Salty atters

10
Page 1 Paleoproterozoic era sediments (2.5-1.6 Ga) contain isotopic evidence the first significant oxygenation event in the world’s atmosphere, largely driven by the increasing dominance of cy- anobacterial photosynthesis. Neoproterozoic sediments (1.0 – 0.542Ga) contain evidence of the second oxygenation event, which is associated with the evolution of widespread multi- cellular life and CaCO 3 /siliceous carapaces. By the end of the Neoproterozoic, the world oceans had chemistries, temperatures and salinities similar to those of the Phanerozoic (Blamey et al., 2016). e intervening Mesoproterozoic (1.6- 1.0 Ga) retains evaporitic residues with aspects of both the late Archean and the Phanerozoic. e oxygenation of Earth’s atmosphere-ocean system occurred in two steps: 1) the Paleoproterozoic “Great Oxygenation Event” Sulphate and oxygenation levels across the Proterozoic Proterozoic (2.5 – 0.542 Ga) saline sediments encompass signifi- cant transitions in evaporite style and chemistry within an evolv- ing atmospheric and oceanic framework. Lithospheric changes tie to a cooling and biologically-evolving earth as earth-scale plate tectonics move to a system set comparable with that op- erating today. e Proterozoic eon is divided into three eras: the Paleoproterozoic, Mesoproterozoic and Neoproterozoic. ick sequences of halite are only found as the actual bedded salts in sediments of the Neoproterozoic (and the Phanerozoic), while calcium sulphate residues and beds occur in all three Proterozoic eras, especially in parts of the Paleoproterozoic and the Neopro- terozoic. Silica mobility and replaced evaporites: 4 - Proterozoic atmospheric transitions and saline microporous chert reservoirs www.saltworkconsultants.com John Warren -Friday, October 7, 2016 Palaeoproterozoic Palaeoproterozoic Neoproterozoic Neoproterozoic Gyr ago 2.5 0 2 4 6 8 10 10 5 0 Sulphate concentration (mM) 1.6 1.0 0.54 “cap carbonate” Initial biospheric oxgenation Lomagundi Event Mesoproterozoic Mesoproterozoic δ 13 C (‰ VPDB) Tulomozero Fm (Reuschel et al., 2012) Figure 1. Proterozoic marine sulphate concentrations. Filled ovals mark data determinations (in part after Kah et al., 2004). Marine sulphate concentrations probably remained mostly below 2mM until about 1.3 Gyr ago (except in the Lomagundi event), rose to, 4.5mM by 1.2 Gyr ago, and possibly as high as 7–10mM by the mid-Neoproterozoic, although near-complete reduction of oceanic sulphate may have occurred during Neoproterozoic glacial intervals (arrow). Increased sulphate concentrations coincide with changes in marine δ 13 C associated with biospheric oxygenation and are marked by the appear- ance of bedded gypsum (stars) in the late Mesoproterozoic and more frequent gypsum deposition in the last two-thirds of the Neoproterozoic. See also Figure aa).

Transcript of Salty atters

Page 1: Salty atters

Page 1

Paleoproterozoic era sediments (2.5-1.6 Ga) contain isotopic evidence the first significant oxygenation event in the world’s atmosphere, largely driven by the increasing dominance of cy-anobacterial photosynthesis. Neoproterozoic sediments (1.0 – 0.542Ga) contain evidence of the second oxygenation event, which is associated with the evolution of widespread multi-cellular life and CaCO3/siliceous carapaces. By the end of the Neoproterozoic, the world oceans had chemistries, temperatures and salinities similar to those of the Phanerozoic (Blamey et al., 2016). The intervening Mesoproterozoic (1.6- 1.0 Ga) retains evaporitic residues with aspects of both the late Archean and the Phanerozoic.

The oxygenation of Earth’s atmosphere-ocean system occurred in two steps: 1) the Paleoproterozoic “Great Oxygenation Event”

Sulphate and oxygenation levels across the Proterozoic Proterozoic (2.5 – 0.542 Ga) saline sediments encompass signifi-cant transitions in evaporite style and chemistry within an evolv-ing atmospheric and oceanic framework. Lithospheric changes tie to a cooling and biologically-evolving earth as earth-scale plate tectonics move to a system set comparable with that op-erating today. The Proterozoic eon is divided into three eras: the Paleoproterozoic, Mesoproterozoic and Neoproterozoic. Thick sequences of halite are only found as the actual bedded salts in sediments of the Neoproterozoic (and the Phanerozoic), while calcium sulphate residues and beds occur in all three Proterozoic eras, especially in parts of the Paleoproterozoic and the Neopro-terozoic.

Salty MattersSilica mobility and replaced evaporites: 4 - Proterozoic atmospheric transitions and saline microporous chert reservoirs

www.saltworkconsultants.com

John Warren -Friday, October 7, 2016

Palaeoproterozoic

Palaeoproterozoic

Neoproterozoic

Neoproterozoic

Gyr ago 2.5

0

2

46

8

10

10

5

0

Sulp

hate

con

cent

ratio

n(m

M)

1.6 1.0 0.54

“cap carbonate”

Initial biosphericoxgenation

LomagundiEvent

Mesoproterozoic

Mesoproterozoic

δ13C

(‰ V

PDB)

Tulomozero Fm (Reuschel et al., 2012)

Figure 1. Proterozoic marine sulphate concentrations. Filled ovals mark data determinations (in part after Kah et al., 2004). Marine sulphate concentrations probably remained mostly below 2mM until about 1.3 Gyr ago (except in the Lomagundi event), rose to, 4.5mM by 1.2 Gyr ago, and possibly as high as 7–10mM by the mid-Neoproterozoic, although near-complete reduction of oceanic sulphate may have occurred during Neoproterozoic glacial intervals (arrow). Increased sulphate concentrations coincide with changes in marine δ13C associated with biospheric oxygenation and are marked by the appear-ance of bedded gypsum (stars) in the late Mesoproterozoic and more frequent gypsum deposition in the last two-thirds of the Neoproterozoic. See also Figure aa).

Page 2: Salty atters

Page 2

(GOE ≈ 2.3 Ga), which refers to the transition from a perva-sively reducing Earth-surface system to one with an oxygenated atmosphere and oxygenated shallow seas, and 2) the “Neopro-terozoic Oxygenation Event” (NOE), when the Earth’s atmo-sphere and ocean are understood to have become persistently oxygenated down to the deep ocean bottom (Turner and Bekker, 2016; Scott et al., 2014). The GOE is indicated by a Proterozo-ic carbon isotope anomaly known as the “Lomagundi event,” a positive carbon isotope excursion between ca. 2.22 and 2.06 Ga, interpreted to be the result of high organic carbon burial and at-tendant accumulation of atmospheric oxygen (Figure 1; Bekker and Holland, 2012)). A long interval spanning the remainder of the Paleoproterozoic and much of the Mesoproterozoic followed the Lomagundi event, typified by lower levels of atmospheric oxygen and little variation in carbon isotope values. This ended in the late Neoproterozoic with dramatic fluctuations, of escalating magnitude, in the biogeochemical carbon cycle and attendant fluctuations within an overall increasing oxygen content (Figure 2). By the end of the Neoproterozoic, not just shallow shelf wa-ters but much of the deep-ocean water column was consistently oxygenated (see article 1 in this series)

Chert evolutionWe have already seen how low levels of oxygen and high levels of CO2 in the Archean favoured the precipitation of nahcolite, and its hydrothermal silica association, atop subsealevel isolated saline sumps in microcontinents and island arcs in a saline wa-terworld (article, 28 August 2016). The hydrothermally-dominat-ed silica of the silica-rich Archean oceans is reflected in more negative d30Si isotope values in widespread marine cherts of that time, compared with most Proterozoic cherts (Chakrabarti et al., 2012; see also Figure 3b). However, shorter-term fluctuat-ing levels of atmospheric oxygen in the Proterozoic also influ-enced drop-out salinities for gypsum in Mesoproterozoic ma-rine brines. In brines derived from the modern, well-oxygenated world oceans, as in evaporite successions deposited throughout the Phanerozoic, the sulphate minerals gypsum and anhydrite precipitated from evaporating seawater after aragonite or calcite, but before halite (see 26 August, 2015, blog for more detail). At lower seawater sulphate levels across the Archean and much of the Proterozoic, gypsum and anhydrite precipitated after halite, even at Na and Cl concentrations similar to those of the modern ocean.

This is why some post-Lomagundi, Paleoproterozoic marine evaporite successions show clear evidence of halite precipitation before gypsum or anhydrite or even an absence of gypsum or anhydrite with halite (e.g., ≈ 1.88 Ga Stark Formation; Pope and Grotzinger, 2003). The post-halite precipitation of calcium sul-phate is construed as evidence for a limited marine sulphate res-ervoir and little atmospheric oxygen (Scott et al., 2014). In con-trast, Lomagundi- age sedimentary successions contain evidence for sulphate precipitation before halite (Melezhik et al., 2005; Bekker et al., 2006; Schröder et al., 2008). Sulphur isotope values of marine sulphates (in CaSO4, barite) and sulphides in marine pyrite also record expansion and contraction of anoxic oceanic settings. That is, a higher burial rate of pyrite in anoxic settings is indicated by a positive shift in the sulphur isotope values of

sulphates, whereas ocean oxygenation creates a negative shift in values (e.g., Claypool et al., 1980; Strauss, 1997). Furthermore, expansion of the area of anoxic oceanic settings decreases the size of the seawater sulphate reservoir, resulting in more variable sulphur isotope values of sulphate evaporites, barites, and other carbonate-associated sulphates (Figure 2; Kah et al., 2004). This applies in particular in the Mesoproterozoic when only tshallow oceanic waters were consistently oxygenated.

When we look at silica mobility and chert styles across a Pro-terozoic milieu of evolving oxygen and sulphate levels we see some aspects similar to the Phanerozoic and others more akin to the high-silica oceans of the Archean. Maliva et al. (2005) and Perry and 2014 show that the latter part of the Paleoproterozoic era (post-Lomagundi) is marked by the end of widespread pri-mary and early diagenetic silica precipitation in normal marine subtidal environments. However, silica precipitation continued apace in the deeper marine in waters that were still anoxic. The Paleoproterozoic is defined by the “rusting” of the shallower parts (shelves and upper slopes) of the world’s ocea,n as dissolved oxy-gen levels increased and the accumulation of widespread Banded Iron Formations (BIFs) occurred, including the huge deposits of NW Australia.

Where and when do we see nodular cauli-flower chert after sulphate in the Protero-zoic?Some of the oldest silicified nodular sulphates with cauliflower chert textures and actual relict anhydrite are found in the Hu-ronian Gordon Lake Formation (≈2.4 Ga; Chandler, 1988). The nodules are commonest in mud chip breccia at the base of sand-stone, siltstone and mudstone upward-fining storm cycles. An-hydrite nodule relicts are composed of a mosaic and meshworks of blocky crystal laths. Earlier laths are preserved in silicified outer rims of many cauliflower chert nodules, with texture align-ments similar to that of Recent displacive sabkha anhydrite nod-ules. Completely silicified nodules are composed of megaquartz, some calcite-cored, or of jasper, with replacement textures iden-tical those documented the Phanerozoic by Milliken (1979). Similarly-textured cauliflower cherts are found in the Mallapu-nyah Formation (1650 Ma) in the Paleoproertozoic sediments of the McArthur Basin in the Northern Territory of Australia (Warren, 1999). Thin sections through 5-30cm diameter Mal-lapunyah nodules (in a redbed host) can still retain small relict highly-birefringent laths of anhydrite, but most of the former felted cores to the nodules are now composed of mimetic silica. There are also older sedimentary chert nodules in the McArthur Basin succession located in units below the level of the Mallapu-nyah, but they are smooth walled not rugose-surfaced features.

Unfortunately, the term cauliflower chert is loosely defined and is used to describe cemented features in Meosproterozoic and earlier sediments and metasediments, which are not true calcium sulphate evaporite replacements.Although termed cauliflower features, they do not have surface textures resembling the flo-rets of a cauliflower (see article 2 in this series - July 31, 2016). For example, aggregates and clusters of growth-aligned barite

www.saltworkconsultants.com

Page 3: Salty atters

Page 3

crystals in the Archean of South Africa are described as cauli-flowers when they should be described as bladed, palmate crystal aggregates (Reimers and Heinrich, 1997). Interstingly, Chowns and Elkins (1974) in a study of cauliflower cherts occurrences across the USA list no examples older than Cambrian. Using a

tighter definition of cauliflower chert and recognising that this term should not be interchangeable with crocodile-skin chert it seems that Proterozoic occurrences of cauliflower chert nodules largely mirror times when oxygen levels were sufficiently high in the world’s ocean to allow sulphate in solution. In the Pa-

www.saltworkconsultants.com

5

35 45

Yudomski event

gypsum-first marine evaporite (Kilian Fm., SSG)

youngest halite-first marine evaporite (Dodo Creek Fm., MMSG)

900

800

700

600

Tonia

nCr

yoge

nian

Trezona anomaly

Ediac

aran

-10 -5 0

Camb

rian

Gaskiers glaciation

Marinoan glaciation

Sturtian glaciation

34δ S ‰ VCDT marine CAS & evaporite

15

500

2510010

[O ] %PAL213δ C ‰ VPDB marine carbonate

Shuram anomaly

Ma(Gorjan et al., 2000; Halverson et al., 2010)(Canfield, 2005;

Shields-Zhou & Och, 2011)

(pink = Halverson, 2006)(red = Halverson et al., 2007, 2010;Macdonald et al., 2010; Saltzman & Thomas, 2012)

oldest thick marine sulphate evaporite successions;

oldest gypsum-first Neoproterozoic evaporite successions

Ten Stone Fm., MMSG Minto Inlet Fm., SSG)

5

4

3

2

1

Islay anomaly

Rasthof anomaly

Maieberg anomaly

Tayshiranomaly

Snail Spring Fm.Ten Stone Fm.

Gayna Fm

5

oldest uncontested metazoan body fossils (Hofmann et al., 1990)

oldest crown-group metazoan biomarkers (Brocks et al., 2014)

oldest petrographically discernable metazoan-grade tissue (Neuweiler et al., 2009)

Bitter Springs anomaly

Halverson, 2006

Araanomaly

Ram Head Fm.

vase-shaped microfossils(Strauss et al., 2014)

CBA

Turner andBekker, 2016

Figure 2. Neoproterozoic geochemical events (after Turner and Bekker, 2016 and references therein; MMSG—Mackenzie Mountains Supergroup; SSG—Shaler Supergroup). (A) General understanding of late Neoproterozoic atmospheric oxygen concentration. Detailed trajectory of Neoproterozoic atmospheric oxygen concentration has yet to be determined, including identification of the inflection point at which an accelerated rise in global Neoproterozoic oxygen concentrations started - see also Figure 1. %PAL—percentage of present atmospheric level. (B) Positions of Gayna, Ten Stone, and Snail Spring Formations relative to established δ13C curves. A composite curve of δ13C values obtained by Turner and Bekker (2016; blue line) fills part of a gap in the pre-existing MMSG data, and delineates new, small-scale pre–Bitter Springs carbon isotope anomaly (BSA) excursions that might be of global significance. VPDB—Vienna Peedee belemnite. (C) Green curve indicates composite δ34S curve for the Ten Stone Formation from Turner and Bekker, 2016). Pale green dots for existing δ34S sulfate (evaporite and carbonate-associated sulfate) data are from compilation of Halverson et al. (2010), but data for sulfate deposited after the BSA (lower Loves Creek Member of the Bitter Springs Formation) have been adjusted to a younger depositional age (younger than ca. 811.5 Ma; Macdonald et al., 2010) than was available when the data were originally published (Gorjan et al., 2000). Numbered intervals in the sulfur isotope field refer to stages in deposition of mid- to late Neoproterozoic bedded sulfate evaporites; horizontal arrows indicate the range of isotopic values for each stage. VCDT—Vienna Canyon Diablo Troilite. Numeral 1 theough 5 refer to: (1) immediately preceding the BSA (Bitter Springs Anomaly); (2) immediately following the BSA - Kilian Formation, Shaler Supergroup; Redstone River Formation, Coates Lake Group [Northwest Territories, Canada]; Upper Roan and Mwashia Subgroups, Central African copper belt; Loves Creek Member, Bitter Springs Formation, Amadeus Basin, Australia; (3) post-Sturtian glaciation (Tapley Hill Formation, Umberatana Group, Adelaide Basin (Australia); (4) post-Marinoan glaciation(Schisto-calcaire Cycle, West Congolian Supergroup, Congo and Gabon; Hüttenberg Formation, Tsumeb Subgroup, Namibia and (5) post-Shuram excursion to Early Cambrian (Hormuz Group, Iran; Ara Formation, Huqf Group, Oman; Hanseran Evaporite Group, northwestern India.

Page 4: Salty atters

Page 4

www.saltworkconsultants.com

leoproterozoic and Mesoproterozoic only the upper parts of the ocean column, including waters covering the world’s continental shelves (and derived evaporite basins) were sufficiently oxygen-ated to allow the formation of cauliflower chert after nodular anhydrite. However in some Neoproterozoic basins, especially if located in sumps in a highly-restricted brine layered seafloor, the levels of anoxia in the ponded bottom brines facilitated the accumulation of laminar microporous chert in association with evaporites or their early replacements

Primary laminated hypersaline silica chert in an evaporite basin at the Precambri-an-Cambrian boundaryAn organic-rich laminated porous chert known as the Athel or Al Shamou silicilyte consists of up to 90% microcrystalline quartz along with dolomite, magnesite, anhydrite and halite (Rajaibi et al., 2015). It occurs at the Precambrian-Cambrian boundary in the subsurface of the South Oman Salt Basin, Sul-tanate of Oman, where it acts as a light-oil reservoir (Ramsey-er et al., 2013; Amthor et al., 2005). Fully encased in variably halokinetic salt masses, it was first discovered during the 1990’s hydrocarbon exploration activities of Petroleum Development Oman. This laminated microporous and variably fractured chert,

has its source of silica and its mode of precipitation tied to an anoxic, sulphur-rich, stagnant and highly saline basin. Its homo-geneous silica distribution and high Si isotope values (avg. d30Si = +0.83 ± 0.28), coupled with a low molar Ge/Si ratio (<0.25 x 10-6) in its microcrystalline quartz matrix imply dissolved silica in concentrated seawater as the Si source, and hydrothermal or biogenic (e.g. sponge-derived) silica are excluded.

Silica precipitation from a seawater-sourced brine was likely the result of a dramatic increase in salinity in response to halokinetic salt dissolution atop and adjacent to the edges of transtensional depressions on a deep basin floor in the South Oman Salt ba-sin, thus markedly reducing the solubility of amorphous silica in these brine-filled seawater depressions. This saturation triggered the formation of silica gel. The gel accumulated at the base of a brine-layer covered basin floor, forming a soft silica-rich layer bound into bacterial mats, giving rise to its fine-scale lamination. The mean number of laminae in this laminated chert is ca. 32 per year suggesting that layering is non-annual and controlled by processes such as fluctuations in nutrient supply, lunar driv-en re-mixing or diagenetic segregation. The transformation of the silica-gel to microcrystalline quartz occurred below 45°C indicating a less than -4.5‰ d18O composition of the pore-wa-ter during microcrystalline quartz formation. The microporous hydrocarbon filled nature of this ancient chert and the fact the

A. C.

B.

Al Fajr-1

Al Kanz-1

Waad NorthBasin

Waad BasinAl N

oor Basin

Makhzoun Basin

Amin Basin

Qaarharail B

asin

Makhzoun No. 1

MarmulNW No.7

Salt edge

Amin Field

Athel DiscoveryWater-bearing AthelAthel Silicilyte mini-basinsCarbonate platformClastics Fault

Washaq BasinAl Noor

Al Shomou

Mazraq South-1

SouthernPlatform

NorthernPlatform

Athel well penetrationAthel wells that didnot penetrate silicilyte

0 10

km

X Y

North

A4

Subsidence

Subsidence

ReservoirsHalite/Evaporites

Shale/Sandstone

Platform CarbonatesBasinal CarbonatesSilicilyte (Basinal)

DeepBasinSouthern Carbonate Domain

EasternFlank

Isolated Buildups

XY

δ30Si

(‰)

2

1

0

1

0 5 10 15Al2O3 (wt%)

Al Shomou (Athel) silicilyteArchean chert

Volcanic-association

marine-derived

Hydrothermal-association

δ30Si (‰)0-1-2-3-4 1 2 3 4

Al Shomou silicilyteSponges

Modern seawater

Hydrothermal �uids

Hydrothermal depositsClays

Sicretes and sandstones

Igneous rocks

RiversRivers

Figure 3. Al Shomou (Athel) silicilyte. A) Athel play map generated after drilling a number of unsuccessful penetrations illustrating the control of structural elements on the distribution of the Athel silicilyte in the South Oman Salt Basin (after Amthor et al., 2005) The oc-currence of silicilyte in this model is restricted to lows between underlying Buah topographic highs which in�uence subsequent salt-cored sea�oor topography (see Warren 2016; Chapter 10, for detail on hydrocarbon association in the silicilyte ). B) δ30Si composi-tion of main modern Si reservoirs. C) δ30Si composition of Early Cambrian Athel Al Shamou) silicilyte samples and Al2O3 content in com-parison with Archaean chert samples with hydrothermal and volcanic associations. Data for both plots B and C derived from Ramseyer et al. (2009, 2013) and references therein.

Page 5: Salty atters

Page 5

hydrocarbon-filled micropores are still distinct after more than 500 million years after they filled (Figure 4d, e) is why when artificially fractured the silicilyte can act as a hydrocarbon reser-voir (See Rajaibi et al. 2015 and Warren, 2016; Chapter 10 for a summary of relevant literature).

What does this mean for other evaporite-associated laminar cherts in the Pha-nerozoic?Beds of laminated chert (not replaced, but precipitated, primary silica accumulations) are unusual across the Phanerozoic, but cauliflower cherts as replacement of calci-um sulphate evaporites are not (Chowns and Elkins, 1974). Spanning the Precam-brian boundary, the Athel Silicilyte is the first Phanerozoic example of a laminar chert-hypersaline association. Then, there is the somewhat younger but world-famous Devonian to early Carboniferous Caballos Novaculite. This chert location was the site of a world-famous set of arguments as its origin, between two world-renown pro-fessors, Dr Earle McBribe and Dr Robert Folk, both on the faculty of the Universi-ty of Texas at that time. Folk argued for a shallow-water peritidal hypersaline deposi-tional setting, McBride for a deep a marine setting but still with possible hypersaline indicator textures (Folk, 1973; Folk and Mcbride, 1976; McBride and Folk 1977).

The Caballos Novaculite outcrops in the Marathon Uplift of Texas, while its lith-ologic and time equivalent, the Arkan-sas Novaculite, outcrops in the Ouachita Mountains of Arkansas and Oklahoma (Figure 5). Novaculite chert) in outcrop is very resistant to erosion so that layers of novaculite stand out as characteristic ridges and dip slopes in the Ouachita and Mar-athon mountains (Figure 5). This outcrop forms and its hard abrasive nature gives it the name novaculite, which in its Latin root novacula, means razor-stone. When some novaculite is fractured in the subsur-face, there is sufficient connected porosi-ty to form a fractured reservoir play, as in Arkansas and Texas. There, some 30 years ago, oil and gas fields such as Isom Springs in Oklahoma and McKay Creek, Pinion and Thistle fields in West Texas were dis-covered in the Caballos and Arkansas no-

vaculite-chert. The chert reservoir is most productive when it is highly fractured, occurs within complex thrust faults and has had any enclosed carbonate material leached from its chert matrix, so creating microporosity (Figure 6; Godo et al., 2011).

Chert beds in the Caballos Novaculite are composed of equant grains of microcrystalline quartz, minor amounts of illite and radiolaria, and trace amounts of pyrite, carbonate and other min-erals and organic matter. The chert beds are generally interpreted

www.saltworkconsultants.com

Figure 4. (A, B): Core photographs of Athel silicilyte illustrating the finely laminated nature of the silicilyte, in places displaying compacted early fractures (B). Scale bar is 4 cm (C): Plane-light thin-section photomicrograph showing a detrital quartz grain in the well-laminated silicilyte (well W2). Note the laminae below the grain are deformed due to the impact (“dropstone”) and following compaction, whilst the laminae covering the grain show onlap and wedge-shaped structures typical for embedding on an uneven surface. Scale bar is 200 mm. (D): SEM image illustrating the texture of a highly porous silicilyte and the sub-microcrystalline nature of quartz, well W1. Scale bar is 10 mm. (E): Close-up of (D) illustrating the euhedral habit of the quartz crystals and the nature of the intercrystalline porosity. Scale bar is 1 mm. Figure is after Amthor et al. (2005).

Page 6: Salty atters

Page 6

as having formed by the silicification and alteration of precur-sor sediment, sometimes massive, other times finely laminated. Some beds retain occasional evidence biogenic silica derived from radiolaria, while underlying levels retain can contain abun-dant siliceous sponge spicules. Fractures and crackle breccias developed in a grey chert following lithification; green siliceous sediment, whose lithification was impeded by clay, filled these pre-orogenic fractures.

Beds of red shale, chert pebble and cobble conglomerate, sand-stone, limestone, dolomite, and lumpy manganiferous and jas-peritic chert make up no more than 3% of the chert and shale members of the Caballos, but still are of controversial origin and environmental significance. The chert conglomerate beds, for ex-ample, are interpreted as tidal-channel deposits by Folk and as mass-flow deposits by McBride. Jasper beds texture are consid-ered bizarre by many geologists who have worked on them: they are lumpy, uneven beds 0.2 to 2 m thick composed of cherry-red chert with local geopetal cavities, contorted laminae. manganif-erous zones, cauliflowerlike quartz-filled nodular cavities some-times with hollow centres and variably filled or partially filled with zebraic chalcedony, lutecite, quartzine, pseudocubic quartz crystals, and filamentous structures resembling algae. These beds are interpreted by Folk as the product of diagenetic alteration of sabkha evaporite nodules and siliceous ooze, during and follow-ing subaerial exposure with soil development, and by McBride as the product of diagenetic alteration of evaporite beds deposited in deep water and sandwiched between radiolarian ooze. Syn-thesis of evidence on the origin of both the novaculite and chert and shale members leads to contrasting interpretations of wa-ter depth during deposition. However, if the evaporite solution breccias, recognised as such by both authors and the cauliflower cherts as replaced diagenetic anhydrite clusters then the depo-sitional setting is akin to that of the Athel Silicilyte (namely a deepwater holokinetic hypersaline evaporite setting not unlike a siliceous DHAL association (see Warren, 2016; Chapter 9, for discussion of the DHAL literature).

Economic implications of defining silicilyte versus novaculite versus tripolite versus diatomaceous oozeWe have already seen that microporous cherts when fractured are possible reservoir rocks as illustrated by the saline-associated Athel Silicilyte and the Caballos Novaculite. The former retains its microporosity because of an early hydrocarbon charge into existing microporosity (Rajaibi et al. 2013; Amthor et al., 2005), while the reservoir quality of the latter was enhanced by diage-netic leaching of finely dispersed carbonate material, likely when it was caught up in the Ouachita Orogeny (Figure 6; Godo et al., 2011). Like the Athel, the Arkansas (Caballos) Novaculite is thought to be self-sourcing in terms of reservoir hydrocarbons (Zemmels et al., 1985).

The term silicilyte is defined by Rajaibi et al. (2013) as a “local-ly-used” term to describe porous organic-rich laminated chert, it is a succession of microcrystalline quartz that is preserved within salt-encased slabs, 300 to 400 m thick, at a depth of 4 to 5 km in the South Oman Salt Basin. As mentioned earlier, the term novaculite comes from its outcrop expression and its “razorstone” properties and does not necessarily have a direct connection to microporosity in the reservoir portions of the unit (Figure 5). Outcrops of weathered microporous chert zones in the upper part of the Arkansas (Caballos) Novaculite are called tripoli or rottenstone (Figure 6). When present in this finely powdered microporous form, it is quarried and crushed for use as a polish-ing abrasive in metalsmithing and woodworking. When present as a very hard dense rock, it can be cut and shaped for use as a whetstone or razorstone. Before European settlement, novaculite was a source for numerous arrow-tips, spear tips and knives.

Etymological variations of the terms tripoli, novaculite and silicilyte as forms of chert, as currently used in the geological literature are interesting and geologically confused. This is par-ticularly true if the writer did not 1) understand there are various origins to laminar sometimes microporous cherts, and 2) that

www.saltworkconsultants.com

Figure 5. Outcrop expression of Carballos Novaculite as dip slopes or �atirons, that gives us the term novaculite (its latin root mean-ing razorstone) This outcrop is located on Highway 385 in the Marathon Basin, Texas (image downloaded from <texasgeological-press.com>).

Page 7: Salty atters

Page 7

there various possible silica sources and precipitation/replacement mech-anisms can be halotolerant bacteria, or other silica sources that can be tied to marine sponges and yet others to radiolaria and diatomaceous oozes. Hence, there is the time-related as-pect of biogenic chert evolution tying back to the source of the silica in some cherts and the presence or lack of sa-linity indicators in a laminar chert and chert nodules (e.g., cauliflower versus crocodile-skin cherts).

A diatomaceous ooze is a form of opaline silica made up of accumula-tions of siliceous frustules of diatoms in normal marine pelagic sediments and when it retains microporosity is sometimes called tripolite or tripolitic earth. Diatomaceous oozes, the precursor to this type of tripolitic earth is often laminated, but is harsh to the feel and scratches glass. To add to the confusion there are microporous mesohaline diatomaceous oozes called the Tripoli unit in many Messinian subbasins.

Tripoli or tripolite powder is also the term used to describe the form of microporous laminar chert used as an abrasive and collected from weathered zones of the Caballos and Arkansas Novaculites. The Palaeozoic age of the Caballos and Arkansas Novaculite means it cannot contain diatoms. Diatoms evolved in the Cretaceous and have been the dominant source of remo-bilised silica in marine chert nodules ever since. Palaeozoic silica remobilized into chert is often related to nearby sponge spicule horizons, or less often to radiolarian beds. The further back in time, and perhaps the more saline the marine bottom water, the greater the likelihood of a microbial association with chert pre-cipitation and remobilisation.

Even in Tertiary strata some microporous diatomaceous earths can have a normal-marine, organic-enriched depositional asso-ciation and can constitute a fractured microporous hydrocarbon reservoir. This is the case with the Miocene-age fractured micro-porous chert reservoirs that produce today in the Santa Barbara Basin of offshore California (Reid and McIntyre, 2001). Cores and coastal outcrops of the Monterey Formation show this type of marine-deepwater diatomaceous ooze is interlayered with mi-crobial (bacterial-archeal) methanogenic organic-rich dolomites. Then, there are the deepwater diatomaceous saline oozes in the Miocene units that immediately underly Messinian evaporites in the Lorca and other sub-basins across the Mediterranean (Rouchy et al., 1998). These diatomaceous organic-rich oozes were deposited as pelagic sediments in saline waters on stratified bottoms that herald the creation of saline bottom water layers related to the onset of hypersaline conditions. Their depositional setting is in restricted basins with increasingly saline bottoms, driven by tectonic isolation and drawdown that soon after pre-cipitated the salts of the Messinian Salinity Crisis. Finally, there are the lacustrine diatomaceous oozes accumulating at the base

of density-stratified water columns in lakes of African Rift Val-ley. This type of laminar ooze occurred in the deeper parts of the lake floor and in Lake Magadi and Lake Natron define units that immediately predate significant lake drawdown episodes that are defined by the type 1 (sodium bicarbonate) evaporite layers. These laminar diatomaceous chert beds contain nodules with the characteristic surface shrinkage textures of crocodile-skin chert (see article 1 in this series of four).

So, in terms of silica with an evaporite association, “one size does not fit all.’ There are multiple saline settings, both depositional and diagenetic, with silica sources evolving with life across the Proterozoic into the Phanerozoic. Interactions between biology and brine chemistry control the accumulation of silica in various forms in a range of evaporite settings, ranging across the ma-rine to deep halokinetic seafloors to lacustrine basins. Once we understand how to recognise cauliflower chert and its possible association with laminar saline cherts of the Proterozoic and the Phanerozoic, then particular chert styles can help to define the evolution of atmospheric oxygen and the saline versus non-sa-line origins of some organic-rich laminar biogenic microporous cherts.

www.saltworkconsultants.com

2610.7 m in the Shell #1 Arivett well

- calcite leached from its novaculite matrix(tripolitic earth)

Subsurface Sample still retains unleached calcite

Thin Section from outcrop sample

500 µm

500 µm

Figure 6. Microporous and non porous varieties of the Arkansas Novaculite (after Godo et al., 2011)

Page 8: Salty atters

Page 8

ReferencesAl Rajaibi, I. M., C. Hollis, and J. H. Macquaker, 2015, Ori-gin and variability of a terminal Proterozoic primary silica pre-cipitate, Athel Silicilyte, South Oman Salt Basin, Sultanate of Oman: Sedimentology, v. 62, p. 793-825.

Amthor, J. E., K. Ramseyer, T. Faulkner, and P. Lucas, 2005, Stratigraphy and sedimentology of a chert reservoir at the Pre-cambrian-Cambrian boundary: the Al Shomou Silicilyte, South Oman Salt Basin: Geoarabia, v. 10, p. 89-122.

Bekker, A., and H. D. Holland, 2012, Oxygen overshoot and recovery during the early Paleoproterozoic: Earth and Planetary Science Letters, v. 317-318, p. 295-304.

Bekker, A., J. A. Karhu, and A. J. Kaufman, 2006, Carbon isotope record for the onset of the Lomagundi carbon isotope excursion in the Great Lakes area, North America: Cambrian, v. 148, p. 145-189.

Blamey, N. J. F., U. Brand, J. Parnell, N. Spear, C. Lécuyer, K. Benison, F. Meng, and P. Ni, 2016, Paradigm shift in determin-ing Neoproterozoic atmospheric oxygen: Geology.

Chakrabarti, R., A. H. Knoll, S. B. Jacobsen, and W. W. Fischer, 2012, Si isotope variability in Proterozoic cherts: Geochimica et Cosmochimica Acta, v. 91, p. 187-201.

Chandler, F. W., 1988, Diagenesis of sabkha-related, sulphate nodules in the early Proterozoic Gordon Lake formation, On-tario, Canada: Carbonates and Evaporites, v. 3, p. 75-94.

Chowns, T. M., and J. E. Elkins, 1974, The origin of quartz ge-odes and cauliflower cherts through the silicification of anhy-drite nodules: Journal Sedimentary Petrology, v. 44, p. 885-903.

Claypool, G. E., W. T. Holser, I. R. Kaplan, H. Sakai, and I. Zak, 1980, The age curves of sulfur and oxygen isotopes in marine sulfate and their mutual interpretation: Chemical Geology, v. 28, p. 199-260.

Folk, R. L., 1973, Evidence for Peritidal Deposition of Devonian Caballos Novaculite, Marathon Basin, Texas: Bulletin American Association Petroleum Geologists, v. 57, p. 702-725.

Folk, R. L., and E. F. McBride, 1976, The Caballos Novaculite revisited Part I: ”Origin of novaculite members”: Journal of Sed-imentary Petrology, v. 46, p. 659-669.

Godo, T. J., P. Li, and M. E. Ratchford, 2011, Exploration for the Arkansas Novaculite Reservoir, in the Southern Ouachita Mountains, Arkansas: AAPG Search and Discovery Article #90124 © 2011 AAPG Annual Convention and Exhibition, April 10-13, 2011, Houston, Texas.

Heinrichs, T. K., and T. O. Reimer, 1977, A sedimentary bar-ite deposit from the Archean Fig Tree Group of the Barberton Mountain Land (South Africa): Economic Geology, v. 72, p. 1426-1441.

Kah, L. C., T. W. Lyons, and T. D. Frank, 2004, Low marine sul-phate and protracted oxygenation of the Proterozoic biosphere: Nature, v. 431, p. 834-838.

McBride, E. F., and R. L. Folk, 1977, The Caballos Novaculite revisited; Part II, Chert and shale members and synthesis: Jour-nal of Sedimentary Research, v. 47, p. 1261.

Melezhik, V. A., A. E. Fallick, D. V. Rychanchik, and A. B. Kuznetsov, 2005, Palaeoproterozoic evaporites in Fennoscandia: implications for seawater sulphate, the rise of atmospheric oxy-gen and local amplification of the delta C-13 excursion: Terra Nova, v. 17, p. 141-148.

Milliken, K. L., 1979, The silicified evaporite syndrome; two as-pects of silicification history of former evaporite nodules from southern Kentucky and northern Tennessee: Journal Sedimenta-ry Petrology, v. 49, p. 245-256.

Pope, M. C., and J. P. Grotzinger, 2003, Paleoproterozoic Stark Formation, Athapuscow Basin, Northwest Canada: Record of cratonic-scale salinity crisis: Journal of Sedimentary Research, v. 73, p. 280-295.

Ramseyer, K., J. E. Amthor, A. Matter, T. Pettke, M. Wille, and A. E. Fallick, 2013, Primary silica precipitate at the Precambri-an/Cambrian boundary in the South Oman Salt Basin, Sultan-ate of Oman: Marine and Petroleum Geology, v. 39, p. 187-197.

Reid, S. A., and J. L. McIntyre, 2001, Monterey Formation por-cellanite reservoirs of the Elk Hills field, Kern County, Califor-nia: Bulletin American Association Petroleum Geologists, v. 85, p. 169-189.

Rouchy, J. M., C. Taberner, M. M. Blanc-Valleron, R. Sprovieri, M. Russell, C. Pierre, E. Di Stefano, J. J. Pueyo, A. Caruso, J. Dinares-Turell, E. Gomis-Coll, G. A. Wolff, G. Cespuglio, P. Ditchfield, S. Pestrea, N. Combourieu-Nebout, C. Santisteban, and J. O. Grimalt, 1998, Sedimentary and diagenetic markers of the restriction in a marine basin: the Lorca Basin (SE Spain) during the Messinian: Sedimentary Geology, v. 121, p. 23-55.

Schroder, S., A. Bekker, N. J. Beukes, H. Strauss, and H. S. van Niekerk, 2008, Rise in seawater sulphate concentration associ-ated with the Paleoproterozoic positive carbon isotope excur-sion: evidence from sulphate evaporites in the 2.2-2.1 Gyr shal-low-marine Lucknow Formation, South Africa: Terra Nova, v. 20, p. 108-117.

Scott, C., B. A. Wing, A. Bekker, N. J. Planavsky, P. Medvedev, S. M. Bates, M. Yun, and T. W. Lyons, 2014, Pyrite multiple-sulfur isotope evidence for rapid expansion and contraction of the early Paleoproterozoic seawater sulfate reservoir: Earth and Planetary Science Letters, v. 389, p. 95-104.

Strauss, H., 1997, The isotopic composition of sedimentary sul-fur through time: Palaeogeography Palaeoclimatology Palaeo-ecology, v. 132, p. 97-118.

www.saltworkconsultants.com

Page 9: Salty atters

Page 9

Turner, E. C., and A. Bekker, 2016, Thick sulfate evaporite ac-cumulations marking a mid-Neoproterozoic oxygenation event (Ten Stone Formation, Northwest Territories, Canada): Geolog-ical Society of America Bulletin, v. 128, p. 203-222.

Warren, J. K., 1999, Evaporites: their evolution and economics: Oxford, UK, Blackwell Scientific, 438 p.

Warren, J. K., 2016, Evaporites: A compendium (ISBN 978-3-319-13511-3) Released Feb. 2016: Berlin, Springer, 1854 p.

Zemmels, I., P. L. Grizzle, C. C. Walters, and F. R. Haney, 1985, Devonian Novaculites as Source of Oil in Marathon-Ouachita Thrust System (abstract): Bulletin American Association Petro-leum Geologists, v. 69, p. 318-319.

www.saltworkconsultants.com

Page 10: Salty atters

Page 10

www.saltworkconsultants.com

John Warren, Chief Technical DirectorSaltWork Consultants Pte Ltd (ACN 068 889 127)Kingston Park, Adelaide, South Australia 5049

www.saltworkconsultants.com