<<

, bleaching, and CO2-driven fl ow

The footprints of ancient CO2-driven fl ow systems: Ferrous carbonate concretions below bleached

David B. Loope* and Richard M. Kettler Department of Earth and Atmospheric Sciences, University of Nebraska, Lincoln, Nebraska 68588, USA

ABSTRACT the vadose zone. Absolute dating of differ- In this paper, we present a conceptual model ent portions of these widespread concretions for bleaching of the sandstone and the down- Iron-rich carbonates and the oxidized could thus reveal uplift rates for a large por- ward transport of iron from the source rock remains of former carbonates (iron-oxide tion of the Plateau. Iron-rich masses in other to the geochemical trap where concretions concretions) underlie bleached Sand- sedimentary rocks may reveal fl ow systems nucleated and grew. First, we will describe stone over large portions of southern . with similar histories. and interpret a variety of iron-rich concre- Iron in the carbonates came from tions. Some still contain carbonate rims on sand grains in the upper Navajo INTRODUCTION and some do not, but we argue below that all that were dissolved when small quantities originated as reduced-iron carbonates. The of methane accumulated beneath the seal- Over a broad portion of the Pla- bleached rocks are striking evidence for a ing . As a second buoyant teau of southwestern , the large-scale fl ow system (Beitler et al., 2005; gas (CO2 derived from contains a wide variety of Chan et al., 2005). We claim that the concre- magmas ) reached the seal and migrated up iron-oxide–cemented masses. Here we explain tions reveal the changing composition, source, dip, it dissolved in the underlying water, why these structures are found down section oxidation state, and migration direction of the enhancing the solution’s density. This water and down dip from bleached sandstone. When fl uids in that system. carried the released ferrous iron and the carbon dioxide dissolves in an aqueous solu- In the Ladbroke Grove Gas Field (Otway methane downward. Carbonates precipi- tion, the density of that solution increases. This Basin, South Australia), the sand-

tated when the descending, reducing water fact has huge implications for subsurface car- stone reservoir contains high levels of CO2 that degassed along fractures. The distribution of bon sequestration (Lindeberg and Wessel-Burg, was derived from a magmatic source (Watson a broad array of iron-rich features made rec- 1997; Ennis-King and Paterson, 2005; Mac- et al., 2004). In a nearby gas fi eld, the same

ognition of the extent of the ancient fl ow sys- Minn and Juanes, 2013), and here we show that sandstone is not exposed to CO2-enhanced pore tems possible. Although siderite is not pre- it is essential to understanding iron diagenesis in waters. Analysis of cores shows that iron-rich served, dense, rhombic, mm-scale, iron-oxide the Navajo Sandstone. carbonate minerals are much more abundant pseudomorphs after ferrous carbonates are North of the Grand Canyon, in southern in the former reservoir than in the latter. These common. Distinctive patterns of iron oxide Utah’s , the upper part of the pore-fi lling carbonates occur mostly as con- were also produced when large (cm-scale), Navajo Sandstone forms the White Cliffs and cretions 2–10 cm in diameter. The Ladbroke poikilotopic carbonate crystals with multi- the lower Navajo forms the upper part of the Grove Field is an excellent analog for geologi-

ple iron-rich zones dissolved in oxidizing Vermillion Cliffs (Fig. 1). Mudrocks and evapo- cal storage of CO2 (Watson et al., 2004), and we waters. Rhombic pseudomorphs are found rites of the Carmel Formation argue here that—before it was exhumed during in the central cores of small spheroids and form an effective seal at the top of the Navajo. the —the Navajo Sandstone contained

large (meter-scale), irregular concretions Previous workers have shown that the upper CO2, methane, and abundant, iron-rich carbon- that are defi ned by thick, tightly cemented Navajo was originally red and that the iron in ate concretions. rinds of iron-oxide–cemented sandstone. the concretions came from iron liberated dur- Ferrous iron, carbon dioxide, methane, The internal structure and distribution of ing the bleaching of the overlying rocks (Beitler and siderite have been widespread in the pore these features reveal their origins as iron- et al., 2005; Chan et al., 2005). We agree with waters of Earth’s sediments and sedimentary carbonate concretions that formed within a these conclusions, but they raise two previ- rocks for billions of , but iron oxides have large-scale fl ow system that was altered dra- ously un answered questions: How, if the iron- been widespread on Earth only since the Early matically during Neogene uplift of the Colo- oxide concretions came from the bleached rocks . Recognition that iron-oxide masses rado Plateau. With rise of the Plateau, the above, was the iron moved down section and in the Navajo Sandstone had ferrous carbon- iron-carbonate concretions passed upward down dip? And, if the concretions formed as pri- ate precursors (Loope et al., 2010, 2011, 2012; from reducing formation water to shallow, mary precipitates from the mixing of reducing Kettler et al., 2011; Weber et al., 2012) and that

oxidizing groundwater fl owing parallel to and oxidizing waters (Beitler et al., 2005; Chan those carbonates were emplaced by CO2-driven modern drainages. Finally they passed into et al., 2005), how was oxidizing water intro- fl ow systems should aid interpretation of other duced and maintained below the widespread iron-oxide accumulations on Earth and on *Corresponding author email: [email protected] reducing waters? other planets.

Geosphere; June 2015; v. 11; no. 3; p. 1–15; doi:10.1130/GES01094.1; 12 fi gures; 1 table. Received 26 June 2014 ♦ Revision received 20 January 2015 ♦ Accepted 25 March 2015 ♦ Published online ___ Month 2015

For permission Geosphere, to copy, contact June [email protected] 2015 1 © 2015 Geological Society of America Loope and Kettler

′ Tmv Miocene volcanics 30 miles N 38° 28 50 km Jna Jc N Tov Oligocene volcanics Tov Te lake beds 24 Jc Jurassic Carmel Fm pre-J Jna Jurassic Navajo SS A DF* qu study site ar ius 12 with former Pla ′ teau 12 road or preserved carbonates Tmv Circle Cliffs Uplift Escalante Anticline Proposed,Jna subsurface paleoflow paths of: 2 89 W 110° 54 mantle-derived CO2 (buoyant; up dip) *

t CC Te l u water3 carrying CO2 (aq) a – 2+ F HCO CH 4 and Fe t 3 n 12 (high density; down dip) u Tov g u t l a s u a n F u 12 E * e a P n e a Jna c n i i - l r r mJ-K c u o

H n

Te * o

DF White M mJ-K

b Cliffs a b i Cliffs Jna a White K RG Sevier Fault

Vermillion * Jc

″ 9 MN* WY Cliffs ZNP * 30 ′ ZT White Cliffs UT NV study WC area * Vermillion Jna BG CO

W 113° 12 Vermillion Cliffs 59 89 * PC pre-J Cliffs * N 37° 00′ AZ Figure 1. Simplifi ed geologic map of southwestern Utah showing distribution of Mesozoic and Cenozoic rocks, major structures, locations of study sites, and postulated, subsurface paleofl ow paths of buoyant, supercritical

CO2 and, in the eastern part of study area, paths of dense water carrying ferrous iron and aqueous CO2 (alter- native explanation for the eastern paths [black arrows] is hydrodynamic fl ow; Loope et al., 2010). Flow paths of dense waters in the western area are unknown. SH—Sand Hollow; RG—Russell Gulch; ZT—Zion Tunnel; 3LC—Three Lakes Canyon; MN—Mollie’s Nipple; BG—Buckskin Gulch; PC—Paria Canyon; CC—Calf Creek; E—Egypt; DF—Dry Fork. ZNP near southwest corner is . For generalized cross section of the Grand Staircase: http:// en .wikipedia .org /wiki /Grand _Staircase -Escalante _National _Monument #mediaviewer /File: Grand _Staircase -big .jpg.

Previous Work primary precipitates formed during the mixing Study Area and Geologic Setting of iron-bearing, reducing waters and oxygen- Chan et al. (2000, 2005), Beitler et al. ated meteoric waters. This hypothesis guided This paper is based on fi eld observations and (2005), and Nielsen et al. (2009) carried out fi eld-based work (Potter and Chan, 2011), analyses of four different types of samples from extensive studies on both the bleaching of the bench experiments (Chan et al., 2007; Barge the Navajo Sandstone at ten localities (Fig. 1): Navajo Sandstone and the iron-oxide concre- et al., 2011), and iron isotope studies (Chan (1) rinded, iron-oxide–cemented concretions; tions within it. Nielsen et al. (2009) showed et al., 2006; Busigny and Dauphas, 2007). (2) sandstone with poikilotopic, iron-zoned that although the upper Navajo Sandstone is Loope et al. (2010, 2011, 2012) and Kettler ferroan cement; (3) calcite concretions bleached in the southern part of Zion National et al. (2011) reinterpreted the concretions as the containing large, rhombic, iron-oxide pseudo- Park (and continuously for 100 km eastward; altered (oxidized) remains of precursor concre- morphs or iron-oxide accumulations along Fig. 1), bleaching does not extend to the north- tions cemented by ferrous carbonate minerals. cleavage planes; and (4) iron-oxide patterns western portion of Zion, where the Navajo Weber et al. (2012) presented evidence that preserved in non-calcareous sandstone. Three retains its early diagenetic, red color. Although iron-oxidizing microbes mediated the oxida- of our sites (Fig. 1) are down dip from the this color-transition zone is only ~2 km wide, tion of the carbonates. Reiners et al. (2014) Escalante and Circle Cliffs Anticlines, along it does not appear to be structurally controlled measured (U-Th)/He ages and element con- the fl ow path defi ned by pipe-like concretions (Nielsen et al., 2009, p. 84). Beitler et al. (2005) centrations in iron oxide from Navajo cements, in the Navajo Sandstone (Loope et al., 2010). proposed that the iron-oxide concretions are concretions, and fracture fi lls. Two lie along the east fl ank of the Kaibab

2 Geosphere, June 2015 Concretions, bleaching, and CO2-driven fl ow upwarp, and four other sites are in fl at-lying sands of years from a faulted reservoir in the vary from less than 1 mm to 3 cm in thickness, strata below the White Cliffs west of the Kaibab Navajo Sandstone (Shipton et al., 2005; Heath and the sandstone surrounded by rinds can be monocline and east of the Hurricane fault. One et al., 2009; Burnside et al., 2013). Wigley et al. iron poor. In many concretions, however, cen- site occupies part of a gentle syncline, just west (2012) and Kampman et al. (2014) showed that trally located cores contain abundant iron-oxide of the Hurricane fault (SH, Fig. 1). the escaping fl uid has extensively bleached the cement. Some of this cement forms dense, rhom- The Navajo Sandstone was deposited near the overlying Jurassic and that bic pseudo morphs (Fig. 2; Loope et al., 2011; western edge of Pangea in a vast, ferroan carbonates are forming at the reac- Kettler et al., 2011). The outer (rind) portion of sand sea (Kocurek and Dott, 1983). Large-scale tion front between bleached and unbleached massive ironstones at Russell Gulch is composed cross strata composed of wind-ripple and dry sandstone. of mm-scale, crinkled bands of iron-oxide– avalanche (grain-fl ow) deposits make up more Three major, N-S–trending normal faults cut cemented sandstone, and cross sections of some than 95% of the formation. In contrast to many Mesozoic rocks of the study area (Fig. 1). Initial of these large masses reveal non-banded cores marine and fl uvial depositional environments, movement along the Hurricane and Sevier faults (Fig. 3A). Parallel, iron-oxide streaks or “comet organic matter was very sparse in the Navajo was probably 15–12 million years ago (Late tails” extend from many spheroids, and the azi- erg, so pore waters remained oxidizing long Miocene; Davis, 1999); the youngest rocks cut muths of the streaks are parallel to the trends of after deposition. by Paunsaugunt fault are 20 million years old elongate, pipe-like concretions in nearby out- The study area lies within the west-central (Bowers, 1991). Offset along these faults clearly crops (Fig. 4). At Calf Creek, boxworks are sur- , in the transition zone to the postdates bleaching of the Navajo Sandstone rounded by thick iron-oxide rinds and contain Basin and Range Province to the west. The (Nielsen et al., 2009). additional rinds where small joints are present Kaibab upwarp, Circle Cliffs upwarp, and Our three westernmost study sites lie within within the structure (Fig. 3B). the Escalante anticline are compressional, the western Grand Canyon volcanic fi eld, an Pipe-like concretions composed of thousands Laramide structures that developed 80–40 Ma, area with hundreds of basaltic cinder cones and of iron-rich, cm-scale, non-calcareous spheroids above reverse faults in basement rocks. Dickin- lava fl ows (Karlstrom et al., 2007; Crow et al., are abundant at Sand Hollow (Figs. 3C and 3D). son et al. (2012) showed that the eroded nose 2008; Karlstrom et al., 2012). The fl ows closest Centimeter-thick, iron-oxide–cemented rinds of the Kaibab upwarp was buried by fl at-lying, to our sites range in age from 1.4 million years delineate the perimeters of many of these pipes. Middle Eocene (ca. 40 Ma) lacustrine strata that to less than 100,000 years (Biek et al., 2003). Pipes are developed in the plane of cross bed- now crop out at high elevation in the northern Most of these fl ows were emplaced during a sin- ding; their trends are oblique to both the strike part of the study area (Fig. 1). The Navajo Sand- gle eruptive cycle (Smith et al., 1999), and many and dip direction of cross beds. stone has been eroded from the southern por- fl owed down paleovalleys and now cap ridges tion of the Kaibab upwarp, and Kaibab because of their high resistance to erosion (the Interpretation of Rinded Concretions is exposed along its crest. inverted valleys of Hamblin, 1970). The rinded, iron-oxide–rich concretions in At least 11 Colorado Plateau anticlines pres- In the past six million years, the Colorado the Navajo Sandstone are the oxidized remains ently contain large volumes of carbon dioxide Plateau has been uplifted ~2 km (Pederson of concretions that were originally cemented

(Haszeldine et al., 2005). Gilfi llan et al. (2008) et al., 2002; Karlstrom et al., 2012, 2014). by siderite (FeCO3). Our previous work (Loope studied the noble gas geochemistry of fi ve natu- During uplift, pore waters of the Navajo Sand- et al., 2010, 2011, 2012; Kettler et al., 2011) has ral CO2 reservoirs from the Colorado Plateau stone—previously buried by thousands of emphasized the importance of siderite in the and Rocky Mountain provinces and concluded meters of younger strata—were recharged by origin of concretions defi ned by dense, iron- that the dominant sources of all fi elds were mag- rain and snow that fell on the high plateaus of oxide rinds or shells that surround uncemented matic. In the transition zone to the Basin and the northern part of the study area. This study centers. Weber et al. (2012) presented evidence Range Province, explosive igneous activity was reports observations of rocks exposed on can- that the rinds on the Navajo concretions are bio- widespread on the Plateau during middle and yon walls, but a large percentage of the Navajo signatures of iron-oxidizing microbes that medi- late Cenozoic time. The lavas exposed in the remains below the regional water table. Parry ated the dissolution and oxidation of siderite. northern part of our study area (Fig. 1) are part of et al. (2009) reported ferrous carbonate minerals The presence of iron-oxide rinds along closely the Marysvale volcanic fi eld, one of the largest in in Navajo Sandstone cores from the Covenant spaced, vertical joints cutting the large concre- western United States. Most calc-alkaline mag- oil fi eld in central Utah. tions at Calf Creek (Fig. 3B; Loope et al., 2011) matic activity occurred during the Oligocene and may help to constrain timing of siderite oxida- Miocene (32–22 Ma), and the less voluminous LATE DIAGENETIC CONCRETIONS tion in these rocks to the late Neogene. Fossen bimodal basalts and rhyolites were emplaced as AND CEMENTS (2007, 2010) hypothesized that, while they were recently as the Holocene (Rowley et al., 2002). buried at depths greater than ~1 km, relatively The Marysvale fi eld is at the eastern edge of a Rinded Iron-Oxide–Cemented Concretions few joints cut the porous and permeable reservoir caldera complex that was centered on what is of the Colorado Plateau; the closely now the Basin and Range Province. Some of the Ironstones, spheroids, boxworks, and pipes spaced joints in the sandstones formed only springs located along faults in the Grand Canyon defi ned by isopachous rinds or shells that are after these rocks underwent uplift and cooling issue mantle-derived CO2 and helium (Crossey densely cemented by iron oxide (Figs. 2 and 3) (Fossen , 2010). Because the Plateau was uplifted et al., 2006, 2009). Work on springs and traver- are widespread in the Navajo Sandstone. At Rus- 2 km in the past six million years (Karlstrom tines shows that although some portions of these sell Gulch, in west-central Zion National Park, et al., 2014), many of the joints in these sand- gases refl ect ongoing processes in the astheno- rinded concretions are present in two of the stones may have formed in the late Neogene. sphere, another strong component was released diagenetic subfacies defi ned by Nielsen et al. Using laboratory experiments, Chan et al. during the Oligocene (Crossey et al., 2009). (2009)—the pink altered subfacies that is prev- (2007) attempted to show that spheroidal, iron- Near the town of Green River, Utah, carbon alent in the middle Navajo and the brown fer- oxide rinds in the Navajo Sandstone could have dioxide has been escaping for hundreds of thou- ruginous subfacies in the lower Navajo. Rinds formed as primary precipitates. They showed

Geosphere, June 2015 3 Loope and Kettler

why rinds cannot be primary precipitates—two A steps, two iron minerals, and two diagenetic environments are required (Fig. 5). Under reducing conditions, closely spaced, siderite- cemented concretions nucleated and grew to form an amalgamated mass. In the labora- tory, Barge et al. (2011) generated intergrown masses of analogous, self-organized spheroids cemented by silver chromate. We view their experiment as a reasonable demonstration of the fi rst step in rind formation (precipitation of primary, intergrown spheroids—in the Navajo case, these were composed of the reduced iron , siderite). The second step began as the Navajo pore water became oxygenated. Iron-oxidizing microbes colonized the convo- luted, outer surface of the already-fused sid- B C erite spheroids (Fig. 5). The microbes slowly dissolved the siderite, and ferrous iron diffused to the perimeter, where it combined with oxy- gen. Iron oxide precipitated, forming a single, inward-thickening rind. Such a rind is thus a product of “batch processing”—its overall geometry refl ects the shape of a preexisting template composed of multiple, intergrown spheroids. If the rinds had originated de novo (without a template), they would now show a growth sequence (Fig. 5D) in which new sphe- roidal rinds were built upon the surfaces of older rinds (a pattern in which spheroids shared walls, as is seen in soap bubbles). Intergrown Navajo spheroids do not, however, share walls (Fig. 2). Iron-rich cores within these spheroids (Fig. 2; previously unreported from the Navajo) provide new evidence supporting our model; Figure 2. Spheroidal concretions with iron-oxide rinds (site RG, the cores are the corroded remains of the larger, Fig. 1). (A) Sawed with well-developed rind, iron-poor original template (Fig. 5). sandstone “moat,” and three central cores. In the fi rst stage (reduc- The southwest-directed “comet-tails” on the ing pore water), three siderite concretions nucleated at nearby sites Russell Gulch spheroids and nearby pipe-like and, with continued growth, merged to form a single mass. In the concretions (Fig. 4) show that alteration of the second stage (oxidizing water), iron oxide precipitated on the perim- siderite took place within groundwater that was eter of the mass and progressively thickened inward as the sider- fl owing southwestward. This is the same paleo- ite was dissolved and receded. Rind thickening ended when mass fl ow direction as that shown by a Late Pleisto- entered the vadose zone. (B) Small, dense, rhomb-shaped iron-oxide cene lava fl ow that fi lled a nearby paleovalley accumulations cementing sand grains within the core zone of con- (Fig. 2A). cretion; these are interpreted as pseudomorphs after siderite (sur- Previous studies have shown that rinds can viving evidence of the fi rst stage of concretion formation). (C) Closer form on other kinds of geologic templates. Fol- view of B (with condenser) shows preserved cleavage planes within lowing van der Burg (1969), Loope et al. (2012) pseudomorph. Sand grains are quartz, and blue is pore-fi lling epoxy. showed that spindle-shaped intraclasts that form lag gravels in fl uvial channels of the Cretaceous (eastern Nebraska) were that if, in an agar matrix, an Fe(NO3)3 solution prior to (secondary) rind formation. The ferric originally composed of sideritic fl oodplain silt is placed adjacent to and surrounded by a KOH iron neutralizes the base (KOH) that served as and developed rinds only after they were depos- solution, iron hydroxide will precipitate at the a chemical trap. In our model (Loope et al., ited in the channels. The abraded shapes of reaction front. Iron diffuses outward, a circu- 2010; Weber et al., 2012), a dissolving sider- the transported Dakota intraclasts dictated the lar rind forms, and the rind thickens inward as ite concretion provides diffusing ferrous iron shapes of the rinds that formed after transport long as iron continues to diffuse from the cen- ions that are trapped at a surrounding, and now surround them. ter. We argue that this experiment demonstrates gradient. Within the cores of spheroids, ironstones, and the importance of a preexisting (primary) con- Complex rinds such as the convoluted one boxworks, the oxidized remains of the original centration of iron in the center of the structure shown in Figure 2A illustrate another reason iron-carbonate crystals are preserved as rhombic

4 Geosphere, June 2015 Concretions, bleaching, and CO2-driven fl ow

Figure 3. Large, rinded, iron- A B rich masses. Hammer is 28 cm oldest rinds hammer long. (A) Block that fell from lava J an adjacent, giant, in situ J ironstone mass in the middle youngest rinds cross-strata defined by innermost rind Navajo Sandstone with a very pseudomorphs thick, laminated rind bounded a core at on its inner surface by an irreg- core ular dissolution surface that cross-str surrounds a non-rinded core. J As with smaller concretions, iron was initially concentrated outermost rind by growth of a ferrous carbon- C ate concretion; the carbonate D was later oxidized from the perimeter inward. Concretion lies below the contact between bedding plane valley-fi lling Pleistocene lava X and Navajo Sandstone (Site rinded, coalesced RG, Fig. 1). (B) In situ iron-rich concretions mass with thick outer (oldest) rind that surrounds multiple younger, thinner rinds and a complexly corroded core. Note 2.2 cm coin that some thin rinds follow small joints (J). These joints are E F interpreted as being very young (late Neogene), postdating for- mation of a large iron-car- bonate concretion (fi rst stage, Fig. 2). Oxidizing waters, fol- lowing bedding and joints, dis- solved the carbonates and sup- ported iron-oxidizing microbes that formed the rinds (second stage, Fig. 2). Preservation of rhombic pseudomorphs in the core suggests it was oxidized above the local water table (see Discussion) (site CC, Fig. 1). (C) Large, pipe-like concretions (white arrows) composed of cm-scale, coalesced, non-calcareous concretions. Perimeters of large pipes are delineated by dense, 1-cm-thick, iron-oxide–cemented rinds (site SH, Fig. 1). Pipes nucleated and grew parallel to the plane of cross-stratifi cation, oblique to direction of both dip and strike. X marks site of photo D. (D) Close-up of cm-scale, non-calcareous, spheroidal concretions cemented by iron oxides that com- prise the interiors of the pipe-like concretions in C. (E) Large, pipe-like concretionary mass (Entrada Sandstone, SE Utah; cf. Garden et al., 2001, fi g. 5) composed of spheroidal, cm-scale concretions cemented by ferroan calcite. (F) Close-up of E showing irregular margin of pipe. Interpretation: Unlike the Navajo concretions in C and D, this Entrada pipe does not have a thick, iron-oxide rind because ferroan calcite (unlike siderite) does not undergo complete dissolution in oxygenated water and does not support the growth of iron-oxidizing microbes. pseudomorphs after siderite or ankerite (Figs. The large, pipe-like concretions at Sand differences can be attributed to the fact that 2 and 3). Pseudomorphs are restricted to the Hollow (Fig. 3C) that are internally composed siderite (unlike ferroan calcite) dissolves com- cores of these concretions and are never visible of thousands of cm-scale spheroids (Fig. 3D) pletely in oxygenated water of near-neutral pH. in the rinds. Rinds formed and thickened only resemble the ferroan-calcite–cemented concre- when ferrous iron could diffuse away from the tionary pipes that Garden et al. (2001) described Poikilotopic, Iron-Zoned Ferroan dissolving (sideritic) centers. We interpret the from the Entrada Sandstone and interpreted as Calcite Crystals pseudomorph-bearing cores as in situ products products of CO2 degassing (Figs. 3E and 3F). of abiotic oxidation of siderite that took place The Navajo spheroids, however, are different in At Egypt and Dry Fork, large (up to 10 m in the vadose zone after cm-scale diffusion of that they are non-calcareous and are surrounded in diameter), dark concretionary masses of ferrous iron ceased (see Discussion and Fig. 12). by thick iron-oxide rinds. We argue that these calcite-cemented, cross-bedded sandstone

Geosphere, June 2015 5 Loope and Kettler A C

B

Figure 4. Flow-direction indicators for oxidizing groundwater (site RG, Fig. 1). (A) Streaks (“comet tails”) extending southwestward, away from spheroidal concretions. (B) Outcrop showing that comet tails are three-dimensional structures preserved in the rock, not surfi cial stains. (C) Pipe-like concretions (defi ned by thin rinds of iron-oxide cement) that trend NE–SW. Note that these elongate rinds share walls and formed sequentially as the conduits carrying oxygenated water expanded radially into a reduced, siderite-bearing aquifer (see Loope et al., 2011, fi g. 8). See Figure 12 for relationship of fl ow direction of oxidizing water to fl ow direction of Pleistocene basalt. are nucleated on numerous shear fractures Figure 5. Evolution of rinded and cored con- Not Observed Observed and accompanying horsetail fractures (Fig. cretions. Siderite concretions (A) nucleated 6A; Fossen, 2010). These structures weather in reducing water at closely spaced sites and A BC into strong relief, are associated with bleached became intergrown. When oxidizing water started to invade, iron-oxidizing microbes sandstone, and trend N70°W—parallel to two M normal faults in the Navajo Sandstone directly colonized the perimeter of the concretion ic robia ion east of Egypt (Utah Geological Survey, 2014). (Weber et al., 2012). As metabolism pro- l Alterat A Centimeter-scale, poikilotopic calcite “sand ceeded, an iron-oxide rind formed on the b i o crystals” cement the bands and surrounding outer surface of the concretion and thickened t i c bedrock (Fig. 6B). Modern lags composed of inward as dissolution caused the siderite to A l individual or twinned crystals are widespread. recede (B). No siderite is observed today on te ra C′ Individual crystals are hexagonal in cross sec- outcrop; the siderite of some concretions was D t io tion. Surface weathering of the individual completely exploited by the microbes (C); n C ′ crystals reveals concentric iron-rich zones other concretions (C ) have oxidized cores siderite C (Fig. 6B). The δ13C values of poikilotopic fer- (c) with abundant iron-oxide pseudomorphs cement roan calcite from Egypt range from –2.78‰ to after siderite. Note thinner rind of C′ com- C –3.48‰ (Table 1). pared to C. Interpretation: cored concre- iron oxide C tions emerged into the vadose zone before cement C the dissolution of siderite and the diffusion C Interpretation of Poikilotopic, Iron-Zoned friable C Ferroan Calcite Cement of ferrous iron to the rind were complete. sandstone Ferroan calcite can be preserved in a sys- A concretion composed of iron-oxide rinds tem where siderite and ankerite are dissolved could theoretically form via a single process because, in a mineral where only a small over thousands of years, within a single diagenetic environment (mixing of oxidizing and percentage of the divalent cations are fer- reducing water; Chan et al., 2007). We argue that such a concretion (D) would have shared rous iron, the amount of acid produced upon (rind) walls, showing that the larger mass formed in a decipherable time sequence (similar oxidation is small compared to the total buff- to soap bubbles). In this example, older rinds are present at the bottom, and younger at the ering capacity of the mineral (Loope et al., top. Spheroidal concretions in the Navajo Sandstone do not share walls; the observed rinds 2010). The presence of iron in the calcite formed via a batch process—the intergrown precursor (cemented by siderite) formed over a crystal lattice indicates this cement was pre- long period of time in a single diagenetic environment and only later (in a second diagenetic cipitated from reducing pore water. As with environment) acted as a template for the (secondary) iron-oxide–cemented rind.

6 Geosphere, June 2015 Concretions, bleaching, and CO2-driven fl ow

A B relatively high δ13C values present in these calcite cements indicate that methane was not the major source of carbon in their formation (Wigley et al., 2012).

Calcite Concretions Containing Large, Iron-Oxide–Filled Rhombic Pseudomorphs or Iron-Oxide Accumulations along Cleavage Planes

At Egypt, Dry Fork, Buckskin Wash, Paria Canyon, and Three Lakes Canyon, many spheroidal calcite-cemented concretions are preferentially developed in the lower portions of grain-fl ow strata. Some spheroids are isolated (Fig. 7C), but others have merged into multi- ples, or have coalesced into large, slab-sided masses (Fig. 7A). Thin sections of the spheroids at Paria Canyon reveal porous, rhombic struc- tures partially fi lled with iron oxide (Fig. 7D). At Dry Fork, white calcite-cemented concretions envelope many distinct iron-oxide–rich rhombs (Fig. 7E). These concretions are surrounded by a speckling of iron-rich masses that are identical in size to the rhombic structures within the con- cretions (Fig. 7F). Nielsen et al. (2009) showed that similar “speckles” are extremely common in the Navajo in Zion National Park. At Buckskin Wash, Paria Canyon, and Three Lakes Canyon, some concretionary masses are dominantly com- posed of coarse-grained calcite, but iron oxide is abundant along cleavage planes and former crys- tal faces. At Buckskin Wash, large concretionary masses cemented by ferroan calcite are abun- dant. The upper, weathered surfaces of some of Figure 6. Concretions cemented by poikilotopic ferroan calcite (site E, Fig. 1). (A) Coarse these masses bear grooves with amplitudes up calcite cementing horsetail fractures. (B) Large poikilotopic calcite “sand crystals” that to one meter and wavelengths from 15 to 60 cm engulfed hundreds of sand grains. Dark bands are iron-rich zones. Note that some iron-rich (Fig. 8A). In cross sections along canyon walls, zones encircle numerous intergrown sand crystals (white arrow). calcite-cemented zones stand out in positive relief but show highly irregular bound aries (Fig. 8B). Resistant rock contains abundant, 6-mm- boxworks and pipe-like concretions in the deformation bands during the early, dilational diameter, spheroidal concretions that are largely Navajo that are associated with joints (Loope stages of the bands when their permeability cemented by dark, ferroan calcite. The outer por- et al., 2010), the association of this ferroan was enhanced. Precipitation of the copious tions of these concretions are cemented by light- calcite with brittle fractures indicates that the ferroan calcite adjacent to the fractures in our colored, iron-free calcite (Fig. 8B). The non- ferroan calcite cements are late diagenetic. fi eld area may have been localized by a rise resistant rock contains iron-stained spheroids In a nearby fi eld area, Parry et al. (2004) in alkalinity of the pore water due to degas- that are calcite free (Fig. 8C). The δ13C values concluded that reducing water bleached por- sing of CO2. Although methane was likely of calcite from Paria Canyon and Dry Fork range tions of the Navajo Sandstone that surround the reductant that bleached the sandstone, the from –1.58‰ to –4.97‰ (Table 1).

Interpretation of Calcite Concretions Containing Large, Rhombic, Iron-Oxide TABLE 1. LOCATIONS, DESCRIPTIONS, AND ISTOPIC DATA FOR CALCITE SAMPLES Pseudomorphs or Iron-Oxide Accumulations Sample δ13C δ18O Location no. Description (‰ VPDB) (‰ SMOW) along Cleavage Planes Egypt 1 Calcite sand crystal –2.78 12.90 These calcite concretions (Figs. 7A and 7C) Egypt 2 Calcite sand crystal –2.81 13.70 aid interpretation of other concretions in the Egypt 3 Calcite sand crystal –3.48 13.26 Dry Fork 1 White discoid –1.58 11.01 Navajo. The sizes and shapes of individual cal- Dry Fork 2 White discoid –3.07 11.76 cite spheroids are similar to those of the non- Paria Canyon 1 Large spheroid –4.97 17.13 calcareous concretions with iron-oxide rinds Note: See Figure 10. SMOW—standard mean ocean water; VPDB—Vienna Peedee belemnite. (Fig. 2A) that we have interpreted as the remains

Geosphere, June 2015 7 Loope and Kettler of former siderite concretions (Loope et al., Interpretation of Iron-Oxide Patterns Fork (Figs. 7E–7G), we interpret them as the 2010, 2011; Weber et al., 2012; this paper). The Preserved in Non-Calcareous Sandstone oxidized remains of ferrous carbonate crystals. fl at-topped, coalesced calcite concretions that As with the rinded concretions, we interpret Nielsen et al. (2009, p. 73) noted that, at Zion, develop in grain-fl ow strata have an overall form these iron-oxide cements as the alteration prod- “speckles” are widespread within the bleached similar to some of the rinded boxworks (com- ucts of precursor ferrous carbonates. Cross sec- portions of the upper Navajo. With their mix- pare Figs. 7A and 7B). tions of many carbonate crystals are rhomb and ing model for the primary (iron-oxide) origin Acid is generated by the hydrolysis that hexagon shaped, and the carbonate crystals that of the speckles, it is necessary to reintroduce accompanies oxidation of Fe+2 to Fe+3 and grow in sandstones are very commonly large oxygenated waters below reducing water so that subsequent precipitation of oxyhydroxides. and poikilotopic (see Fig. 6B). We interpret the iron oxide can be emplaced as a primary phase Carbonate minerals with more than 50 mol% patterns defi ned by the polygons (Fig. 9) as the in a previously bleached zone (“overprinting” Fe are likely to be completely dissolved by a remains of poikilotopic carbonate euhedra— of Nielsen et al., 2009). Our altered-precursor self-promoting oxidation reaction (Loope et al., “sand crystals” that engulfed hundreds of sand interpretation of these iron-oxide–cemented 2010, p. 1002). Dissolution of carbonates with grains. These euhedra contained both iron-rich masses is consistent with their presence in less iron consumes acid and is likely to lead to and iron-poor zones. All of the carbonate in bleached sandstone, because both bleaching of precipitation of iron-free calcite. We therefore these crystals has been dissolved, leaving only the sandstone (dissolution of iron oxide) and interpret the cm-scale concretions at Dry Fork iron oxide to mark the former positions of iron- cementation by ferrous carbonates necessarily and Paria Canyon (Figs. 7C–7G) to have had rich zones. Acid generated during oxidation of took place in reducing waters. <50 mol% Fe. the iron-rich zones (Loope et al., 2010) would We interpret the “rhombic structures” within have aided dissolution of the adjacent iron-poor DISCUSSION these concretions as pseudomorphs after either calcite cement. siderite (rhombs fi lled with iron oxide) or The “speckles” that Nielsen et al. (2009) Carbon Dioxide Giants or ankerite (rhombs retaining considerable pore described from Zion National Park are abun- Broad Flow Paths? space and partially fi lled by iron oxide). The dant in the Navajo in other parts of the Colorado “speckles ” that lie outside the concretions (Fig. Plateau and, based on their geometry and their Using satellite imagery and fi eld-based spec- 7F) are the same size and general shape as those resemblance to the pseudomorphs surrounding tral refl ectance measurements, Beitler et al. that are enclosed in calcite, but their margins are and within calcite concretions at Egypt and Dry (2003, 2005) argued that the bleached sand- not as well defi ned. We interpret them as pseudo- morphs after isolated iron-carbonate crystals. The grooved surfaces and irregular shapes of sandstone masses at Buckskin Gulch that Figure 7 (on following page). Carbonate and iron-oxide cements and concretions. (A) Wedge- are cemented by ferroan calcite (Figs. 8A and shaped toes of eolian grain fl ows (GF; site E, Fig. 1) are preferentially cemented by ferroan cal- 8B) could indicate large-scale amalgamation of cite and surrounded by uncemented wind-ripple deposits (wr). White arrows point to planar , cemented zones. Another possibility is that they upper surface of cemented wedge—the former avalanche face of an eolian dune. Cementation indicate subsurface dissolution of ferroan calcite began with spheroidal concretions and ended with slab-sided masses. (B) Boxwork concre- by acidic water. Acid could have been generated tion (iron oxide, no preserved carbonates; site CC, Fig. 1) that is bounded by a thick iron- either during siderite oxidation or by dense, oxide rind with a planar, upper surface (white arrows; top of grain-fl ow ). Internal

CO2-charged water descending from the nearby compartments are defi ned by iron oxide that precipitated along two perpendicular sets of Kaibab monocline. The iron-free rims on the joints (J). Two cores (c) occupy the centers of the two visible compartments. Interpretation: small concretions (Fig. 8C) could be the result Siderite preferentially developed in grain-fl ow strata (as did calcite in A); upon oxidation, the of a shift from reducing to oxidizing water or iron-oxide rind followed the original grain fl ow–wind-ripple boundary (the upper edge of the could have precipitated from oxidizing water as former, siderite-cemented mass). (C) Oblate-spheroidal concretions in wind-ripple deposits ferroan calcite underwent dissolution. at Paria Canyon (PC; Fig. 1). (D) Plane-light micrograph from a concretion (as in C and E). As with the calcite from Egypt, the high Calcite (c) is preserved, but the rhombic, iron-rich areas (white arrows) are pseudomorphs δ13C values of the Dry Fork and Paria Canyon of an iron-rich carbonate phase (probably ankerite) that was originally in optical continuity samples indicate that little of their carbon was with the calcite but was dissolved, leaving iron oxide (black) and pore space (blue). These derived from methane oxidation. rhombs resemble those seen dispersed in sandstones that lack calcite cement—the “speckles” of Nielsen et al. (2009) and in the core stones shown in Figure 2. (E) Sawn halves of a single Iron-Oxide Patterns Preserved in calcite-cemented concretion from Paria Canyon (as in C and D). Black arrow points to a Non-Calcareous Sandstone central, rhomb-shaped, iron-rich zone. White arrows point to a thin, rhomb-shaped, iron- oxide–rich zone that defi nes the perimeter of a large, heavily altered carbonate crystal. Inter- At Zion Tunnel and Russell Gulch, iron- pretation: Two zones (with the same poikilotopic crystal) that were originally composed of oxide cements form polygonal masses (mm siderite or ankerite dissolved in oxidizing pore water, leaving iron oxide to mark their former scale) and patterns (cm scale). Some patterns positions. Small, dark structures are pseudomorphs (ps) also shown in D and G. (F) Sand- comprise nested polygons defi ned by narrow stone exposed at Dry Fork with isolated, dark speckles (S) and large concretions with dense, bands of iron-oxide cement (Fig. 9). Scattered calcite-cemented cores (black arrows) and perimeters composed of closely spaced speckles small masses produce a “speckled” appearance (white arrows). (G) Sawed sections of the large white cores of concretions shown in (F). The (Nielsen et al., 2009; Fig. 7E). These workers dense, calcite-cemented sandstone contains numerous scattered, 3- to 4-mm-diameter, iron- noted that most of the “speckles” are within oxide masses (black arrows) interpreted as pseudomorphs after iron-rich carbonate crystals. their pink and their brown diagenetic facies but These closely resemble the iron-rich areas shown in D and the “speckles” shown in F; all are also occur in their white facies. interpreted as rhombic pseudomorphs after siderite or ankerite.

8 Geosphere, June 2015 Concretions, bleaching, and CO2-driven fl ow

A wr GF B

J C wr GF GF wr J C wr

C D C Q

C C

0.5 mm E F S

S 5 cm

G 3 cm

Figure 7.

Geosphere, June 2015 9 Loope and Kettler

stones across many of the Colorado Plateau anti- clines mark the positions of “exhumed hydro- A carbon giants.” No bitumen, however, has been reported from bleached Navajo Sandstone, and only a very small quantity of bitumen is present in the Entrada Sandstone (at one site). We inter- δ13 iron oxide pret the C values (Fig. 10) of the carbonates associated with the bleached rocks as evidence that a portion of the carbon was derived from organic sources. This evidence, combined with the widespread surviving and altered carbonate concretions, leads us to agree with Haszeldine et al. (2005) that bleaching was accomplished via large volumes of carbon dioxide and a lesser amount of methane. calcite Wigley et al. (2012) studied bleached Entrada Sandstone and carbonate cements associated

with active CO2 seeps along a fault near the crest of an anticline at Green River, Utah. They B concluded that the Entrada was bleached by an aqueous solution in which carbon was pre-

uGF dominantly oxidized (CO2; 93%), with lesser uGF amounts of methane (7%): uGF uGF 20Fe2O3(hematite) + 5CH4(aq) + + 64CO2(aq) + 19H2O + 11H (aq) = (1) 30Fe2+ + 10FeHCO + + 59HCO – . uGF (aq) 3(aq) 3(aq) The δ13C values from cement samples at Green River (Wigley et al., 2012) were only slightly heavier than those reported by Beitler et al. (2005). Because our δ13C values (mean of –3.12‰), are very near those reported by Wigley et al. (2012), they, too, are more consis- C tent with bleaching by a fl uid carrying mostly inorganic carbon. Nielsen et al. (2009) emphasized that the uGF bleaching of the upper Navajo could not have taken place under static conditions—sustained fl uid fl ow was necessary to strip iron oxide from sand grains. Furthermore, Parry (2011) showed that, because of the low solubility of ferrous iron, the growth of iron-rich concretions in the Navajo Sandstone required advective fl ow through the aquifer for thousands of years. We propose that, over a large portion of our study area, two linked, density-driven fl ow systems operated to bleach the sandstone and to trans- port suffi cient iron to precipitate the concre- tions: (1) buoyant up-dip fl ow of methane and Figure 8. Possible evidence for subsurface dissolution of late diagenetic cement composed supercritical carbon dioxide; and (2) convective of ferroan calcite. (A) Upper surface of a broad mass of calcite- and iron-oxide–cemented fl ow driven by the enhanced density of waters sandstone. (B) Irregular distribution of ferroan calcite cement in Navajo cross strata. carrying aqueous CO2 in contact with overly-

Grain-fl ow strata are preferentially cemented, but large, irregular patches of uncemented ing, migrating or trapped supercritical CO2 (Fig. grain fl ows (uGF) suggest that some cements have been dissolved. (C) Boundary between 11). Dissolution can yield a solution that is ~5 cemented (below dashed line) and uncemented grain fl ows. Small spheroids within uGFs are wt% CO2; Bickle (2009) notes that if the reser- not calcareous and are stained by iron oxide. Note the thin rims of light-colored, iron-poor voir has high permeability, a 50-m-thick layer calcite on each ferroan-calcite–cemented spheroid below the dashed line. of CO2 can be dissolved in ten years. Upslope

migration continually exposes migrating CO2 to unsaturated waters, making dissolution more

10 Geosphere, June 2015 Concretions, bleaching, and CO2-driven fl ow A B

GF

GF GF GF GF GF GF

WR

C D

WR

WR

Figure 9. Patterned iron-oxide cements in non-calcareous sandstone (site RG, Fig.1). (A) Grain-fl ow (GF) tongues prefer- entially cemented by iron oxide (cf. Fig. 7A). (B) Close-up grain fl ow (as in A) with nested polygonal patterns defi ned by iron-oxide bands. Arrowed bands surround multiple pattern elements. Interpretation: Iron-oxide bands are the oxidized and dissolved remains of large, iron-zoned, euhedral carbonate crystals. (C) Large swath of iron-oxide stain that cuts across wind-ripple crossbedding in the white subfacies of the Navajo Sandstone. Upper margin of swath grades into patterned iron oxide as shown in B and “speckles” (Figs. 7E–7G). Lower margin (arrow) contains distinct, dark iron-oxide bands. (D) Close-up of band arrowed in C. Interpretation: The dark, continuous band (arrows) is the oxidized and dissolved remnant of an iron-rich zone that grew within sandstone that was densely cemented by carbonate minerals with varying iron content. WR—wind-ripple deposits. effi cient; descent of the density-enhanced water –2 (charged with aqueous CO2) can arrest the Figure 10. Plot of carbon and upslope fl ow of the buoyant fl uid (MacMinn and oxygen isotopes for calcite –4 Juanes, 2013; Fig. 11). The great extent and rel- cements and concretions from atively uniform thickness of the bleached rock the Egypt, Dry Fork, and Paria –6 on the Colorado Plateau (e.g., the White Cliffs, Canyon localities (red dots; see

Fig. 1) refl ect the interaction of these two fl ow Table 1) and comparison with (‰ VPBD) –8 3 systems. The low northward dip of the Navajo values from other sandstones

Sandstone along the east-west–oriented White in the region. Triangles mark CaCO –10 Navajo Sandstone C this study Cliffs of southwestern Utah (Fig. 1) suggests values from Chan et al. (2000), 13 δ –12 that CO2 migrated southward toward a broad, Eichhubel (2009), and Garden Entrada Sandstone now eroded structural high. The abundance of et al. (2001. Modified from Wigley et al. (2012) –14 concretions along the east fl ank of the Kaibab Wigley et al. (2012). VPDB— Jurassic sandstones Eichhubl et al. (2009) upwarp indicates that CO2 very likely accumu- Vienna Pee Dee belemnite; –16 lated along the crest of that structure. Rather VSMOW—Vienna standard 6 9 12 15 18 21 24 27 than developing only within or immediately mean ocean water. δ18O (‰ VSMOW) CaCO3 below giant CO2 reservoirs, however, a large

Geosphere, June 2015 11 Loope and Kettler

water by oxygenated meteoric water, oxidation Jna of iron-rich carbonates, and drainage of much of the Navajo aquifer. T-1 unaltered, water-saturated, red sandstone, After they were eroded during latest Creta- no concretions seal (Jc) supercritical CO2 ceous and early Paleogene time, tilted Creta- ceous rocks along the Kaibab monocline were buried by fl at-lying, Middle Eocene lacustrine bleached rock Jna sediment (ca. 40 Ma; Eaton et al., 2011; Dick- inson et al., 2012). The Navajo aquifer was plumes of dense water carrying dissolved intact at this stage; its highest elevation (crest T-2 CO2, methane, and ferrous iron (Fe 2+ ) of the Kaibab Plateau) would have approxi- mated that of the Eocene lake beds. Post-uplift, those beds crop out along the southern edge of bleached rock Jna Aquarius Plateau, just east of the Paunsaugunt fault (Fig. 1; presently ~3100 m). Oligocene T-3 red, pink, and brown sandstone, volcanics started to erupt at 32 Ma and buried

with siderite and ferroan calcite concretions the lake beds. After release of supercritical CO2 3° from intrusive magma and its migration into the Navajo, dissolution of carbon dioxide into the Figure 11. Model for bleaching of upper Navajo Sandstone and precipitation of ferrous con- pore waters immediately below the supercriti- cretions in the middle Navajo. T-1: Moving up dip, just below the sealing Carmel Forma- cal CO2 slowed or stopped its buoyant, up-dip tion, buoyant, supercritical carbon dioxide and a small quantity of methane starts to accu- migration (MacMinn and Juanes, 2013; Fig. mulate in the upper Navajo Sandstone. Dissolution of CO2 into underlying water begins. 11). Methane dissolved in the water bleached T-2: Supercritical carbon dioxide in contact with underlying formation water continues to the upper Navajo and, as CO2(aq) accumulated, move up dip and to dissolve. As aqueous carbon dioxide adds mass to the solution, the dense the fl uid moved downward, transporting ferrous fl uid moves downward as fi nger fl ow. Methane reduces hematite on the rims of sand grains, iron and remaining methane into lower strata. releasing ferrous iron into solution and bleaching the rock. The descending water transports As extensional faulting started to break up the ferrous iron and remaining methane downward. T-3: Carbon dioxide is depleted, and down- western Colorado Plateau (12–15 Ma; Davis, ward fl ow has ceased. Siderite and ferroan calcite have precipitated in the middle Navajo 1999), degassing adjacent to joints would have (perhaps facilitated by localized degassing of carbon dioxide). Up-dip extent of bleached increased alkalinity of the water, thereby trig- rock marks limit of CO2 migration, a possible explanation for the presence of unbleached gering precipitation of carbonate minerals. sandstone in northernmost Zion National Park (noted by Nielsen et al., 2009). Final thick- Eventually, block faulting ruptured and offset ness of bleached zone refl ects quantity of methane available for reduction of ferric iron the seal formed at the top of the Navajo, releas- during descent of plumes. Modifi ed from MacMinn and Juanes (2013, fi g. 1). Jna—Jurassic ing the buoyant fl uids and ending convective Navajo Sandstone; Jc—Jurassic Carmel Formation. fl ow within the Navajo aquifer. Recently published (U-Th)/He ages of Colo- rado Plateau iron-oxide accumulations indi- proportion of the bleached rocks and carbon- the Grand Canyon, thereby generating southern cate they precipitated between 0.21 and 25 Ma ate concretions most likely developed along Utah’s Grand Staircase. To bleach the upper (Reiners et al., 2014). These are likely oxidation gently sloping fl ow paths along which the CO2 Navajo across the crest of the Kaibab upwarp ages—helium accumulation had to wait until and methane never reached a structural trap (the from the west of the Hurricane fault to the Kai- after precursor siderite was altered to iron oxide. dense water descended before reaching a trap or bab monocline—a distance of 110 km—required Entry of oxygenated, meteoric water into the the ancient land surface; Fig. 11). a continuous, confi ned aquifer. Bleaching the Navajo initiated rind-forming, microbial oxida- Hydrodynamic fl ow may have also played upper Navajo, precipitating ferrous carbonates, tion of siderite. Collapse in the Basin and Range a role in the bleaching and emplacement of and oxidizing those carbonates required the fol- led to headward erosion and deep dissection of concretions. The positions of the Escalante lowing sequence of geologic events: (1) com- Plateau strata. Lucchitta et al. (2011) made a anticline and the Aquarius Plateau were suit- pressive stresses tilted reservoir and sealing strong case that a river fl owing southwestward able for hydrodynamic fl ow through the Navajo rocks into broad folds; (2) methane from ther- from Colorado was coursing through the Kaibab Sandstone during portions of the middle and late mally mature strata migrated northward along upwarp in what is now eastern Grand Canyon

Cenozoic (Loope et al., 2010). A sister anticline with supercritical CO2 released from magma; by the Middle Miocene. Karlstrom et al. (2014) just west of the Escalante structure (the Upper (3) waters with enhanced density fl owed down- calculated that, by 6 Ma, the river crossing the Valley anticline) presently has hydrodynamic ward from the base of migrating or trapped Kaibab upwarp had already incised all Mesozoic

fl ow from the Aquarius Plateau (Allin, 1993). supercritical CO2; (4) extensional stresses pro- strata and had reached the Redwall duced normal faults and fractures that allowed Limestone. From these interpretations, we con-

Sequence of Events and Their Relationships degassing of CO2, facilitating nucleation and clude that density-driven, convective fl ow within to Landform Evolution growth of ferrous carbonate concretions and the Navajo aquifer started ca. 30 Ma (strata had

cements; (5) thermal expansion during continu- been tilted and CO2 delivery initiated). Precipita- Clarence Dutton’s (1882) “Great Denuda- ing uplift generated numerous, closely spaced tion of ferrous carbonates probably ended across tion” was a long erosive episode during which joints (Fossen, 2010); and (6) dissection of the most of the study area between 15 and 6 Ma Mesozoic strata were stripped northward from Colorado Plateau led to fl ushing of reducing (breach of sealing strata and infl ux of oxygen-

12 Geosphere, June 2015 Concretions, bleaching, and CO2-driven fl ow ated, meteoric water). Ongoing uplift of the Colo- tary canyons (Fig. 12; Biek et al., 2003). Rinded table below the oxidizing concretions (Fig. 12). rado Plateau ensures that cooling will continue to concretions with central cores (including both Rind formation then ceased, and any remain- generate new, closely spaced joints and that mete- small spheroids and large ironstone masses) are ing siderite in these concretions was oxidized oric water will displace reducing waters remain- abundant in Navajo outcrops that are laterally abioti cally in the vadose zone. ing in structurally lower parts of the Navajo, and adjacent to and 30–150 m below the base of the any siderite surviving there will be oxidized. Lava Point fl ow (Figs. 2A, 3A, and 12). “Comet CONCLUSIONS tails” (Fig. 4A) extend southwestward from the Concretion Morphology and small spheroids. These relationships suggest an Bleaching of the upper Navajo Sandstone Timing of Aquifer Drainage explanation for why rind formation goes to com- along the White Cliffs and precipitation of pletion in some concretions but is incomplete in iron-rich concretions of the same formation in The Lava Point basalt fl ow in west-central others (the concretions that retain central cores; the Vermillion Cliffs took place in a vertically Zion National Park (dated at 1.02 ± 0.02 Ma) Fig. 5): Rinds were forming on sideritic concre- confi ned, near-horizontal aquifer in which two followed paleodrainages cut into the Navajo tions at ca. 1 Ma—a time when they lay within fl ow systems interacted. Buoyant fl uids (super- Sandstone (Biek et al., 2003). Due to its resis- oxidizing, southwest-fl owing groundwater, a critical carbon dioxide sourced from Oligo- tance to erosion, this rock now caps mesas that few tens of meters below the water table. Post- cene–Miocene intrusions and small amounts of stand 400 m above the Virgin River and its tribu- lava incision of Russell Gulch lowered the water methane) moved up dip along the sealing strata

Quaternary Figure 12. Map and cross sec- Jc Qu undivided tion showing distribution Qu N (within central Zion National 2 Park; Site RG, Fig. 1) of Late Pleistocene basalt, Navajo Jna Qb basalt flow Sandstone, capping strata of Qb 2 the Carmel Formation, and rinded concretions with cen- 2 tral cores (see Figs. 2 and Jurassic 3A). Basalt flowed down the Jc Carmel Fm paleovalley of North Creek (a tributary of the Virgin River) Jc that was cut into the Navajo Sandstone (Biek et al., 2003). Jurassic Jna Modern canyon fl oors of North Qb Navajo SS Creek, Russell Gulch, and the A Virgin River are ~400 m lower Russell Gulch than the base of the basalt. Tri- Jurassic- angles mark approximate posi- B Kayenta & tions of ancient (upper) and Jkm Moenave modern groundwater tables. Formations Cores of concretions record timing of water-table fall below Jna level of concretions. This halted rinded concretions with cores the cm-scale diffusion of water that is necessary for continued thickening of the rinds. Cores paleoflow were then oxidized abioti- Jkm Left Fork, North Creek (groundwater) cally (Fig. 5). Flow direction of groundwater (red arrows) W113.02° paleoflow was discerned from concretions (basalt) (Fig. 4). Note near parallelism 1.6 km cinder cones of groundwater fl ow direction, N37.02° flow direction of basalt, and fl ow direction of North Creek. Lava Point Flow Jc Modified from Interactive A 1.02 my Qb B Geologic Map of Utah (http:// Jn .utah .gov /map /geomap /interactive /viewer). ~400 m Jn

Geosphere, June 2015 13 Loope and Kettler

of the Carmel Formation and bleached the upper Barge, L.M., Hammond, D.E., Chan, M.A., Potter, S., zircons, paleocurrent trends, and sandstone petro- Petruska, J., and Nealson, K.H., 2011, Precipitation facies: Geosphere, v. 8, no. 4, p. 854–880, doi:10 .1130 Navajo. As CO2 dissolved into the underlying patterns formed by self-organizing processes in porous /GES00763 .1 . water, the added density caused that solution media: Geofl uids, v. 11, p. 124–133, doi:10 .1111 /j Dutton, C.E., 1882, Tertiary History of the Grand Canyon to sink down section, carrying ferrous iron and .1468 -8123 .2010 .00324 .x . with Altas: U.S. Geological Survey, Government Print- Bickle, M.J., 2009, Geological carbon storage: Nature Geo- ing Offi ce, 264 p. methane with it. Upon loss of CO2 along joints, science, v. 2, p. 815–818. Eaton, J.G., Tibert, N.E., and Biek, R.F., 2011, First mam- siderite and ferroan calcite concretions precipi- Beitler, B., Chan, M.A., and Parry, W.T., 2003, Bleaching mals and ostracodes from the Paleogene Claron For- of Jurassic Navajo Sandstone on Colorado Plateau mation, southwestern Utah: Geological Society of tated, sequestering copious carbon in the middle Laramide highs: Evidence of exhumed hydrocarbon America Abstracts with Programs, v. 43, no. 4, p. 77. Navajo. Siderite is now absent from Navajo out- supergiants?: Geology, v. 31, p. 1041–1044. Eichhubl, P., Davatzes, N.C., and Becker, S.P., 2009, Struc- crops, but the ferrous iron and carbon isotopes Beitler, B., Parry, W.T., and Chan, M.A., 2005, Fingerprints tural and diagenetic control of fl uid migration and of fl uid fl ow: Chemical diagenetic history of the Juras- cementation along the Moab Fault, Utah: American in the preserved calcite concretions are indepen- sic Navajo Sandstone, southern Utah, U.S.A: Journal Association of Petroleum Geologists Bulletin, v. 93, dent testimony to the presence of reducing, CO2- of Sedimentary Research, v. 75, p. 547–561, doi: 10 p. 653–681. charged pore water. Post-Miocene uplift of the .2110 /jsr .2005 .045 . Ennis-King, J., and Paterson, L., 2005, Role of convective Biek, R.F., Willis, G.C., Hylland, M.D., and Doelling, mixing in the long-term storage of carbon dioxide in Colorado Plateau and collapse of the Basin and H.H., 2003, Geology of Zion National Park, Utah, in deep, saline aquifers: Society of Petroleum Engineers Range Province led to canyon cutting and drain- Sprinkel, D.A., Chidsey, T.C., and Anderson, P.B., eds., Journal, v. 10, no. 3, p. 349–356. Geology of Utah’s Parks and Monuments: Utah Geo- Fossen, H., 2007, Deformation bands in sandstone: A age of pore water from the Navajo Sandstone. logical Association Publication 28, p. 107–135. review: Journal of the Geological Society, London, Thermal contraction during uplift led to prolifer- Bowers, W.E., 1991, Geologic map of Bryce Canyon v. 164, p. 755–769, doi: 10 .1144 /0016 -76492006 -036 . ation of closely spaced joints that fractured many National Park and vicinity, southwestern Utah: U.S. Fossen, H., 2010, Structural Geology: Cambridge, Cam- Geological Survey Miscellaneous Investigations Series bridge University Press, 463 p. (p. 86–88; 137, 147). of the large siderite concretions and became con- Map I-2108, scale 1:24,000. Garden, I.R., Guscott, S.C., Burley, S.D., Foxford, K.A., duits for fl uid fl ow. When oxygenated meteoric Burnside, N.M., Shipton, Z.K., Dockrill, B., and Elam, Walsh, J.J., and Marshall, J., 2001, An exhumed R.M., 2013, Man-made versus natural CO leakage: A palaeo-hydrocarbon migration fairway in a faulted water reached the siderite concretions, it altered 2 400 k.y. history of an analogue for engineered geologi- carrier system, Entrada Sandstone of SE Utah, USA: them into rinded, iron-oxide concretions. Many cal storage of CO2: Geology, v. 41, p. 471–474. Geofl uids, v. 1, p. 195–213, doi: 10 .1046 /j .1468 -8123 of the large concretions retain central cores with Busigny, V., and Dauphas, N., 2007, Tracing paleofl uid cir- .2001 .00018 .x . culations using iron isotopes: A study of hematite and Gilfi llan, S.M.V., Ballentine, C.J., Holland, G., Blagburn, iron-oxide pseudomorphs after siderite. Cored, concretions from the Navajo Sandstone (Utah, D., Lollar, B.S., Stevens, S., Schoell, M., and Cassidy, pseudomorph-bearing concretions that lie USA): Earth and Planetary Science Letters, v. 254, M., 2008, The noble gas geochemistry of natural CO2 below valley-fi lling, Late Pleistocene lava fl ows p. 272–287, doi: 10 .1016 /j .epsl .2006 .11 .038 . gas reservoirs from the Colorado Plateau and Rocky Chan, M.A., Parry, W.T., and Bowman, J.R., 2000, Diage- Mountain provinces, USA: Geochimica et Cosmo- strongly suggest that much of the siderite was netic hematite and manganese oxides and fault-related chimica Acta, v. 72, p. 1174–1198, doi:10 .1016 /j .gca oxidized very recently (<1 Ma)—a conclusion fl uid fl ow in Jurassic Sandstones of southeastern Utah: .2007 .10 .009 . consistent with some of the recently reported American Association of Petroleum Geologists Bulle- Hamblin, W.K., 1970, Late Cenozoic basalt fl ows of the tin, v. 84, p. 1281–1310. western Grand Canyon, in Hamblin, W.K., and Best, (U-Th)/He ages from Navajo Sandstone concre- Chan, M.A., Bowen, B.B., and Parry, W.T., 2005, Red rock M.G., eds., The Western Grand Canyon District: Utah tions (Reiners et al., 2014). When the concretions and red planet diagenesis: Comparisons of Earth and Geological Society Guidebook to the Geology of Utah, concretions: GSA Today, v. 15, no. 8, doi:10 no. 23, p. 21–38. emerged above the water table, the cm-scale dif- :1130 /1052–5173 (2005) 015 <4 :RRARPD>2.0 .CO;2 . Haszeldine, R.S., Quinn, O., England, G., Wilkinson, M., fusion needed for rind formation was no longer Chan, M.A., Johnson, C.M., Beard, B.L., Bowman, J.R., and Shipton, Z.K., Evans, J.P., Heath, J., Crossey, L., possible. The siderite remaining in concretion Parry, W.T., 2006, Iron isotopes constrain the pathways Ballentine, C.J., and Graham, C.M., 2005, Natural and formation mechanisms of terrestrial oxide concre- geochemical analogues for carbon dioxide storage in cores was therefore oxidized in the vadose zone, tions: A tool for tracing iron cycling on Mars?: Geo- deep geological porous reservoirs, a United Kingdom preserving the pseudomorphs. These relation- sphere, v. 2, p. 324–332, doi: 10 .1130 /GES00051 .1 . perspective: Oil & Gas Science and Technology, v. 60, ships make it likely that some siderite remains Chan, M.A., Ormo, J., Park, M., Stich, M., Souza-Egipsy, V., p. 33–49, doi: 10 .2516 /ogst :2005004 . and Komatsu, G., 2007, Models of iron oxide concre- Heath, J.E., Lachmar, T.E., Evans, J.P., Kolesar, P.T., and in the subsurface and that oxidation continues in tion formation: fi eld, numerical, and laboratory com- Williams, A.P., 2009, Hydrogeochemical characteriza- some areas. Further absolute dating of different parisons: Geofl uids, v. 7, p. 356–368, doi: 10 .1111 /j tion of leaking carbon-dioxide-charged fault zones in portions of these concretions could test the ideas .1468 -8123 .2007 .00187 .x . east-central Utah, with implications for geologic car- Crossey, L.J., Fischer, T.P., Patchett, P.J., Karlstrom, K.E., bon storage, in McPherson, B.J., and Sundquist, E.T., presented here and could reveal uplift rates for a Hilton, D.R., Newell, D.L., Huntoon, P., Reynolds, eds., Carbon Sequestration and Its Role in the Global large portion of the Colorado Plateau. Bleached A.C., and de Leeuw, G.A.M., 2006, Dissected hydro- Carbon Cycle: Geophysical Monograph Series, v. 183, logic system at the Grand Canyon: Interaction between p. 147–158. sandstones, iron-oxide–cemented concretions, deeply derived fluids and plateau aquifer waters Kampman, N., Bickle, M., Wigley, M., and Dubacq, B., and ferroan calcite concretions in other strata in modern springs and travertine: Geology, v. 34, 2014, Fluid fl ow and CO2-fl uid-mineral interactions p. 25–28, doi: 10 .1130 /G22057 .1 . during CO -storage in sedimentary basins: Chemical may reveal fl ow systems with similar confi gura- 2 Crossey, L.J., Karlstrom, K.E., Springer, A.E., Newell, D., Geology, v. 369, p. 22–50, doi:10 .1016 /j .chemgeo tions and histories. Hilton, D.R., and Fischer, T., 2009, Degassing of man- .2013 .11 .012 .

tle-derived CO2 and He from springs in the southern Karlstrom, K.E., Crow, R.S., Peters, L., McIntosh, W., ACKNOWLEDGMENTS Colorado Plateau region—Neotectonic connections Raucci, J., Crossey, L.J., Umhoefer, P., and Dunbar, and implications for groundwater systems: Geological N., 2007, 40Ar/39Ar and fi eld studies of Quaternary We greatly appreciate the helpful reviews by Peter Society of America Bulletin, v. 121, p. 1034–1053, doi: basalts in Grand Canyon and model for carving Grand 10 .1130 /B26394 .1 . Canyon: Quantifying the interaction of river incision Mozley and Laura Crossey. We also thank Jim Elder, Crow, R., Karlstrom, K.E., McIntosh, W., Peters, L., and and normal faulting across the western edge of the Jon Mason, Jodi Norris, and Shirley Yik for help dur- Dunbar, N., 2008, History of Quaternary volcanism Colorado Plateau: Geological Society of America Bul- ing fi eldwork and Andrew Hutsky and Tracy Frank and lava dams in western Grand Canyon based on lidar letin, v. 119, p. 1283–1312, doi:10 .1130 /0016 -7606 for help with isotopes. Kevin Miller and Carolyn analysis, 40Ar/39Ar dating, and fi eld studies: Implica- (2007)119 [1283 :AAFSOQ]2.0 .CO;2 . Shelton (Grand Staircase–Escalante National Monu- tions for fl ow , timing of volcanic events, Karlstrom, K.E., Coblentz, D., Dueker, K., Ouimet, W., ment) and Kezia Nielsen and Jock Whitworth (Zion and lava dams: Geosphere, v. 4, p. 183–206, doi:10 Kirby, E., Van Wijk, J., Schmandt, B., Kelley, S., National Park) provided encouragement and logistical .1130 /GES00133 .1 . Lazear, G., Crossey, L.J., Crow, R., Aslan, A., Darling, assistance . Davis, G.H., 1999, Structural Geology of the Colorado Pla- A., Aster, R., MacCarthy, J., Hansen, S.M., Stachnik, teau Region of Southern Utah: Geological Society of J., Stockli, D.F., Garcia, R.V., Hoffman, M., Keon, America Special Paper 342, 157 p. R., Feldman, J., Helzer, M., and Donahue, M.S., and REFERENCES CITED Dickinson, W.R., Lawton, T.F., Pecha, M., Davis, S.J., the CREST Working , 2012, Mantle-driven Gehrels, G.E., and Young, R.A., 2012, Provenance of dynamic uplift of the Rocky Mountains and Colorado Allin, D.L., 1993, Upper Valley, in Hill, B.G., and Bereskin, the Paleogene Colton Formation () and Plateau and its surface response: Toward a unifi ed S.R., eds., Oil and Gas Fields of Utah: Utah Geological Cretaceous–Paleogene provenances evolution in the hypothesis: Lithosphere, v. 4, p. 3–22, doi: 10 .1130 Association Publication 22, 192 p. Utah foreland: Evidence from U-Pb ages of detrital /L150 .1 .

14 Geosphere, June 2015 Concretions, bleaching, and CO2-driven fl ow

Karlstrom, K.E., Lee, J.P., Kelley, S.A., Crow, R.S., Crossey, MacMinn, C.W., and Juanes, R., 2013, Buoyant currents and mineral resources of the Marysvale volcanic fi eld, L.J., Young, R.A., Lazear, G., Beard, L.S., Ricketts, arrested by convective dissolution: Geophysical southwestern Utah: Cedar City, Utah, Geological Soci- J.W., Fox, M., and Shuster, D.L., 2014, Formation of Research Letters, v. 40, p. 2017–2022, doi:10 .1002 /grl ety of America Field Guide, Rocky Mountain Section the Grand Canyon 5 to 6 million years ago via integra- .50473 . Annual Meeting, p. 131–170. tion of older palaeocanyons: Nature Geoscience, v. 7, Nielsen, G.B., Chan, M.A., and Petersen, E.U., 2009, Dia- Shipton, Z.K., Evans, J.P., Dockrill, B., Heah, J., Williams, p. 239–244, doi: 10 .1038 /ngeo2065 . genetic coloration facies and alteration history of the A., Kirchner, D., and Kolesar, P.T., 2005, Natural leak-

Kettler, R.M., Loope, D.B., and Weber, K.A., 2011, Follow Jurassic Navajo Sandstone, Zion National Park, south- ing CO2-charged systems as analogs for failed geologic the water: Connecting a CO2 reservoir and bleached western Utah: Utah Geological Association, v. 38, storage reservoirs, in Thomas, D.C., and Benson, S.M., sandstone to iron-rich concretions in the Navajo Sand- p. 67–96. eds., Carbon Dioxide Capture for Storage in Deep stone of south-central Utah, USA: REPLY: Geology, Parry, W.T., 2011, Composition, nucleation, and growth of Geologic Formations, Volume 2, p. 699–712. v. 39, no. 11, p. e251–e252, doi: 10 .1130 /G32550Y .1 . iron oxide concretions in the Jurassic Navajo Sand- Smith, E.L., Sanchez, A., Walker, J.D., and Wang, K., 1999, Kocurek, G., and Dott, R.H., Jr., 1983, Jurassic paleogeog- stone, Utah: Sedimentary Geology, v. 233, p. 53–68, Geochemistry of mafi c magmas in the Hurricane vol- raphy and paleoclimate of the central and southern doi: 10 .1016 /j .sedgeo .2010 .10 .009 . canic fi eld, Utah—Implications for small- and large- Rocky Mountains regions, in Reynolds, M.W., and Parry, W.T., Chan, M.A., and Beitler, B., 2004, Chemical scale chemical variability of the lithospheric mantle: Dolly, E.D., eds., Mesozoic Paleogeography of the bleaching indicates episodes of fl uid fl ow in defor- The Journal of Geology, v. 107, p. 433–448, doi: 10 West-Central United States: Rocky Mountain Section, mation bands in sandstones: American Association of .1086 /314355 . SEPM (Society for Sedimentary Geology), p.101–116. Petroleum Geologists Bulletin, v. 88, p. 175–191, doi: Utah Geological Survey, 2014, Interactive Geologic Map of Lindeberg, E., and Wessel-Burg, D., 1997, Vertical con- 10 .1306 /09090303034 . Utah, http://geology .utah .gov /maps /geomap /interactive

vection in an aquifer column under a gas cap of CO2: Parry, W.T., Chan, M.A., and Nash, B.P., 2009, Diagenetic /viewer/index .html. Energy Conversion and Management, v. 38, p. S229– characteristics of the Jurassic Navajo Sandstone in the van der Burg, W.J., 1969, The formation of rattle stones and S234, doi:10 .1016 /S0196 -8904 (96)00274 -9 . Covenant oil fi eld, central Utah thrust belt: American the climatological factors which limited their distribu- Loope, D.B., Kettler, R.M., and Weber, K.A., 2010, Follow Association of Petroleum Geologists Bulletin, v. 93, tion in the Dutch Pleistocene, 1. The formation of rattle

the water: Connecting a CO2 reservoir and bleached p. 1039–1061, doi: 10 .1306 /04270908170 . stones: , Palaeoclimatology, Palaeo- sandstone to iron-rich concretions in the Navajo Sand- Pederson, J., Karlstrom, K., McIntosh, W.C., and Sharp, W., ecology, v. 6, p. 105–124, doi:10 .1016 /0031 -0182 stone of south-central Utah, USA: Geology, v. 38, 2002, Differential incision of Grand Canyon related to (69)90007 -8 . p. 999–1002, doi: 10 .1130 /G31213 .1 . Quaternary faulting—Data from U-series and Ar-Ar Watson, M.N., Zwingmann, N., and Lemon, N.M., 2004, Loope, D.B., Kettler, R.M., and Weber, K.A., 2011, Mor- dating: Geology, v. 30, p. 739–742, doi: 10 .1130 /0091 The Ladbroke Grove-Katnook carbon dioxide natural

phologic clues to the origin of iron-oxide-cemented -7613 (2002)030<0739 :DIOTGC>2 .0 .CO;2 . laboratory: A recent CO2 accumulation in a lithic sand- concretions in the Navajo Sandstone, south-central Potter, S.L., and Chan, M.A., 2011, Iron mass transfer and stone reservoir: Energy, v. 29, p. 1457–1466, doi:10 Utah: The Journal of Geology, v. 119, p. 505–520, doi: fl uid fl ow patterns in Jurassic Navajo Sandstone, south- .1016 /j .energy .2004 .03 .079 . 10 .1086 /661110 . ern Utah, U.S.A: Geofl uids, v. 11, p. 184–198, doi:10 Weber, K.A., Spanbauer, T.A., Wacey, D., Kilburn, M.R., Loope, D.B., Kettler, R.M., Weber, K.A., Hinrichs, N.E., and .1111 /j .1468 -8123 .2011 .00329 .x . Loope, D.B., and Kettler, R.M., 2012, Biosigna- Burgess, D.T., 2012, Rinded iron-oxide concretions: Reiners, P.W., Chan, M.A., and Evenson, N.S., 2014, tures link microorganisms to iron mineralization in a Hallmarks of altered siderite masses of both early and (U-Th)/He geochronology and chemical compositions paleoaquifer: Geology, v. 40, p. 747–750, doi: 10 .1130 late diagenetic origin: Sedimentology, v. 59, p. 1769– of diagenetic cement, concretions, and fracture-fi lling /G33062 .1 . 1781, doi: 10 .1111 /j .1365 -3091 .2012 .01325 .x . oxide minerals in Mesozoic sandstones of the Colo- Wigley, M., Kampman, N., Dubacq, B., and Mickle, M., Lucchitta, I., Holm, R.F., and Lucchitta, B.K., 2011, A Mio- rado Plateau: Geological Society of America Bulletin, 2012, Fluid-mineral reactions and trace metal mobi-

cene river in northern and its implications v. 126, no. 9–10, p. 1363–1385, doi:10 .1130 /B30983 .1 . lization in an exhumed natural CO2 reservoir, Green for the and the Grand Canyon: GSA Rowley, P.D., Cunningham, C.G., Anderson, J.J., Steven, River, Utah: Geology, v. 40, p. 555–558, doi: 10 .1130 Today, v. 21, no. 10, p. 4–10, doi: 10 .1130 /G119A .1 . T.A., Workman, J.B., and Snee, L.W., 2002, Geology /G32946 .1 .

Geosphere, June 2015 15