.78 7.16 1.17 878 812 79 168

1904 Oban_DRT RB 52.54 0.94 10.11 0.32 0.09 15.91 5.11 3.60 2.62 0.80 0.77 5.66 1.24 1029 914 86 177

1906 Holyrood R 48.85 1.02 9.52 0.32 0.14 15.98 5.23 4.38 2.67 1.08 0.61 8.44 1.50 846 838 118 114

1907 Haddo_House R 49.26 1.00 9.55 0.28 0.15 15.51 5.37 4.65 2.63 1.02 0.70 7.90 1.71 866 903 101 103

1911 Ben_Loyal R 48.93 0.89 9.46 0.28 0.11 16.80 5.46 4.76 2.63 0.98 0.80 7.61 0.99 898 1308 55 153

1921 Haroldswick_Raster1 R resistant feature and sand 51.32 0.99 10.19 0.30 0.13 16.43 6.63 3.54 2.74 0.80 1.16 4.38 1.20 663 505 87 126

1921 Haroldswick_Raster2 R dark sticks and rock 51.52 0.86 9.68 0.32 0.13 17.64 5.92 2.94 2.68 1.00 1.05 4.74 1.22 871 645 105 136

1921 Haroldswick_Raster3 R dark sticks, rock and sand 47.67 0.89 9.47 0.28 0.14 22.22 6.70 3.84 2.55 0.60 1.01 3.49 0.95 640 590 38 159

1921 Haroldswick_Raster4 R resistant feature and rock 51.09 0.79 9.67 0.28 0.08 18.44 6.26 2.63 2.60 0.86 0.96 4.85 1.25 1117 565 76 231

1921 Haroldswick_Raster5 R resistant feature and rock 53.36 0.85 9.88 0.35 0.10 15.87 6.17 2.65 2.57 0.90 1.24 4.39 1.26 946 639 124 159

1922 Raasay R resistant features 50.44 0.90 9.87 0.31 0.11 15.90 5.45 4.39 2.68 0.83 0.73 6.81 1.29 1038 1123 108 145

1925 Assynt_RP R 48.00 0.96 9.41 0.30 0.17 16.52 6.07 5.36 2.51 0.86 0.82 7.60 1.12 935 1069 135 97

1927 Craighead_DRT RB 50.47 1.00 9.86 0.28 0.20 15.71 6.44 3.80 2.54 1.00 0.75 6.47 1.17 968 1159 101 79

1929 Banff R 54.26 1.01 10.64 0.32 0.16 16.43 5.41 3.03 2.80 1.07 0.65 3.12 0.85 904 818 22 103

1931 Unst_DRT RB 51.72 1.00 9.92 0.30 0.11 14.25 4.03 4.54 2.58 0.94 0.89 7.91 1.57 833 665 165 130

©2020 American Geophysical Union. All rights reserved.


1932 Aberfoyle R 52.34 1.09 10.48 0.36 0.21 10.17 6.57 4.21 3.26 0.97 1.07 7.22 1.68 1832 688 217 48

1934 Macleans_Nose R resistant feature 47.64 0.91 9.35 0.25 0.07 19.59 5.41 3.90 2.61 0.85 0.70 6.89 1.52 1293 787 156 280

1934 Ross_of_Mull_DRT RB CaSO4 veins 50.78 0.96 9.65 0.31 0.10 14.12 4.06 4.98 2.46 0.95 0.74 9.02 1.62 925 644 148 141

1937 Rona_raster1 RB CaSO4 vein with dark material 43.22 0.81 8.34 0.29 0.08 12.62 3.98 9.04 2.23 0.71 0.79 16.31 1.38 686 542 142 158

1937 Rona_raster3 RB CaSO4 vein with dark material 35.73 0.65 6.97 0.20 0.10 11.43 4.09 12.91 2.00 0.53 0.73 23.27 1.06 587 503 121 114

1937 Rona_raster4 RB CaSO4 vein with dark material 35.40 0.80 7.43 0.30 0.08 11.02 4.27 13.09 2.18 0.59 0.89 22.61 1.17 471 419 113 138

1937 Rona_raster2 RB CaSO4 vein with dark material 37.03 0.79 7.16 0.28 0.13 12.20 4.39 12.30 2.07 0.56 0.72 21.08 1.08 581 486 135 94

1938 Funzie_APXS R 47.65 0.92 9.20 0.27 0.10 16.90 6.08 4.47 2.67 0.86 0.99 8.13 1.42 1768 695 151 169

1940 Mallaig_offset R 48.34 1.00 9.67 0.35 0.33 17.88 7.17 4.22 2.44 0.69 0.93 5.37 1.17 797 1286 111 54

1940 Mallaig R 48.88 1.00 9.45 0.33 0.35 17.03 7.10 4.29 2.44 0.74 0.90 6.04 1.15 906 1214 118 49

1940 Knoydart R 50.55 1.16 10.25 0.35 0.12 15.76 6.16 4.22 2.56 1.02 0.81 5.26 1.44 1024 1223 87 131

1943 Eaval R 49.11 0.95 9.38 0.29 0.27 17.30 6.49 4.37 2.41 0.75 0.62 6.48 1.29 904 1122 125 64

1945 Loch_Gairloch R 50.46 1.06 9.90 0.34 0.11 17.17 4.91 4.40 2.49 0.83 0.84 5.88 1.33 869 397 140 156

1947 Thurso R 49.51 1.10 9.62 0.33 0.10 18.23 5.98 3.99 2.62 0.72 0.72 5.28 1.55 891 781 158 182

1947 Loch_Tay R 48.72 1.10 9.62 0.33 0.09 19.16 5.81 3.55 2.56 0.79 0.80 5.54 1.67 899 839 78 213

1950 Balmedie_DRT RB CaSO4 vein/concordant/dark inclusions 48.61 1.00 8.92 0.36 0.11 9.13 3.81 7.74 2.04 0.58 1.02 15.44 0.90 341 1039 84 83

1954 Glen_Roy R 48.26 1.00 9.23 0.31 0.10 17.81 5.73 4.41 2.62 0.98 0.73 6.95 1.59 1020 752 136 178

1954 Skara_Brae R 48.74 1.06 9.27 0.30 0.12 17.92 5.55 4.29 2.65 0.99 0.73 6.40 1.71 1088 782 108 149

1959 Arnaboll R 47.59 1.15 9.44 0.30 0.12 18.51 5.59 4.24 2.50 1.05 0.75 6.80 1.71 1004 870 121 154

1963 Newmachar_offset RB fine nodules 46.99 0.84 9.15 0.28 0.12 15.73 4.87 5.67 2.71 0.94 1.04 9.94 1.40 829 1088 116 131

1963 Newmachar_DRT RB fine nodules 50.20 0.93 9.52 0.30 0.12 15.03 4.63 4.96 2.72 1.00 0.76 7.99 1.54 979 1248 104 125

1966 Forties R resistant features 49.34 0.88 9.64 0.26 0.11 16.55 5.24 4.47 2.71 0.96 0.77 7.57 1.24 896 911 80 150

1966 Lake_Orcadie_DRT RB fine nodules 50.53 0.90 9.66 0.29 0.14 14.26 5.26 4.67 2.63 1.00 0.73 8.26 1.33 1040 1152 127 102

1972 Lake_Orcadie_Offset_2 RB fine nodules 52.07 0.95 10.12 0.29 0.15 13.88 5.23 4.53 2.66 0.90 0.74 6.96 1.23 904 1110 114 93

1972 Lake_Orcadie_Offset_1 RB fine nodules 47.41 0.87 9.22 0.29 0.16 14.36 5.99 5.74 2.57 0.94 0.87 9.93 1.28 1024 1144 122 90

©2020 American Geophysical Union. All rights reserved.


1975 Benbecula R CaSO4 vein/sand/rock 34.81 0.74 7.00 0.38 0.22 14.70 5.78 12.63 2.11 0.41 0.62 19.70 0.72 482 450 65 67

1975 Rockall R 49.66 0.99 9.52 0.30 0.14 13.75 5.92 5.53 2.62 0.95 0.79 8.35 1.18 1087 1012 118 98

1978 Lake_Orcadie_Drill_Tailings DT failed drill attempt 45.06 0.90 8.81 0.29 0.13 17.90 4.07 7.14 2.71 0.91 0.66 10.03 1.04 1043 1162 129 138

1979 Lake_Orcadie2_Drill MD failed drill attempt 50.02 1.04 9.83 0.32 0.18 15.45 6.41 4.60 2.74 0.94 0.78 6.16 1.25 984 1002 91 86

1980 Lake_Orcadie2_DRT RB 50.30 0.94 9.51 0.31 0.14 14.93 5.16 4.94 2.66 0.91 0.71 7.83 1.30 1075 1002 102 107

1984 Lake_Orcadie2_post_Drill MD failed drill attempt 50.59 1.00 9.97 0.31 0.14 15.80 4.78 4.79 2.90 1.04 0.57 6.62 1.15 1011 988 91 113

1988 Barkeval R concordant CaSO4 46.28 0.86 9.08 0.27 0.11 16.51 6.33 4.75 2.82 1.20 0.86 8.77 1.71 746 1342 38 150

1988 North_Harris R 48.03 0.95 9.45 0.29 0.11 17.01 6.10 4.34 2.73 0.99 0.83 7.06 1.74 832 1602 23 155

1991 Seaforth_Head R 50.45 1.14 9.54 0.29 0.28 17.18 5.00 4.76 2.38 0.99 0.52 6.01 1.09 879 1037 37 61

1993 Stirling_Castle R 50.17 1.08 10.03 0.28 0.21 17.11 5.14 4.30 2.49 1.03 0.90 5.84 1.16 640 1145 62 81

1995 Paisley R 48.86 0.99 9.32 0.27 0.29 17.15 5.46 4.65 2.70 1.02 0.77 7.02 1.24 817 981 22 59

1995 Durness RB 48.57 0.96 9.23 0.27 0.28 17.02 5.52 4.53 2.80 1.01 0.81 7.01 1.72 704 845 36 61

2160 Tayvallich R rough texture 47.48 1.00 9.01 0.33 0.21 19.66 5.81 3.94 2.61 0.91 0.94 6.54 1.28 911 969 52 94

2165 Trollochy R 48.30 1.00 9.08 0.29 0.23 15.37 6.23 4.40 2.59 1.09 0.63 8.80 1.69 988 1199 57 67

2165 Burn_O_Vat R 48.63 0.99 9.39 0.27 0.22 16.04 6.63 3.84 2.68 1.07 0.77 7.87 1.35 941 935 61 73

2165 Portobello R nodules 51.46 0.97 10.13 0.29 0.17 17.28 5.26 4.02 2.68 1.05 0.68 4.99 0.78 961 699 32 102

2168 Inverness_offset RB nodules 46.67 0.95 9.12 0.31 0.21 13.78 5.68 6.43 2.57 0.95 0.85 11.02 1.16 942 994 178 66

2168 Inverness_DRT RB nodules 47.12 0.97 9.44 0.34 0.22 15.44 6.14 5.98 2.61 0.91 0.81 8.78 0.96 931 969 122 70

2217 Grange R CaSO4 patch with dark inclusions/rock 32.24 0.66 6.58 0.20 0.14 18.78 4.76 11.21 2.26 0.58 0.70 20.42 1.23 720 810 172 134

2217 Inverness_minidrill MD failed drill attempt 48.42 0.94 9.53 0.30 0.22 14.41 5.65 5.92 2.55 1.04 0.72 9.17 0.87 894 924 113 66

2220 Findon R 50.20 0.84 9.83 0.33 0.23 13.39 5.49 5.47 2.39 1.04 0.68 8.57 0.94 965 1289 106 58

2220 Calgary_DRT RB small nodule 49.79 0.94 9.70 0.29 0.23 13.70 5.23 5.11 2.50 1.05 0.72 9.35 1.04 1074 1596 79 60

2223 Highfield_triage R 53.25 0.75 10.59 0.30 0.20 16.20 5.74 3.44 2.75 1.00 0.67 3.30 0.99 852 649 106 81

2223 Highfield_offset RB 48.87 0.88 9.28 0.28 0.13 15.86 4.33 5.53 2.40 0.80 0.79 9.56 1.00 754 603 79 122

2223 Highfield_DRT RB 53.54 0.96 10.30 0.33 0.16 16.13 5.03 3.54 2.68 0.95 0.88 4.18 1.07 1055 731 96 101

©2020 American Geophysical Union. All rights reserved.


2245 Highfield_dump_offset DBA 50.59 0.94 9.45 0.30 0.14 18.60 4.20 4.83 2.58 0.85 0.73 5.61 0.93 1020 700 67 133

2245 Highfield_dump_centre DBA 51.58 0.88 9.64 0.28 0.15 17.31 4.29 4.84 2.57 0.89 0.79 5.63 0.92 930 661 60 115

2247 Highfield_Drill_Tailings DT 52.38 0.96 10.02 0.29 0.14 16.97 4.58 4.19 2.68 0.92 0.78 4.86 0.99 963 707 92 121

2254 Springside R 50.32 0.97 9.62 0.28 0.10 17.43 4.85 4.49 2.45 1.06 0.76 6.31 1.10 854 885 22 174

2255 Woodhill_DRT RB 47.29 0.93 9.33 0.29 0.10 17.76 5.56 4.71 2.62 0.98 0.89 7.79 1.45 803 871 42 178

2258 Corrieshalloch_Gorge R brecciated? 46.44 0.99 8.79 0.28 0.07 17.07 4.83 4.97 2.54 0.68 0.86 10.70 1.54 838 726 156 244

2258 Rock_Hall R brecciated? 42.84 0.98 8.60 0.30 0.12 15.81 6.40 6.89 2.67 0.55 0.95 11.93 1.74 737 705 138 132

2288 Rock_Hall_dump_offset DBA brecciated? 37.65 0.99 7.54 0.33 0.10 20.28 3.92 8.19 2.07 0.64 0.87 14.96 2.06 1025 1006 232 203

2291 Rock_Hall_dump_corrected DBA brecciated? 37.37 1.00 7.33 0.31 0.08 20.00 3.97 8.64 2.13 0.62 0.87 15.28 2.10 995 919 229 250

2292 Rock_Hall_Drill_Tailings DT brecciated? 35.43 0.94 7.21 0.30 0.10 19.77 4.14 9.16 2.03 0.57 0.94 16.85 2.28 877 877 229 198

2295 Bothwell_DRT RB brecciated? 40.69 0.94 8.01 0.28 0.09 15.92 5.50 7.60 2.54 0.51 0.98 14.90 1.80 754 700 179 177

2299 Melrose R descent to clay-bearing unit 51.13 1.24 9.84 0.37 0.06 19.84 4.34 3.34 2.63 0.83 0.76 4.35 1.05 812 709 57 331

2300 Linlithgow R descent to clay-bearing unit 48.66 1.16 9.34 0.30 0.06 19.88 5.32 3.24 2.55 0.98 0.82 5.76 1.59 1000 1010 155 331

2301 Puddledub R descent to clay-bearing unit 46.76 1.12 9.47 0.32 0.09 19.20 5.30 4.43 2.39 1.01 0.86 7.24 1.46 895 1596 235 213

2301 Loch_Ness_DRT RB descent to clay-bearing unit 48.34 1.16 9.73 0.34 0.09 19.98 4.57 3.55 2.53 1.14 0.75 5.41 1.99 907 1825 390 222

See Table S1 for errors

Grey highlighted analyses indicate APXS targets from within blue/grey bedrock areas identified from orbital, colour stretched HiRise imagery

Italics indicate APXS targets associated with successful VRR drill holes

1
R - as is, unbrushed rock surface

RB - brushed rock surface

DT - drill tailings (fines deposited around the hole during drilling; typically sampled from <2 cm depth

MD - shallow test or failed drill; APXS deployed over drill hole

DBA - drill bit assembly (DBA) dump pile; dumped from the DBA using feed extended drilling and feed extended sample transfer techniques

©2020 American Geophysical Union. All rights reserved.

You might also like