You are on page 1of 16

Sedimentary Geology 233 (2011) 5368

Contents lists available at ScienceDirect

Sedimentary Geology
j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / s e d g e o

Composition, nucleation, and growth of iron oxide concretions


W.T. Parry
Department of Geology and Geophysics, University of Utah, 115 South 1460 East, Room 383, Frederick A. Sutton Building, Salt Lake City, Utah 84112-0101, United States

a r t i c l e

i n f o

Article history:
Received 4 June 2010
Received in revised form 20 October 2010
Accepted 26 October 2010
Available online 3 November 2010
Editor: B. Jones
Keywords:
Concretions
Goethite
Lepidocrocite
Sandstone diagenesis

a b s t r a c t
Iron oxide concretions are formed from post depositional, paleogroundwater chemical interaction with iron
minerals in porous sedimentary rocks. The concretions record a history of iron mobilization and precipitation
caused by changes in pH, oxidation conditions, and activity of bacteria. Transport limited growth rates may be
used to estimate the duration of uid ow events. The Jurassic Navajo Sandstone, an important hydrocarbon
reservoir and aquifer on the Colorado Plateau, USA, is an ideal stratum to study concretions because it is
widely distributed, well exposed and is the host for a variety of iron oxide concretions.
Many of the concretions are nearly spherical and some consist of a rind of goethite that nearly completely lls
the sandstone porosity and surrounds a central sandstone core. The interior and exterior host-rock sandstones
are similar in detrital minerals, but kaolinite and interstratied illitesmectite are less abundant in the
interior. Lepidocrocite is present as sand-grain rims in the exterior sandstone, but not present in the interior of
the concretions.
Widespread sandstone bleaching resulted from dissolution of early diagenetic hematite grain coatings by
chemically reducing water that gained access to the sandstone through fault conduits. The iron was
transported in solution and precipitated as iron oxide concretions by oxidation and increasing pH. Iron
diffusion and advection growth time models place limits on minimum duration of the diagenetic, uid ow
events that formed the concretions. Concretion rinds 2 mm thick and 25 mm in radius would take place in
2000 years from transport by diffusion and advection and in 3600 years if transport was by diffusion only.
Solid concretions 10 mm in radius would grow in 3800 years by diffusion or 2800 years with diffusion and
advection.
Goethite (-FeO (OH)) and lepidocrocite (-FeO (OH)) nucleated on K-feldspar grains, on illite coatings on
sand grains, and on pore-lling illite, but not on clean quartz grains. Model results show that regions of
detrital K-feldspar in the sandstone that consume H+ more rapidly than diffusion to the reaction site
determine concretion size, and spacing is related to diffusion and advection rates of supply of reactants Fe2+,
O2, and H+.
2010 Elsevier B.V. All rights reserved.

1. Introduction
Diagenetic redistribution and recrystallization of iron minerals in
sandstone aquifers and hydrocarbon reservoirs are records of
paleouid ow paths, timing, and uid composition (Beitler et al.,
2005; Eichhubl et al., 2004). Redistribution and recrystallization leads
to prominent color changes and mobilized iron was precipitated as
concretions. Theoretical solubility and stability calculation together
with the chemical and mineralogical composition of concretions
constrains solution characteristics, precipitation mechanisms, and the
size of spherical concretions. Theoretical calculations of growth times
can be used to estimate duration of uid ow events. Study of
concretions has implications for developing a genetic model of
formation, diagenesis of iron minerals, and hydrocarbon migration.

Tel.: +1 801 277 4124.


E-mail address: wparry@comcast.net.
0037-0738/$ see front matter 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.sedgeo.2010.10.009

The objectives of this study of concretions are to test three


hypotheses. First, the iron oxide concretions initiate or nucleate on a
single seed. This hypothesis will be tested by petrographic analysis to
identify nucleation sites and a conceptual model that accounts for
rinds of iron oxide cemented sandstone that surrounds uncemented
sandstone. Second, the concretions form quickly and record brief
diagenetic events. This hypothesis will be tested by numerical
calculations of concretion growth rates using estimates of temperature, diffusion coefcient, hydraulic conductivity, hydraulic gradient,
and solution composition. Third, concretions grow by diffusion, and
reactant transport limits the growth rate, not the rate of precipitation
or recrystallization of iron oxide. Calculation of mass balance of iron in
the sandstone, examination of the iron distribution in host sandstone,
and comparison of transport limited growth with precipitation and
recrystallization rates of iron oxide will test this hypothesis.
Testing these hypotheses requires selection of sandstone strata
that host concretions, and determination of the chemical, mineralogical, and petrographic characteristics of concretions and surrounding

54

W.T. Parry / Sedimentary Geology 233 (2011) 5368

host rock. Datable materials are notably absent so formation times


must be deduced from numerical modeling. Theoretical solubility and
stability calculations identify precipitation mechanisms.
1.1. The Navajo Sandstone
The Jurassic Navajo Sandstone abundantly displays the effects of
redistribution and recrystallization of iron minerals and is host for a
complex variety of iron and manganese precipitates including many
diverse iron oxide concretions. The Navajo Sandstone is widely
distributed and well exposed on the Colorado Plateau of Utah (Fig. 1)
and is an ideal stratum for a study of concretions. Methods, calculation,
interpretations, and conclusions developed for the Navajo Sandstone
are applicable to iron oxide concretions in other sandstone strata.
The Navajo Sandstone and stratigraphically equivalent Nugget and
Aztec Sandstone, the largest eolian dune deposit in North America,
cover an area greater than 3.5 105 km2 (Blakey, 1994; Blakey et al.,
1988). The Navajo, an important aquifer and hydrocarbon reservoir,
has produced 288 million barrels of oil and 5.1 trillion cubic feet of gas
(Chidsey and Morgan, 2005). The Navajo Sandstone is a well-sorted,
ne-grained quartz arenite. Color varies from moderate reddish
brown and moderate orange pink altered to white or pale orange with
later superimposed diffuse and concretionary iron oxides and
carbonate (Beitler et al., 2005).
Regional studies of Navajo Sandstones and stratigraphically
equivalent Nugget Sandstone describe early diagenetic hematite,
clay, and calcite and dolomite cements. Later burial diagenesis
includes compaction, quartz and feldspar overgrowths, and porelling illite and kaolinite. (Beitler et al., 2005; Bergosh et al., 1982;
Bowen, 2005; Chan et al., 2000; Jordan, 1965; Lindquist, 1983, 1988;
Net, 2003; Parry et al., 2007, 2009; Tillman, 1989).

42

Utah

200 Miles

41

The Navajo Sandstone, normally red from early diagenesis


(Walker, 1967, 1975, 1979), exhibits widespread chemical reduction
fronts formed by migrating reducing aqueous uids (Beitler et al.,
2003, 2005; Bowen et al., 2007; Chan et al., 2000; Garden et al., 2001).
Chemical reduction produces a color change from moderate reddish
brown and moderate orange pink to white or very pale orange along
uid ow pathways (Beitler et al., 2005; Eichhubl et al., 2004). The
color is a function of iron oxide abundance, grain size, and mineralogy.
Iron mobilized by chemical reduction is precipitated at the oxidationreduction front as iron mineral cement in the sandstone. Cemented
sandstone occurs as spheroidal forms from mm to cm scale, rinds,
buttons, disks, and large cylindrical pipes or columns some up to 10 m
high. Strata bound layers of iron oxide cemented sandstone and
conglomerate extend for 10s of meters (Chan et al., 2000, 2004, 2005).
Smaller concretions are more closely spaced than larger concretions
(Chan et al., 2004). Iron oxide also lines northeast striking, vertical
joints. Rind concretions up to 12 cm in diameter typically exhibit a
spheroidal rim b1 mm up to 1 cm thick of iron oxide lling the
sandstone porosity surrounding interior uncemented sandstone. Solid
concretions cemented from center to rim are seldom larger than
1.5 cm in diameter. Numerical simulations of advection combined
with diffusion produces periodic self-organized nucleation centers
through Liesegang-type double-diffusion of iron from interaction of
reduced formation water and oxygen from shallow, fresh water Chan
et al. (2007). Unique spherical concretions in the Navajo Sandstone
are similar in appearance to hematite concretions at Meridiani
Planum on Mars (Chan et al., 2004; Orm et al., 2004). Sefton-Nash
and Catling (2008) give a detailed analysis of the time scale for
formation of the Martian concretions.
In laboratory bench tests, precipitation of hydrated iron hydroxides forms rinds around an initial spherical source of iron (Chan et al.,
2007). Chemical gradients between the inside and outside of
precipitated spheres cause diffusion of Fe towards the outer perimeter
of the sphere forming a rind. The rind then grows inward due to
diffusion within the sphere.
Isotopic composition of Fe in the concretions is consistent with
chemical reduction of early diagenetic Fe3+ mediated by bacteria and
later precipitated by complete oxidation of aqueous Fe2+ (Busigny
and Dauphas, 2007; Chan et al., 2006). However organic compounds
could not be detected by gas chromatographymass spectrometry
analysis (Souza-Egipsy et al., 2006).
2. Stratigraphy

Navajo Sandstone Outcrop

hlands

40

Sevier H
ig

39

S
L

Former

PB
GS C
HF

SF

CV
H

38

K
114

113

112

37
111

110

109

Fig. 1. Map of Utah's portion of the Colorado Plateau showing areas of Navajo Sandstone
outcrop, major normal faults, and structural uplifts. A = Abajo Mountains, B = Boulder
Mountain, C = Circle Cliffs Uplift, CV = Caineville, H = Henry Mountains, HF = Hurricane fault K = Kaibab Uplift, L = LaSal Mountains M = Monument Uplift, P = Paunsaugunt fault; S = San Rafael Uplift; SF = Sevier Fault, GS = Sample location in the Grand
Staircase Escalante National Monument, U = Uncompahgre Uplift.

The Navajo Sandstone is overlain by 1.8 to 4.2 km of marine,


uvial, eolian, and lacustrine sediments (Fig. 2) (Doelling et al., 2000;
Hintze and Kowallis, 2009). The Carmel Formation, a sabkha sequence
of sandstone, siltstone, mudstone, limestone, anhydrite, and gypsum,
above the Navajo Sandstone forms a seal. The uvial Chinle Formation
and the uvial to marine Moenkopi Formation that lie beneath the
Wingate Formation are ne-grained aquitards. The aquitards above
and below the Navajo Sandstone inhibit uid access except along
faults or surface exposures.
3. Structure
The study area is located in the Colorado Plateau province, a
roughly circular area of about 384,000 km2 with many high plateaus
and isolated mountains. Elevations range from 900 to 4200 m and
average 1600 m. Monoclinal folds, anticlines, domes, basins and faults
deform the generally at lying Navajo Sandstone (Hunt, 1956). The
structural framework of the Colorado Plateau and the study area is
dominated by Laramide deformation (Dickinson and Snyder, 1978;
Dickinson et al., 1988) that occurred from uppermost Cretaceous to
Middle Eocene and later Basin and Range normal faults (Davis, 1999;
Hunt, 1956). Laramide compression formed the Kaibab, Circle Cliffs,

W.T. Parry / Sedimentary Geology 233 (2011) 5368

Tert iary

Age

Thickness (m)
Claron Fm

Pine Hollow, Grand Castle


Canaan Peak Fms

Cret aceous

55

427

0-396

Lacustrine limestone and dolomite

Fluvial conglomerate, sandstone, and


mudstone

Kaiparowits Fm

610-914

Lacustrine sandstone and siltstone

Wahweap Fm

305-457

Fluvial and lacustrine mudstone, siltstone,


sandstone, and conglomerate

274-549

152-229
Tropic Shale
Dakota and Cedar Mtn Fms 1-313

Permian T r i a s s i c

Jurassic

Morrison Fm
Entrada Ss.

0-290
0-305

Carmel-Page Fms

55-317

Navajo Sandstone

396-457

Kayenta Fm
Wingate Sandstone

46-107
30-107

Marine mudstone and shale


Fluvial conglomerate, sandstone, siltstone,
mudstone. Lacustrine limestone
Eolian sandstone
Marine limestone, sandstone
siltstone, gypsum, and eolian sandstone

Eolian sandstone

Fluvial sandstone, siltstone


Eolian sandstone

Chinle Fm

130-283

Fluvial conglomerate, sandstone,


claystone, mudstone, bentonite

Moenkopi Fm

134-351

Fluvial, eolian, and coastal plain sandstone,


siltstone, mudstone, limestone, gypsum

Kaibab Limestone

Marine dolomite

White Rim Sandstone

Eolian sandstone

Fig. 2. Stratigraphy of the Grand Staircase-Escalante National Monument study area. Strata and thicknesses from Doelling et al. (2000).

Escalante, Monument, and San Rafael monoclines (Davis, 1999; Hunt,


1956). The Escalante monocline, with a steeply dipping west limb and
gentle east limb, lies to the west of the study area and the Circle Cliffs
Uplift, with a steeply dipping east limb, lies to the east (Fig. 1). The
Escalante River and its tributaries have cut deep canyons into the west
limb of the Circle Cliffs Uplift. The study area is located on the gentle
west limb of the Circle Cliffs Uplift (Fig. 1).
Basin and Range extensional faulting began about 15 Ma and
continues today (Davis, 1999). Three major Basin and Range faults are,
from west to east, the Hurricane fault, the Sevier fault and the
Paunsaugunt fault (Fig. 1). The Paunsaugunt fault is the easternmost
of the Basin and Range faults and is nearest the study area, 55 km to
the east (Fig. 1). Boulder Mountain, a portion of the Aquarius Plateau
protected from erosion by a volcanic cap lies north of the study area.

Uplift of the Colorado Plateau region began in the middle Eocene and
lasted until the Late Miocene elevating and tilting the area and
subjecting it to erosion, which continues today (Hunt, 1956). Mean
uplift of the Colorado Plateau near the study area since the Late
Cretaceous is estimated at near 2000 m (Pederson et al., 2002; Sahagian
et al., 2002). The present physiography of mesas, plateaus and deeply
incised canyons is the result of nearly continuous erosion throughout
Cenozoic time that has removed thousands of meters of sedimentary
rock and developed the present groundwater ow system.
4. Chemical, mineralogical, and petrographic methods
The location of the study is an area of abundant concretions that
reach 5 cm in diameter located in the Spencer Flat area of the Grand

56

W.T. Parry / Sedimentary Geology 233 (2011) 5368

Staircase Escalante National Monument (Fig. 1). The sample site is in


the Jurassic Navajo Sandstone on the gently dipping west limb of the
Circle Cliffs Uplift. The Navajo Sandstone in the study area contains
numerous iron oxide concretions some of which weather out and
form an extensive lag deposit (Fig. 3A). The concretions consist of a
rind of iron oxide cemented sandstone surrounding interior, uncemented sandstone (Fig. 3B). Two large samples approximately
50 20 20 cm each that included a rind concretion and surrounding
host rock (Fig. 3C, E) were subdivided in the laboratory to examine
characteristics of the concretion interior, concretion rind, and host
rock. A series of sub-samples were collected systematically 2 to 4 cm
apart from the concretion center to 28 cm from the center in sample
GS1a (Fig. 3D) and concretion center to 13 cm from the center in
sample GS1b (Fig. 3C).
Mineralogy and textures were determined by petrographic
examination of thin sections, petrographic point counting, and X-ray
diffraction. Thin sections were stained for potassium feldspar (Bailey
and Stevens, 1960). Thin sections were examined with a standard
petrographic microscope. Modal mineralogy was determined by a
combination of petrographic point counting (1000 points per thin
section, 300 m step) and least squares t of mineralogy, mineral
chemistry, and whole-rock chemistry using computer software (Parry
et al., 1980). The computer program incorporates whole-rock
chemistry and petrographic point counting to produce mineral
abundance estimates that are consistent with both petrography and
chemistry.
Whole-rock samples were analyzed for major elements by
inductively coupled plasma atomic emission spectroscopy. Analytical
accuracy and laboratory procedures conform to international standards ISO 9001:2000 and ISO 17025:2005. Analyses are accurate to
the signicant gures reported (Table 1).

Clay minerals were extracted from core samples by mild grinding


by hand and 2 micron particle size was separated by centrifugation.
Oriented smears on glass slides were X-rayed with a Rigaku 2000 V Xray spectrometer using a copper anode X-ray tube. Scanning speed
was 2 per minute. X-ray scans were done following air-drying and
vapor glycolation at 60 C for 12 h.
5. Chemical, mineralogical, and petrographic results
5.1. Chemical composition
Chemical analyses of sampling points (Fig. 3E) reveal that the
interior sample contains slightly less Al2O3 and K2O than the exterior
samples (Table 1, Fig. 4). Samples from GS1a contain 0.03 to 0.08% C,
but no carbonate minerals were detected in petrographic or X-ray
diffraction examination (Table 1). Only the interior uncemented
sandstone and one exterior host sandstone sample contained
detectable S. Phosphorous was detected in all exterior samples at
concentrations of 0.01 to 0.07% (Table 1). Fe2O3 varies from 1.47
inside the concretion to 28.3 wt.% in the concretion rind to 1.70 wt.%
outside the concretion (Table 1, Fig. 4). No iron-depleted zone is
present outside the concretion within 28 cm of the concretion or more
than half the distance to the next concretion.
5.2. Petrography and mineralogy
Sample GS1a averages 89% rounded quartz grains and 5.7% Kfeldspar grains (Table 1). Quartz grains reach 300500 m in diameter
and in some cases quartz overgrowths are present. The rind
concretions consist of a thin rind of goethite-cemented sandstone
that is 19.6 vol.% goethite (Table 1) surrounding an uncemented

Fig. 3. Photographs of the sample locality showing the location of detailed sampling. A. Lag deposit of ~ 5 cm diameter concretions that have weathered from the Navajo Sandstone. B.
Examples of concretion cross-sections of various sizes. C. Sample GS1b at the location of collection. D. Sample GS1a at the collection location. Detailed sampling points are shown. E.
Sample GS1a in the laboratory following collection of detailed samples from points shown.

W.T. Parry / Sedimentary Geology 233 (2011) 5368

57

Table 1
Chemical composition and modal mineralogy of GS1a samples.
Sample

rind*

10

SiO2
Al2O3
Fe2O3
CaO
MgO
Na2O
K2O
C
S
P2O5

94.7
1.96
1.47
0.07
0.07
0.12
1.14
0.05
0.02
b 0.01

63.5
1.93
28.3
0.06
0.08
0.16
0.83

94.6
2.11
1.23
0.04
0.07
0.07
1.20
0.03

94.0
2.32
1.51
0.05
0.07
0.07
1.32
0.03

93.8
2.12
1.46
0.04
0.07
0.05
1.22
0.03
0.01
0.01

92.8
2.11
1.04
0.09
0.06
0.06
1.20
0.08

94.4
2.14
1.44
0.03
0.07
0.05
1.22
0.08

94.3
2.20
1.17
0.03
0.07
0.05
1.24
0.03

94.5
2.25
1.30
0.03
0.08
0.05
1.26
0.03

94.2
2.15
1.40
0.03
0.07
0.07
1.22
0.03

94.0
2.33
1.70
0.03
0.08
0.05
1.30
0.05

0.02

94.2
2.14
1.55
0.05
0.07
0.06
1.22
0.04
0.01
0.02

0.04

0.01

0.02

0.01

0.07

0.02

74.5
0.3
4.6
0.6
2.9
0.03
0.4
16.8

71.7
0.1
5.3
0.7
1.7
0.6
0.5
19.3

76.0
0.5
6.2
0.7
1.7
0.4
0.9
14.1

75.5
0.5
5.2
0.7
1.4
0.6
1.2
14.9

74.9
0.5
5.4
0.5
1.4
0.3
1.4
15.6

Modal Mineralogy (Volume %)


Quartz grns
75.5
Overgrowth
0.1
Kfeldspar
5.1
Goethite
0.7
Illite rims
1.7
Pore illite
0.4
Kaolinite
0.1
Porosity
16.4

69.4
6.6
19.1

3.2
1.8

0.01

74.9
0.5
5.2
0.7
1.6
0.6
1.0
15.6

73.3
0
5.3
0.6
2.0
0.3
1.0
17.6

73.5
0.9
6.1
0.6
1.6
0.3
1.2
15.9

74.1
0.5
5.7
0.7
1.0
0.5
1.2
16.4

71.5
0.3
6.4
0.8
1.3
0.4
0.8
18.5

*rind thickness of 0.7 mm prevented large enough sample for chemical analysis. This analysis is a thicker rind from Bowen (2005).

center. The rind porosity is almost completely lled with precipitated


goethite (Fig. 6A, B). The higher magnication photo of Fig. 6B shows
that little porosity remains open. Samples average 1.4% FeO (OH).
Goethite in the concretion rim occurs in three textures: a
microgranular texture with crystal size from less than 1 m up to
5 m that forms a coating on quartz grains; an equigranular texture
with grain size up to 15 m occurs in the centers of former pore space;
and a radial-brous texture with sweeping extinction 20 to 40 m in
size that forms on quartz and feldspar grains. Porosity of the goethite
rim is estimated to be 1.8%.
Clay minerals consist of illite and kaolinite (Fig. 5A, B). Illite in thin
section occurs as a pore lling and as rims on quartz grains. Illite rims are
absent from the quartz grains in the rind or masked by semi opaque
goethite. Illite rims on quartz grains are present adjacent to kaolinite
lled porosity. Scattered red hematite grains less than 0.01 m are
present within illite coating. Clay mineral X-ray diffraction traces reveal
two ordered interstratications of illite with smectite. The rst (RN 3, 3
or more illite layers together before a smectite layer is encountered)
contains 710% interstratied smectite and is present in both GS1a and
GS1b. The second (R1) contains 70% illite and 30% smectite and is
present in the exterior of the GS1b concretions along with the R N 3 illite
(Fig. 5A, B). The concretion rind contains slightly more kaolinite than
3.5

Weight or Volume %

K2O

28.3 %

Al2O3

rind

Fe2O3
Kaolinite

2.5
2
1.5

6. Numerical growth time models


Chemical and mineralogical study did not detect dateable materials in
the concretions. The rate of growth of the spherical concretions can be
calculated by numerical growth models that constrain the minimum
duration of diagenetic uid events involved in formation. Goethite lling
the porosity in the concretions forms from solution by a combination of
diffusion and advective transport of reactants through the porous
sandstone, concurrent precipitation reactions, and recrystallization of
the precipitate to goethite. Growth time calculations require estimates of
the temperature of transport and precipitation, identity, stability,
solubility, and precipitation mechanism of the iron oxide forming the
concretionary cemented sandstone, hydrologic conductivity of the
sandstone, hydraulic gradient driving advection, advective velocity, and
concentration of iron in solution in the groundwater.
6.1. Temperature

1
0.5
0

interior and exterior sandstones and no illite (Table 1, Fig. 4). Samples
average 2.1% illite and 1.4% kaolinite (Table 1). Some K-feldspar grains
are altered to kaolinite and illite.
Modal mineralogy of sequential samples is remarkably uniform.
However, the sample from the concretion interior contains less
kaolinite and slightly less illite than the exterior samples and more Kfeldspar (Table 1, Fig. 4). Exterior host sandstone samples commonly
show a light tan iron oxide coating on sand grains. X-ray diffraction
analysis of these samples identied lepidocrocite (-FeO (OH))
(Fig. 5A, B). Illite grain coatings are present inside the tan iron oxide
coating. Also the concretion interior sandstone sample is free of
lepidocrocite that is common in exterior samples.
Iron oxide precipitate clearly nucleates on K-feldspar in the host
sandstone (Fig. 6E, F). In some cases the iron oxide precipitate
completely surrounds K-feldspar grains (Fig. 6F).

10

15

20

25

30

Distance from Concretion Center (cm)


Fig. 4. Plot of chemical constituents and minerals as a function of distance from a
concretion center.

Diffusion and chemical reactions are temperature dependent


processes. Few constraints on temperature of formation of the
concretions are available. Goethite in 10's of micrometer particles is
stable relative to hematite up to 60 C (Navrotsky et al., 2008). Burial
histories for the region suggest depth of burial of the Navajo
Sandstone did not exceed 1.8 to 4.2 km (Fig. 2) limiting the maximum
temperature to 60 to 100 C (Garden et al., 2001; Huntoon et al., 1999;
Nuccio and Condon, 2000). Hematite is the main iron oxide
precipitate in the Navajo Sandstone in the Covenant oil eld where

Quartz Illite 95/Smectite


K-feldspar

Kaolinite

Lepidocrocite

Kaolinite

Illite 95/Smectite

W.T. Parry / Sedimentary Geology 233 (2011) 5368

Illite 95/Smectite

58

250 Counts
GS1a10
GS1a9

Intensity

GS1a8
GS1a7
GS1a6
GS1a5
GS1a4

GS1a3
GS1a2
GS1a1
GS1a0

10

15

20

30

25

K-feldspar

Quartz Illite 95/Smectite

Kaolinite

Quartz

Illite 95/Smectite

Illite 70/Smectite

Lepidocrocite

Kaolinite

Illite 95/Smectite

Illite 70/Smectite

2 (CuK)

1250

Intensity(Counts)

1000

GS1b6

GS1b5

750
GS1b4
GS1b3

500
GS1b2

250

GS1b1

GS1b0

0
5

10

15

20

25

30

2 (CuK)
Fig. 5. X-ray diffraction patterns of 2 clay extracts. A. X-ray diffraction patterns of -2 extracts of sample GS1a. B. X-ray diffraction patterns of 2 extracts of sample GS1b.

the temperature is 87 C (Parry et al., 2009). Mobilization of iron is


associated with hydrocarbon migration that took place in Late
Cretaceous to Early Tertiary and precipitation of iron oxide concretions must follow iron reduction and mobilization. Temperature in the

Navajo Sandstone cooled to 50 C, a temperature at which goethite


could precipitate, during exhumation in the late Tertiary (Huntoon
et al., 1999). The stability diagram and diffusion models have been
calculated using a temperature of 50 C.

W.T. Parry / Sedimentary Geology 233 (2011) 5368

59

Fig. 6. Photomicrographs of thin sections. A. Goethite cemented rim on sample GS1a. B. Goethite cement in concretion rim. C. Goethite nucleating on K-feldspar (yellow) in sample
GS1a-1 outside the rim. D. Goethite nucleating on K-feldspar in sample GS1a-4 outside the rim. E. Goethite nucleating on K-feldspar in sample GS1a-4 outside the rim. F. Goethite
protective layer surrounding K-feldspar in sample GS1a-1 outside the rim.

6.2. Solubility and precipitation mechanisms


A genetic model of iron oxide concretion formation should include
stability, solubility, and precipitation mechanism. Goethite is stable
with respect to magnetite at higher oxidation and lower pH (Fig. 7).
Pyrite is stable at lower oxidation where sulfur is reduced as well as
iron. Goethite is in equilibrium with aqueous Fe2+ species at higher

pH values as log fO2 decreases. Lepidocrocite is metastable with


respect to goethite (dashed line on Fig. 7).
Solubilities of goethite and lepidocrocite were calculated using a
moderately saline water composition (ionic strength 0.18) from the
Navajo Sandstone aquifer in southeastern Utah (Spangler et al., 1996).
Activity coefcients for aqueous species were calculated with
Geochemists Workbench (Bethke, 1998). Stability constants, free

60

W.T. Parry / Sedimentary Geology 233 (2011) 5368

1000000

100000

FeSO 4

20
et

Goethite

as

ta

bl

Le

pi

do

40

cr

oc

ite

0.1
0.01
0.001

Pyrite

Pyrrhotite
50C
3

10

11

12

Fig. 7. Log fO2-pH diagram showing the stability of goethite relative to pyrite, magnetite,
siderite, and solution. A dashed line shows metastable lepidocrocite. The diagram was
constructed using Geochemists Workbench (Bethke, 1998). The diagram was
constructed using log aFe = 6 (about 0.05 mg/l), log aSO4 = 3 (about 100 mg/l).

energies, activity coefcients and equilibrium constants used in the


solubility calculations are given in Table 2.
Goethite has a solubility minimum that increases in pH from 7 to 9
as log fO2 decreases from 20 to 50 (Fig. 8). The solubility of
lepidocrocite is more than 10 times greater than goethite at all pH
values (Fig. 8). The solubility minimum for lepidocrocite at log fO2 =
20 is at pH = 7 to 7.5. No change in solubility above log fO2 = 20
is evident in the calculations as the solution is fully oxidized in the pH

Table 2
Parameters used in model calculations.
Reaction or species

Stability Constants
Log1
Fe3+ + H2O = Fe(OH)2+ + H+
+
Log 2
Fe3+ + 2H2O = Fe(OH)+
2 + 2H
Log 3
Fe3+ + H2O = Fe(OH)3 + 3H+
Log 4
Fe3+ + H2O = Fe(OH)-4 + 4H+

10

12

pH

pH

Parameter

Magnetite

70

80
2

10

20

60

=fO2

Fe

50

++

100

Lepidocrocite
log fO =-20

g
lo

log fO2(g)

30

1000

40
=fO2
log

10000

it e
eth -50
Go f =
O2
log

Solubility (nanograms Fe/kg)

10

Value (50 C)

Source

1.57
5.73
12.95
20.26

1
1
1
1

Fig. 8. Calculated solubility of goethite and lepidocrocite. Both minerals have a


solubility minimum that shifts with shifting fO2. Parameters used in the calculations are
given in Table 2.

range from 4 to 12. Precipitation of goethite can be initiated by an


increase in pH or an increase in fO2 or a combination (Figs.7 and 8).
Iron species in solution calculated as a function of pH at log fO2 =
20 illustrate Fe++ species are important at low pH and lower values
of log fO2 (Fig. 9). At pH = 4 the dominant species are Fe++ (52%), Fe
o
(OH)++ (37%), then Fe (OH) +
2 (10%), followed by Fe (OH) 3 (1%), and
at pH above about 7.5 Fe (OH) 4 becomes the dominant species.
Aqueous complexes of Fe with Cl, SO4 and HCO3 have negligible
concentrations except at low pH values.
6.3. Hydraulic conductivity and hydraulic gradient
The product of hydraulic gradient and hydraulic conductivity
yields the advective water velocity needed in the growth time models.
Paleo groundwater and hydrocarbon ow is assumed to be in the
direction of the paleosurface (Sanford, 1995). Topographically driven
groundwater ow affecting the study area in Late Cretaceous was
northeast toward the study area from the Sevier highlands approximately 300 km west of the study area (Sanford, 1995) (Fig. 1). During
this time, the Navajo Sandstone was near sea level and the Sevier

100
0.1268
0.3344
0.3006
1.000
0.7308
0.7334
1.000

2
2
2
2
2
2
2

Gibbs Free Energy


Gof
Fe2+
Gof
Fe3+
Gof
O2(g)
Gof
H+
o
G f
FeO(OH) Goethite
Gof
Fe)(OH) Lepidocrocite

21,231 cal
2,431
1,233
0
117,384
115,098

3
3
3
3
4
4

Log Equilibrium Constants


Goethite
FeO(OH) + 2H+ = Fe2++ 1/4O2 + 3/2H2O
Lepidocrocite FeO(OH) + 2H+ = Fe2++ 1/4O2 + 3/2H2O
Log Keq
Fe3+ + 1/2H2O = Fe2+ + 1/4O2 + H+

6.873
5.327
6.395

5
5
5

Fe(OH)4-

80
70
60
50
40

Fe++

30
20

Fe(OH)3

++
10 Fe(OH)

1. Liu and Millero (1999); 2. Geochemist's Workbench; 3. SUPCRT92 Johnson et al.


(1992); 4. Majzlan et al. (2003a,b); 5 Calculated.

Fe(OH)2+

90

Proportion of Species

Activity Coefcients

Fe3+

Fe2+

Fe(OH)2+

Fe(OH)3

Fe(OH)+
2 , Fe(OH)4, Fe(OH)3,

Fe(OH)+

Fe(OH)2

10

12

pH
Fig. 9. Fe speciation as a function of pH at log fO2 = 20. At elevated pH the complex Fe
(OH)-4 increases solubility from the minimum value. Solubility increases substantially as
pH is lowered below the solubility minimum.

W.T. Parry / Sedimentary Geology 233 (2011) 5368

highlands were as high as 3.2 km (DeCelles and Coogan, 2006)


resulting in a topographic gradient of 0.01. From Middle Eocene to
Late Miocene, groundwater ow was westward from the crest of the
Circle Cliffs Uplift. From late Miocene to the present, ow was
generally southwestward from the Boulder Mountain portion of the
Aquarius Plateau (Sanford, 1995).
The hydraulic gradient from the Circle Cliffs Uplift can be
estimated by restoring the strata above the exposed Permian bed in
the core of the uplift. The present crest of the uplift is 197 m above the
study area. Restoring the strata from Permian to the Jurassic Navajo
Sandstone (Fig. 2) indicates that the Navajo Sandstone in the restored
crest of the uplift would have been 1044 m above and 34 km
northeast of the sample location for a hydraulic gradient of 0.03.
The Boulder Mountain portion of the Aquarius Plateau lies 42.5 km
north and 1543 m above the sample location. Abundant lakes and
springs from 3358 m elevation down to 1879 m suggest the present
hydraulic gradient is near the topographic gradient of 0.036.
Numerous studies of the important Navajo Sandstone regional
aquifer include measurements of hydraulic conductivity. The measurements nearest the study area at Caineville (Fig. 1) show that the
Navajo Sandstone is hydraulically isotropic with an average hydraulic
conductivity of 0.15 m/d (54.8 m/yr) (Hood and Danielson, 1979). The
product of hydraulic conductivity of 54.8 m/yr and the hydraulic
gradient of 0.036 yields the ow velocity of 2 m/yr.
6.4. Groundwater composition
The iron concentration in groundwater must be estimated to
calculate growth times. Iron concentrations ranges from 1.8 to 5 mg/
l in the region 100 km northeast of the study area (Hood and
Danielson, 1981) and 1.5 to 19 mg/l in the Paradox Basin area 200 km
east of the study area (Spangler et al., 1996). Water produced from the
Covenant oil eld Navajo Sandstone reservoir contains 14 mg/
l dissolved Fe (Chidsey et al., 2007). The following calculations use a
value of 5 mg/l, the maximum observed nearest the study area by
Hood and Danielson (1981).
6.5. Transport controlled concretion growth time
Numerical growth time models can be used to deduce the length of
time required for formation of concretions and the duration of the
uid ow events that form them. Concretions to be modeled are solid
concretions up to 1.5 cm in diameter and rind concretions with a thin
rind of precipitated matter surrounding a spheroidal core containing
little of the precipitate. Mass of Fe used in calculations of growth times
for solid concretions assume that concretion mineral matter cements
the sandstone from center to edge, but no assumption of a nucleus is
made. To calculate the time for growth of the rind, the time for growth
of a spherical concretion of the radius of the interior is rst calculated,
and then the time for growth of the radius of the exterior of the rind is
calculated. The difference between the two growth times is the time
required to grow the rind. The growth time for concretions was
calculated using a temperature of 50 C, 5 mg/l dissolved Fe, diffusion
only as a minimum no ow condition, and a ow velocity of 2 m/yr
using the following 5 equations.
Growth time models described below include formation of solid
concretions and rind concentrations formed by diffusion and by a
combination of diffusion and advection of reactants. The growth times
not only indicate the length of time required for growth of the
concretions, but also place limits on the minimum duration of the
uid ow events that led to the formation of the concretionary iron
oxide cements.
Eq. (1) was developed for Fe supplied to the concretion by
diffusion only (Berner, 1968, Lasaga, 1998) assuming the volume
fraction of precipitated iron oxide in the concretion, Fp, equals the
porosity . Terms and values used in the calculations are dened in

61

Table 3
Variables in calculation of concretion growth times in Eqs. (1)(5) (note all variables
converted to meters, years, moles).
R = radius of concretion
D = diffusion coefcient. Here we follow Berner (1980) and Drever (1997) to
correct for tortuosity. The effect of tortuosity on the diffusion coefcient is
approximated by the effect on electrical conductivity which is measured by
Archie's Law and the formation factor F. The diffusion coefcient for a porous
sediment, Dsed = Dsolution/F, where F is the formation factor. The formation
factor F = 2. Further, the effective diffusion coefcient is Dsed/ so
Deffective = Dsolution. Signicant Fe is present in solution only as Fe2+ species.
Precipitation is initiated by either an increase in fO2 or pH. The diffusion
coefcient for Fe2+ at 50 C in aqueous solution is 1.5 10 5 cm2/sec (Oelkers
and Helgeson, 1988). Correcting this for tortuosity and converting to
appropriate units of m2/year produces Deffective = 7 10-3 m2/year.
Fc = Volume fraction of the host rock occupied by concretions. Concretions 5 cm in
diameter are spaced approximately 40 cm apart. Therefore, each concretion
occupies 65.45 cm3 in a volume of rock 4188.8 cm3 or a volume fraction of
0.0156.
Fd = Volume fraction of the host rock occupied by dissolving source material.
Average hematite concentration in original red Navajo Sandstone is 0.5
weight %. For a sandstone that is 89% quartz, 5.7% K-feldspar, 2.1% illite, 1.4%
kaolinite and 15% porosity, the volume fraction hematite source material is
0.002. The average sandstone in sample GS1a is 1.3% Fe2O3, however much of
this is introduced and is present as lepidocrocite.
Fp = Volume fraction of the concretion occupied by precipitated material.
Precipitated goethite occupies all of the porosity. Fp then = 0.15 (see below
for porosity)
= Volume fraction of the host rock occupied by pore space (porosity). Porosity of
undeformed Navajo Sandstone ranges from 10 to 30% (Cooley et al., 1969;
Cordova, 1978; Hood and Patterson, 1984; Freethey and Cordy, 1991; Shipton
et al., 2002). Average porosity of dune facies Navajo Sandstone in the Covenant
oil eld, central Utah is 12% (Chidsey et al., 2007). We have chosen a nominal
15% porosity for these calculations.
= Molar volume of precipitate (goethite) = 2.08 10 5 m3/mole
c cR = Concentration of Fe2+ far away from a concretion - concentration at the
concretion. Concentration of Fe2+ in oil eld reservoir produced water is
12 mg/l at the Covenant oil eld (Chidsey et al., 2007), and concentration
in the model calculations of Parry et al. (2004) where dissolved CH4 is
reacted with Navajo Sandstone reaches a maximum of 14 mg/kg. We
choose a conservative 5 mg/kg for these calculations = .0895 moles/m3.
cL = concentration at the outer edge of the depleted zone.
U = Pore uid velocity. The average hydraulic conductivity of 58.4 m/yr and
hydraulic gradient of 0.03 were used to calculate a darcian velocity of 2 m/yr as
discussed in the text.

Table 3. Derivation of the equation assumes a linear concentration


gradient of less than 0.05 mole/l over some large distance.
t=

R2
2vDc cR

Dissolving original red hematite grain coating near the concretion


could have formed a concretion and an iron-depleted zone. The
equation for this is (Berner, 1980):
R
t=

2 Fp
Fd


1

2vDcL cR

As the volume fraction of dissolving material, Fd, becomes smaller


the time necessary for concretion growth becomes larger. For the
values of Fp = 0.15 and Fd = 0.002 in the Navajo Sandstone host rock,
the concretion growth times estimated from equation 2 are greater
than equation 1 by a factor of 74. However, chemical and
mineralogical analysis of host rock near a concretion shows no
depleted zone (Table 1, Fig. 4).
Eq. (3) accounts for the supply of iron by both diffusion and
owing groundwater (Berner, 1968, 1980).



0:715
D
R 0:715U
+
1 + RU
D
t=
1:715Uvc cR

D
0:715U

W.T. Parry / Sedimentary Geology 233 (2011) 5368

The model for simple diffusion (Berner, 1968) was modied to


account for a steady state concentration gradient that is not linear. The
volume fraction of precipitated material within the concretion, Fp,
and the volume fraction of dissolving material, Fd, within the host
rock were also introduced (Wilkinson, 1991; Wilkinson and Dampier,
1990).
0
r
t=

F
1 d
Fp

2@

!1 1
3
A
4

2vDcL cR

7000
6000

Equation 4
Equation 1

5000
4000
3000

Equations 3 and 5

2000
1000

Pore-uid ow and nonlinear steady state concentration gradient


were introduced in Eq. (5) (Wilkinson and Dampier, 1990).

t=

9000
8000

Gowth Time (years)

62

h

D
R 0:715U
1+


RU 0:715
D

D
0:715U

i
1
1 FFd 3

1:715Uvc cR

Rind Thickness (mm)

5000

6.6. Oxidation rate of Fe2+


The calculated growth times assume transport controls the rate of
growth of the concretion. If the rate of growth is instead controlled by
chemical reaction, then these growth times are minimum values. The
chemical reactions that formed the concretions are oxidation rate of
Fe2+ and the rate of conversion of precipitated ferric hydroxide to
goethite. Above pH = 4 oxidation rate is related to the overall reaction
(Langmuir, 1997)
2+

+ :2502 + 2:5H2 O = FeOH3 + 2H

Fe

The empirical rate law for the reaction is (Langmuir, 1997):




d Fe2 +
dt

= k



Fe2 +
H 2

PO2

40000

Growth Time (years)

35000
n

tio

30000

a
qu

io

t
ua

25000

Eq

20000

nd

15000

ion
uat

3a

Eq

10000
5000
0

10

15

20

25

30

Concretion Radius (mm)


Fig. 10. Growth time for FeO(OH) concretions calculated for diffusion and diffusion
coupled with pore water ow.

Growth Time (years)

Rind Thickness (mm)

The results of the calculations for each equation are shown in


Figs. 10 and 11. Eq. (1) and (4) produce nearly identical results for
diffusion-only growth, and Eq. (3) and (5) produce nearly identical
results for concretion growth by diffusion and advection. For example,
calculations suggest that a solid concretion with a radius of 25 mm
grows by diffusion in about 24,000 years (Fig. 10). A 25 mm radius
concretion would form in 15,000 years by diffusion and advection.
Growth time is 3,600 years for a 2 mm thick rind with a radius of
25 mm formed by diffusion (Fig. 11A). The two models involving both
diffusion and uid ow predict the 2 mm thick rind formed in slightly
less than 2,000 years. Concretions 25 mm in radius form rinds 1 mm
thick in 1,000 years to 5,000 years for a 5 mm thick rind when
advection accompanies diffusion (Fig. 11B).

4000

5
4

3000

3
2000

1000

0
5

10

15

20

25

Concretion Radius (mm)


Fig. 11. A. Growth times for rinds of various thickness and 5 cm diameter calculated for
diffusion and diffusion coupled with pore water ow. B. Growth time for FeO(OH)
concretion rinds of various thicknesses and diameters calculated for diffusion coupled
with pore water ow.

The rate of oxidation of ferrous ion is greater at high pH and PO2.


The rate constant, k+, is 1.2 10 11 mol2/(bar day) at 20 C and the
activation energy is estimated to be 23 kcal/mol (Langmuir, 1997).
The rate constant at 50 C is therefore estimated to be 4.71 10 10
mol2/(bar day). If pH and PO2 are specied the rate law becomes
pseudo rst order. With pH = 6 the pseudo rst order rate constant,
k', is 94/day for PO2 = 0.2 and 0.00471 for PO2 = 1 10 5 yielding
half-lives of .007 days and 147 days respectively. Bacterial mediation
of the oxidation reaction increases the rate by a factor of one million
(Langmuir, 1997).
The precipitate, Fe (OH)3, recrystallizes to hematite or goethite at a
rate that is dependent on pH, temperature, and solution composition.
At 50 C and pH = 6.4, the half life for recrystallization is 8.3 days and
recrystallization is 99% complete in 55 days calculated with the rate
law given in Shaw et al. (2005) and the rate constant of Fischer (1971)
reported in Shaw et al. (2005).
Concretion growth was transport controlled because of the rapid
oxidation of Fe2+ with PO2 above 1 10 5, the rapid conversion of Fe
(OH)3 to goethite, and the lack of a depleted zone surrounding
concretions.
7. Conceptual reaction-nucleation model
Concretion growth models previously proposed incorporate
diffusion and advective transport of reactants, but do not consider
chemical reaction with minerals in the host rock, nor do the growth
models explain the spacing between concretions or the presence of
the rinds. The purpose of this section is to explain Liesegang-Ostwald
self-organization and the role of pore solution reacting with Kfeldspar in concretion nucleation and growth.

W.T. Parry / Sedimentary Geology 233 (2011) 5368

Rate = As k aH 1Q = K in mole = cm = s
The rate constant at an elevated temperature is calculated using
the Arrhenius equation

kt

E = RT

= Ae

Numerous eld studies have demonstrated that the natural


feldspar dissolution rates are 2 to 5 orders of magnitude slower
than laboratory rates under similar conditions of pH and temperature
(e. g. Blum and Stillings, 1995; Velbel, 1993; White and Brantley,
2003; Zhu, 2005). K-feldspar grains from the Navajo Sandstone at
some localities are coated with clay minerals that drastically retard
feldspar reaction (Zhu, 2005). However, the small rock volumes in a
concretion interior considered here and the lack of clay coatings on
the K-feldspar in host sandstone suggest reaction rates may not
approach the slowest rates observed in eld studies. Calculations
presented below are based on a reaction rate that is 3 orders of
magnitude slower than the laboratory rate.
K-feldspar reacts with H+ that diffuses from solution in surrounding porous rock. The concretions are regularly spaced with spacing
that varies with the concretion size (Chan et al., 2004). The rate at
which H+ would diffuse into the volume of reacting K-feldspar was
Fig. 12. Conceptual model of the Ostwald-Liesegang cycle that is thought to govern
concretion spacing. Iron bearing water owing to the right and oxygen from the oxygen
rich water on the right diffuses upstream in A. The saturation index rises in B and C and
reaches the nucleation threshold in D causing precipitation of goethite and lowering Fe
content to very low values. Fe diffuses to the concretion and the concretion grows. The
saturation index falls to the equilibrium value in E. As the front moves downstream,
oxygen diffusion no longer reaches the concretion in E so that a new saturation index
rises in F and G, reaches the nucleation threshold in H and a new concretion forms.

calculated from the diffusion distance of the half-spacing between the


concretions, and the difference in H+ concentration from ambient
solution pH and H+ at the solubility minimum. The H+ diffusion ux

Nucleation Threshold

1.5
Fe2+
1.0

Equilibrium Saturation
Moving front

0.5

O2(aq)

1.5

Fe2+

1.0
O2(aq)

0.5
0

Goethite saturation level

1.5
Fe2+
1.0
O2(aq)

0.5
0

1.5

Fe2+

1.0
O2(aq)

0.5

Goethite Saturation Level

The reactants for precipitation of goethite, Fe2+ and O2, cannot be


carried within the same solution; two separate solutions are
necessary. Precipitation can be initiated by addition of oxygen or
consumption of H+ by reactions with K-feldspar. Iron concretions and
bands form by an Ostwald-Liesegang cycle (Ortoleva, 1994) and were
interpreted for the Navajo Sandstone by Chan et al. (2000, 2007). In
this cycle, reducing solutions owed upward along faults into the
Navajo Sandstone, reduced the iron, and dissolved the hematite. The
reducing water now charged with Fe2+ owed through the Navajo
Sandstone, reacted with K-feldspar, and encountered oxygenated
groundwater (Fig. 12A). Upstream diffusion of O2 together with
downstream diffusion of Fe2+ and consumption of H+ by reaction
with K-feldspar elevated the reaction product until it reached a level
of supersaturation that promoted nucleation and precipitation of iron
oxide (Fig. 12B-D). The reaction product then fell to its equilibrium
value (Fig. 12E). In this model reactants Fe and O2 were supplied by
diffusion, H+ was consumed by the K-feldspar reaction and the
concretion grew. The Fe concentration at the concretion precipitation
site was very low at goethite solubility (Fig. 8) so Fe diffused to the
concretion from all directions causing nearly spherical growth. When
the advection front of reducing, Fe2+-charged water proceeded
downstream, diffusion of O2 was no longer effective in maintaining
saturation, so a new supersaturation-nucleation level was reached
downstream and growth of a new concretion began (Fig. 12F-I).
Therefore, spacing and size of the concretions depends on reactant
Fe2+, O2, and H+ supply rates.
For K-feldspar reaction to raise the pH and cause precipitation of
goethite, H+ must be consumed more rapidly than it can be supplied
by diffusion. The following conceptual model incorporates reaction
with host-rock K-feldspar and evaluates reaction rates compared with
H+ diffusion rates to the site of reaction. The rate at which H+ is
consumed for the volume of K-feldspar contained in the interior of
rind concretions of various sizes was calculated using the following
rate model (Bethke, 1998). Variables are shown in Table 4.

63

1.5
Fe2+
1.0
O2(aq)

0.5
0

1.5
Fe2+
1.0
O2(aq)

0.5
0

1.5

Fe2+

1.0
O2(aq)

0.5
0

1.5
Fe2+
1.0
O2(aq)

0.5
0

1.5
Fe2+
1.0
O2(aq)

0.5
0

I
Distance

W.T. Parry / Sedimentary Geology 233 (2011) 5368

Table 4
Variables used in calculation of the reaction model.
2

As = K-feldspar surface area; 374 cm /g (57.9 in /g) calculated using the equation of
Brantley and Mellott (2000) for mean grain size of 0.125 mm (0.0049 in).
k+ = the laboratory rate constant at 25 C is 1.168 10 14 mol/cm2/s. The reaction
model assumes a dissolution rate 3 orders of magnitude slower than the
laboratory rate.
aH+ = activity of hydrogen ion.
Q = reaction quotient.
K = equilibrium constant.
n (exponent on H+) = 0.5 (Blum and Stillings, 1995)
A = pre-exponential factor = 1.347 10-5.
E = activation energy = 51.7 kJ/mol for acid solutions (Blum and Stillings, 1995).
R = gas constant.
T = absolute temperature.
D = diffusion coefcient.
= porosity.
rb = outer radius of H+ depleted zone (half spacing between concretions).
rs = radius of concretion.
cb = concentration of H+ at the outer radius of the depleted zone.
cs = concentration of H+ at the concretion surface.

is calculated using the equation for spherical diffusion developed by


Lasaga (1998):
Flux = D

rb cb cs
rs rb rs

Variables in this equation are in Table 4, and the main components


of the physical-chemical conceptual model are illustrated in Fig. 13.
The results of these calculations (Fig. 14) illustrate that the rate of
diffusion of H+ to the concretion and the volume of sandstone
determines the rate of H+ consumption by reacting K-feldspar. The

Log (H+ Flux or Reaction) (moles/yr)

64

-7.5

ld

K-fe

(cb-cs) mol/cm3

-8.5

re
spar

=5

n pH

actio

-8.0
23x10-13
12x10-13

ction

r rea

spa
K-feld

-9.0

-7

pH=6

4x10-13

-9.5
-10.0

10

15

20

25

Concretion Radius (mm)


Fig. 14. Comparison of H+ diffusion with H+ consumption by reaction with K-feldspar
volumes of various diameters. As K-feldspar volume increases, lower pH is required for
H+ diffusion to the reacting K-feldspar. If K-feldspar consumes H+ faster than diffusion
can supply it, then the pH raises causing precipitation of hydrous iron oxide.

pH at precipitation of ferric oxide is assumed to be at the solubility


minimum (Fig. 8). When the rate of consumption exceeds the rate of
supply, the pH will rise and cause precipitation of ferric oxide. For the
lowest H+ concentration difference (cb-cs = 4 10 13 mol/cm3)
considered, the rate of consumption of H+ by reacting K-feldspar at
pH = 6 or 7 exceeds the rate of supply of H+ by diffusion at a
concretion radius of 14 mm (Fig. 14). If K-feldspar reacts at pH = 5 the
radius is 20 mm. For an intermediate H+ concentration difference of
12 10 13 mol/cm3, the concretion size when K-feldspar consumption of H+ exceeds supply is 21 mm radius for K-feldspar reacting at
pH 6 to 7 or 12 mm in radius if the reaction takes place at pH = 5. For
the greatest H+ concentration difference of 23 10 13 mol/cm3, a
volume of sandstone 29 mm in radius consumes H+ faster than
supplied by diffusion if the reaction takes place at pH = 6 or 7 and
18 mm if the reaction takes place at pH = 5. A single K-feldspar grain
can raise the pH and nucleate goethite only if the pH is very close to
the saturation value. For a solution pH level far from goethite
saturation, a larger mass of K-feldspar is needed to consume H+
faster than supplied by diffusion, so that pH will rise and cause
precipitation. With this model, concretion size is related to the pH of
solution reacting with K-feldspar and the pH gradient surrounding the
concretion. Concretion spacing is determined in the LiesegangOstwald cycle by the upstream diffusion of O2 and ow velocity.
8. Mass balance calculations

Fig. 13. Conceptual reaction-nucleation model. A. regional geological cross section


showing chemically reducing uids gaining access to the Navajo Sandstone along a fault
conduit, bleaching the sandstone, and precipitating concretions at the chemical
reaction front. B. Diagram of an array of concretions showing dimensions used in the
reaction-nucleation calculations.

Mass balance estimates presented below evaluate the possibility


that the iron in concretions is derived from diffusion of reactants in
static pore uid, from redistribution of iron by diffusion in
immediately surrounding rock, or from pore uid ow. The spacing
between spherical concretions is related to the concretion diameter.
The average spacing of 5 cm diameter concretions is 40 cm (Chan et
al., 2004). The volume of a 40 cm diameter volume of porous rock
surrounding each concretion is 33,000 cm3. The porosity of the
sandstone is approximately 15% so the pore volume is 5,000 cm3.
Chemical modeling of reduction of hematite in the original sandstone
by dissolved methane (Parry et al., 2004) suggests the maximum
dissolved iron in the pore space is near 15 mg/l. Static pore uid
surrounding a concretion could contain a maximum of 0.075 g of Fe. A
5 cm diameter concretion with a rind 2 mm thick has a rind volume of
14 cm3 that is 20% goethite and contains 8 g of Fe, but 0.075 g of Fe is
available in the surrounding pore uid. Fe must therefore be supplied
by pore uid ow (100 pore volumes) or redistribution of Fe mineral
in the surrounding rock. Navajo Sandstone with hematite grain
coatings from early diagenesis originally contained on average 0.53
wt. % Fe2O3 (range 0.18 to 1.25%, Beitler et al., 2005) or 350 g of Fe,
ample to supply the Fe in the concretion rind. If the Fe in host rock

W.T. Parry / Sedimentary Geology 233 (2011) 5368

near a concretion were dissolved, transport of the Fe to the concretion


would be related to distance with more Fe transported from shorter
distances producing an iron depletion halo in the host rock. However,
no Fe-depleted halo is apparent surrounding the concretions (Table 1,
Fig. 4). The host rock surrounding the concretion averages 1.39% Fe as
goethite and lepidocrocite nuclei on illite-coated quartz grains and Kfeldspar grains from advective addition of iron. The original iron was
not redistributed to form the concretion rind but iron was added to
the rock by advection. The source of the iron is chemical reduction of
hematite coatings on sand grains by a reducing, hydrocarbon-bearing
solution. The reduced iron is transported in solution to the site of
oxidation and concretion formation.
9. Discussion
Concretions represent an end stage in diagenesis of iron minerals
in red beds. The iron mineral diagenesis begins with an early stage of
reddening that forms as ferrous silicates oxidize and form hematite
coatings on detrital grains during deposition and early burial.
Hydrocarbons from underlying source beds ow into the sediment
along fault conduits and react with the hematite coatings to
chemically reduce and mobilize iron and bleach the sandstone
(Parry and Blamey, 2010). Chemical reduction and iron removal
changes the color from reddish brown to nearly white recording uid
ow pathways. The mobilized iron remains in solution during
transport until an oxidizing solution is encountered. The iron in
solution is then precipitated as a result of oxidation. Bacteria may
mediate both the chemical reduction and oxidation of iron.
Concretion growth rates calculated in numerical growth models
are sensitive to concentration of reactants in solution, temperature,
and uid ow rate, but do not consider reaction of uids with host
rock minerals such as K-feldspar, identity of precipitate, nor timing of
concretion growth. Growth times using the iron concentration in the
modeling of Chan et al. (2007) that are 10 orders of magnitude less
than those used here produce growth times that exceed the Jurassic
age of the Navajo Sandstone. Doubling the iron content of groundwater to 10 mg/l cuts growth time by half. Increasing temperature to
100 C doubles the iron diffusion coefcient and decreases growth
time by a factor of 2, but hematite would be more likely to precipitate
instead of goethite. Decreasing the ow velocity below the 2 m/yr
used in the models produces growth times between the diffusion only
calculations and the diffusion with advection.
Petrographic observation of FeO(OH) nuclei forming on K-feldspar
and illite suggests that H+ consumption by reacting silicates as well as
oxidation causes the precipitation. Reacting K-feldspar could cause a
pH gradient to form if H+ consumption rate exceeds the rate of H+
supply by diffusion. Calculations suggest that the concretion rind is
related to ambient solution pH. Low solution pH reacting with regions
of K-feldspar results in growth of a rind with a large diameter, because
the larger volume of K-feldspar consumes H+ at a greater rate than
can be supplied by diffusion; pH then rises causing precipitation of
goethite. The increased pH caused by K-feldspar reaction of smaller
volumes with higher pH solutions causes precipitation of goethite
with smaller volumes of K-feldspar and thus smaller concretions.
The goethite rim forms a protective layer of reduced porosity and
permeability that inhibits chemical reactions with K-feldspar in the
concretion interior with exterior solutions. The porosity of the goethite
rim is 1.8% (Table 1) reducing permeability by 5 orders of magnitude
(Chidsey et al., 2007). Thus, solutions that precipitated lepidocrocite in
host sandstone were inhibited from reaching the concretion interior.
Clay mineralogy differs in interior versus exterior samples. The interior
is protected from reactants in solution that form kaolinite from
alteration of K-feldspar so the interior contains one-third to one-half
the kaolinite as the exterior samples (Table 1, Fig. 4). Formation of 1%
kaolinite in the exterior host rock would require approximately 1 m. y.
from rates of K-feldspar reaction shown in Fig. 14.

65

Lepidocrocite present in host sandstone outside concretion rinds is


10 times more soluble than goethite (Fig. 8) so some mechanism must
prevent precipitation of lower solubility goethite. Precipitation of
lepidocrocite over goethite is favored by phosphate, high oxidation
rate, and pH above 7, and inhibited by dissolved Si and dissolved CO2
(Carlson and Schwertmann, 1990; Cumplido et al., 2000; Schwertman
and Thalmann, 1976). The host sandstone contains up to 0.07 wt. %
P2O5 (Table 1). Phosphate concentrations in groundwater from the
Navajo Sandstone are seldom above 0.03 mg/l, but do reach a
maximum of 0.16 mg/l (Hood and Danielson, 1979, 1981; Hood and
Patterson, 1984). The ratio P/Fe varies from b0.003 to 0.016 at the
assumed Fe concentration used in the modeling. The rate of oxidation
of Fe2+ increases with an increase in P/Fe ratio up to 0.2 (Cumplido et
al., 2000). High oxidation rate is therefore suggested as the most likely
cause of metastable lepidocrocite precipitation rather than more
stable goethite.
9.1. Timing of concretion formation
Formation of goethite concretions involves mobilizing iron from
hematite rims on detrital grains of quartz and feldspar that formed
during early diagenesis (Fig. 15). Then, the iron precipitates as
concretions when the iron-bearing solutions encountered oxidizing
conditions. Concentration of Fe3+ in equilibrium with iron oxides is
negligibly small so the likely mechanism of mobilization is chemical
reduction of the iron to Fe2+ by organic carbon. Other reductants
considered include hydrogen sulde that would immobilize the iron
as pyrite (Fig. 7), methane, or organic acids (Garden et al., 2001). The
timing of the mobilization step was related to hydrocarbon migration,
for example, the iron content of produced water from the Covenant oil
eld reservoir in the Navajo Sandstone is 14 mg/kg (Chidsey et al.,
2007) near the maximum that can be dissolved by chemical reduction
at near-neutral pH.
Hydrocarbon migrated from Mississippian source beds during
Cretaceous thrust faulting into the Lower Permian White Rim
Sandstone reservoir during Middle Tertiary Laramide deformation
(Huntoon et al., 1999) (Fig. 15). Oil migrated into the Navajo
Sandstone in the Covenant oil eld during Cretaceous to Early Tertiary
(Chidsey et al., 2007). In the Moab area, hydrocarbon inux and iron
reduction occurred at 55-60 Ma (Garden et al., 2001).
The original porosity of the Navajo Sandstone is unknown, but
could have been as much as 49% (Atkins and McBride, 1992). The
average present intergranular volume (sum of intergranular porosity
and cements) is 20.8% (Table 1), the intergranular volume predicted
for 2 to 4 km of burial (Lander and Walderhaug, 1999). The iron oxide
cementation is expected to preserve the intergranular volume at the
time of cementation and must have taken place after 2 to 4 km of
burial compaction. The Navajo Sandstone is overlain by 1.9 to 4.2 km
of Jurassic to Eocene strata (Fig. 2). An inux of oxidizing meteoric
water would have been caused by local topographically driven ow
gradients that accompanied uplift and exhumation of the Colorado
Plateau. Uplift of the Colorado began in middle Eocene and continued
to late Miocene subjecting the area to exhumation and development
of the Colorado River system that continues today.
9.2. Bacterial mediation of iron reduction and oxidation
Formation of lepidocrocite and goethite by microbial Fe2+
reduction does not differ from abiotic Fe2+ oxidation with O2; but
solution geochemistry is responsible for the mineralogy of most Fe3+
precipitates (Larese-Casanova et al., 2010). Iron oxides in two typical
iron oxide concretions from Utah resemble known biosignatures, but
organic compounds could not be detected with gas chromatographymass spectrometry analysis (Souza-Egipsy et al., 2006). Host rock
samples for the present study average 0.04 wt. % C, but no carbonate
minerals could be detected. The isotopic composition of Fe in the

66

W.T. Parry / Sedimentary Geology 233 (2011) 5368

Fig. 15. Paragenesis diagram showing major diagenetic and tectonic events affecting the Navajo Sandstone.

concretions provides the only evidence of bacterial involvement in the


iron diagenesis system. The Fe-oxide concretions show a relatively
large range in 56Fe from + 0.9 to 2.0 (Chan et al., 2006). The
negative values result not from simple, closed-system remobilization
of 0 Fe-oxide, but instead formed in an open system. The
concretions probably formed by complete oxidation of aqueous
Fe2+, and the negative 56Fe must have resulted from negative values
of Fe2+ in solution that resulted from mobilization of Fe by chemical
reduction mediated by Fe3+ reducing bacteria (Chan et al., 2006).
Mediation of iron oxidation reactions by bacteria has no effect on the
concretion growth models, because the growth is limited by transport
of iron not by the rate of oxidation.
10. Conclusions
A physical, chemical, diffusion, and advection transport model has
been applied to mobilization of iron and deposition of hydrous iron
oxide concretions. Iron oxide concretions form in porous sandstone by
chemical reduction and solution of iron minerals, transport of Fe2+ in
solution, and precipitation of hydrous ferric oxide by oxidation.
Concretions form quickly in 1000's of years by diffusion and
advection, and K-feldspar consumption of H+ initiates nucleation.
The Jurassic Navajo Sandstone, host for the iron oxide concretions
in this study, averages (volume % one standard deviation) 74 2 %
quartz, 5.6 0.6 % K-feldspar, 2.1 0.2 % illite, 1.1 0.8 % kaolinite
and 0.7 0.1 % FeO(OH) as goethite or lepidocrocite. Concretions
consist of an outer rind of goethite almost completely lling the
porosity (19.6 volume %) in sandstone. A smaller amount of goethite is
also disseminated within the interior of the concretions. Lepidocrocite
is present in the exterior sandstone as light tan rims on sand grains.
Concretions form by diffusion and advection of iron that was
previously dissolved and mobilized during chemical reduction.
Calculated solubility of goethite and lepidocrocite are extremely low
and signicant iron mobility requires reduction to Fe2+ species. A
calculated mineral stability diagram shows that precipitation would
result from an increase in pH and an increase in log fO2. Formation
times for rind concretions vary with rind thickness, concretion
diameter, and pore uid ow velocity. For example, assuming a
temperature of 50 C, ow velocity of 2 m/yr, and 5 mg/l Fe in
solution, a 2 mm thick rind on a 25 mm radius concretion requires
2000 years to form from diffusion and advection and 3600 years to
form from diffusion only (Fig. 11A). A concretion with a 10 mm radius
consisting of completely cemented sandstone is near the largest solid
concretion observed and requires 2800 years to form from diffusion
and an advection velocity of 2 m/yr and 3800 years to form from
diffusion alone (Fig. 10). Concretion growth times are limited by Fe

transport, because rates of precipitation and recrystallization of iron


minerals are very rapid compared to transport rates. Bacterial
mediation of the reactions by bacteria increases the rate of
precipitation and thus has no effect on growth time.
Mass balance calculations show that the iron in a rind or solid
concretion cannot be supplied by dissolved iron in a static pore uid.
Chemical analysis of host sandstone surrounding concretions did not
detect an iron depletion halo. Iron in the concretions cannot result
from redistribution of iron in the immediately surrounding sandstone,
but must be supplied by pore uid ow.
Petrographic study identied goethite nuclei on detrital K-feldspar
grains. Consumption of H+ by pore uid reaction with the feldspar
elevated the pH to cause precipitation in oxidizing solutions.
Comparison of K-feldspar reaction rates (assumed to be 1000 times
slower than laboratory rates) to H+ diffusion rates at various assumed
H+ concentration differences between precipitation site and the
midpoint between adjacent concretions results in an estimated
concretion size. A 14 mm radius concretion would form at H+
concentration difference of about 4 10-13 mol/cm3, a 21 mm radius
concretion would form at H+ difference of about 12 10-13 mol/cm3,
and a 29 mm radius concretion would form at H+ difference of about
23 10-13 mol/cm3, because K-feldspar regions of each of these
diameters would consume H+ faster than supplied by diffusion.
The concretions complete the record of uid ow events,
oxidation-reduction conditions, pH, and microbial activity involved
in iron mineral diagenesis. The record began with early diagenetic
reddening followed by bleaching by solutions containing hydrocarbon
and Fe transportation to the site of concretion formation. Petrographic, chemical, mineralogical, and modeling studies have improved the
limited understanding of conditions, processes, and time scales in
which iron oxide concretions form.
Acknowledgments
This material is based upon work supported by the National
Aeronautics and Space Administration under Agreement No.
NNG06GI10G issued through the Science Mission Directorate Fundamental Research Program, Mars Fundamental Research. Reviews by P.
Eichhubl, E. Sefton-Nash, M. A. Chan, K. Milliken, and M. Wilkinson are
greatly appreciated.
References
Atkins, J.E., McBride, E.F., 1992. Porosity and packing of Holocene river, dune, and beach
sands. American Association of Petroleum Geologists Bulletin 76, 339355.
Bailey, E.H., Stevens, R.E., 1960. Selective staining of K-feldspar and plagioclase on rock
slabs and thin sections. American Mineralogist 45, 10101025.

W.T. Parry / Sedimentary Geology 233 (2011) 5368


Beitler, B., Chan, M.A., Parry, W.T., 2003. Bleaching of Jurassic Navajo Sandstone on
Colorado Plateau Laramide highs: Evidence of exhumed hydrocarbon supergiants?
Geology 31, 10411044.
Beitler, B., Parry, W.T., Chan, M.A., 2005. Fingerprints of uid ow: chemical diagenetic
history of the Jurassic Navajo Sandstone, Southern Utah. U.S.A. Journal of
Sedimentary Research 75, 547561.
Bergosh, J.L., Good, J.R., Hillman, J.T., Kolodzie, S., 1982. Geological characterization of
the Nugget Sandstone, Anschutz Ranch East. In: Nielson, D.L. (Ed.), Overthrust Belt
of Utah. : Utah Geological Association Publication, 10. Utah Geological Association,
Salt Lake City, Utah, pp. 253265.
Berner, R.A., 1968. Rate of concretion growth. Geochimica et Cosmochimica Acta 32,
477483.
Berner, R.A., 1980. Early diagenesis: A theoretical approach. Princeton University Press,
Princeton. 241 pp.
Bethke, C.M., 1998. The Geochemist's Workbench. University of Illinois. 184 pp.
Blakey, R.C., 1994. Paleogeographic and tectonic controls on some Lower and Middle
Jurassic erg deposits, Colorado Plateau. In: Caputo, M.V., Peterson, J.A., Franczyk, K.J.
(Eds.), Mesozoic systems of the Rocky Mountain region, U.S.A. Rocky Mountain
Society for Sedimentary Geology, Denver, Colorado, pp. 273298.
Blakey, R.C., Peterson, F., Kocurek, G., 1988. Synthesis of late Paleozoic and Mesozoic
eolian deposits of the Western Interior of the United States. Sedimentary Geology
56, 3125.
Blum, A.E., Stillings, L.L., 1995. Feldspar dissolution kinetics. In: White, A.F., Brantley, S.L. (Eds.),
Chemical weathering rates of silicate minerals: Reviews in Mineralogy, 31, pp. 291351.
Bowen, B.B., 2005, Sandstone bleaching and iron concretions: An index to uid
pathways and diagenetic history of Jurassic Navajo Sandstone, southern Utah.
Ph. D. thesis, University of Utah Salt Lake City, Utah, 221 pp.
Bowen, B.B., Martini, B.A., Chan, M.A., Parry, W.T., 2007. Reectance spectroscopic
mapping of diagenetic heterogeneities and uid-ow pathways in the Jurassic
Navajo Sandstone. American Association of Petroleum Geologists Bulletin 91,
173190. doi:10.1306/08220605175.
Brantley, S.L., Mellott, N.P., 2000. Surface area and porosity of primary silicate minerals.
American Mineralogist 85, 17671783.
Busigny, V., Dauphas, N., 2007. Tracing paleouid circulations using iron isotopes; a
study of hematite and goethite concretions from the Navajo Sandstone (Utah, USA).
Earth and Planetary Science Letters 254, 272287.
Carlson, L., Schwertmann, U., 1990. The effect of CO2 and oxidation rate on the
formation of goethite versus lepidocrocite from and Fe(II) system at pH 6 and 7.
Clay Minerals 25, 6571.
Chan, M.A., Parry, W.T., Bowman, J.R., 2000. Diagenetic hematite and manganese oxides
and fault-related uid ow in Jurassic sandstones, southeastern Utah. American
Association of Petroleum Geologists Bulletin 84, 12811310.
Chan, M.A., Beitler, B., Parry, W.T., Orm, J., Komatsu, G., 2004. A possible terrestrial
analogue for haematite concretions on Mars. Nature 429, 731734. doi:10.1038/
nature02600.
Chan, M.A., Beitler Bowen, B., Parry, W.T., Orm, J., Komatsu, G., 2005. Red rock and red
planet diagenesis: Comparisons of Earth and Mars concretions. GSA Today 15, 410.
doi:10.1130/1052-5173(2005) 015[4:RRARPD]2.0.CO;2.
Chan, M.A., Johnson, C.M., Beard, B.L., Bowman, J.R., Parry, W.T., 2006. Iron isotopes
constrain the pathways and formation mechanisms of terrestrial oxide concretions:
A tool for tracing iron cycling on Mars? Geosphere 2, 324332. doi:10.1130/
GES00051.1.
Chan, M.A., Orm, J., Park, A.J., Stich, M., Souza-Egipsy, V., Komatsu, G., 2007. Models of
iron oxide concretion formation: eld, numerical and laboratory comparisons.
Geouids 7, 356368. doi:10.1111/j.1468-8123.2007.00187.x.
Chidsey, T.C., Morgan, C.D., 2005. Major oil plays in Utah and vicinity. Quarterly
technical progress report 2005. Utah Geological Survey. 64 pp.
Chidsey Jr., T.C., DeHamer, J.S., Hartwick, E.E., Johnson, K.R., Schelling, D.D., Sprinkel, D.A.,
Strickland, D.K., Vrona, J.P., Wavrek, D.A., 2007. Petroleum geology of Covenant oil
eld, central Utah thrust belt. In: Willis, G.C., Hylland, M.D., Clark, D.L., Chidsey Jr., T.C.
(Eds.), Central Utahdiverse geology of a dynamic landscape: Utah Geological
Association Publication, 36, pp. 273296.
Cooley, J.E., J. Harshbarger, J.W., Akers, J.P., Hardt, W.F., 1969. Regional hydrogeology of the
Navajo and Hopi Indian reservations, Arizona, New Mexico, and Utah, with a section
on vegetation by O. N. Hicks. U. S. Geological Survey Professional Paper 400-B, 61 p.
United States Government Printing Ofce, Washington, D. C.
Cordova, R.M., 1978. Ground-water conditions in the Navajo Sandstone in the Central
Virgin River Basin, Utah. Utah Department of Natural Resources Technical
Publication no. 40, 64 p.
Cumplido, J., Baron, V., Torrent, J., 2000. Effect of phosphate on formation of nanophase
lepidocrocite from Fe(II) sulfate. Clays and Clay Minerals 48, 503510.
Davis, G.H., 1999. Structural geology of the Colorado Plateau region of southern Utah,
with special emphasis on deformation bands. Geological Society of America Special
Paper, 342. 157 pp.
DeCelles, P.G., Coogan, J.C., 2006. Regional and kinematic history of Sevier fold-andthrust belt, central Utah. GSA Bulletin 118, 841864.
Dickinson, W.R., Snyder, W.S., 1978. Plate tectonics of the Laramide orogeny. In:
Matthews, V.I.I.I. (Ed.), Laramide folding associated with basement block
faulting in the western United States: Geological Society of America Memoir,
151, pp. 355366.
Dickinson, W.R., Klute, M.A., Hayes, M.J., Janecke, S.U., Lundin, E.R., McKittrick, M.A.,
Oliveres, M.D., 1988. Paleogeographic and paleotectonic setting of Laramide
sedimentary basins in the central Rocky Mountain region. Geological Society of
America Bulletin 100, 10231039.
Doelling, H.H., Blacket, R.E., Hamblin, A.H., Powell, J.D., Pollock, G.L., 2000. Geology of Grand
Staircase-Escalante National Monument, Utah. In: Sprinkel, D.A., Chidsey, T.C.,

67

Anderson, P.B. (Eds.), Geology of Utah's Parks and Monuments: Utah Geological
Association Publication, #28, pp. 189231.
Drever, J.I., 1997. The geochemistry of natural waters, Surface and groundwater
environments, Third edition. Prentice Hall, Upper Saddle River, N. J. 436 pp.
Eichhubl, P., Taylor, W.L., Pollard, D.D., Aydin, A., 2004. Paleo-uid ow and
deformation in the Aztec Sandstone at the Valley of Fire, Nevada - Evidence for
the coupling of hydrologeologic, diagenetic, and tectonic processes. GSA Bulletin
116, 11201136.
Fischer, W. R., 1971. Modellversuch zur Bilung und Auosung von Goethit und
amorphen Eisenoxiden im Boden. Dissertation, Techn, Univ. Munchen.
Freethey, G.W., Cordy, G.E., 1991. Geohydrology of Mesozoic rocks in the upper Colorado
River Basin - excluding the San Juan Basin - in Arizona, Colorado, New Mexico, Utah
and Wyoming. U. S. Geological Survey Professional Paper 1411-C, 118 p. United
States Government Printing Ofce, Washington, D. C.
Garden, I.R., Guscott, S.C., Burley, S.D., Foxford, K.A., Walsh, J.J., Marshal, J., 2001. An
exhumed palaeo-hydrocarbon migration fairway in a faulted carrier system,
Entrada Sandstone of SE Utah, USA. Geouids 1, 195213.
Hintze, L.G., Kowallis, B.J., 2009. Geologic history of Utah. Brigham Young University
Geology Studies Special Publication, 9. Provo, Utah, 225 pp. Department of
Geological Sciences Brigham Young University, Provo, Utah.
Hood, J.W., Danielson, T.W., 1979. Aquifer tests of the Navajo Sandstone near Caineville,
Wayne County, Utah. State of Utah Department of Natural Resources Technical
Publication, no. 66. 69 pp.
Hood, J.W., Danielson, T.W., 1981. Bedrock aquifers in the lower Dirty Devil basin area,
Utah with special emphasis on the Navajo Sandstone. State of Utah Department of
Natural Resources Technical Publication, 68. 143 pp.
Hood, J.W., Patterson, D.J., 1984. Bedrock aquifers in the northern San Rafael Swell area,
Utah with special emphasis on the Navajo Sandstone. State of Utah Department of
Natural Resosurces Technical Publication, 78. 127 pp.
Hunt, C.B., 1956. Cenozoic geology of the Colorado Plateau. U. S. Geological Survey
Professional Paper, 279. Washington, D. C., 99 pp. United States Government
Printing Ofce, Washington, D. C.
Huntoon, J.E., Hansley, P.L., Naeser, N.D., 1999. The search for a source rock for the giant
tar sand triangle accumulation, southeastern Utah. American Association of
Petroleum Geologists Bulletin 83, 467495.
Johnson, J.W., Oelkers, E.H., Helgeson, H.C., 1992. SUPCRT92 software package for
calculating the standard molal thermodynamic properties of minerals, gases,
aqueous species, and reactions from 1 to 5000 bars and 0 to 1000 C. Computers
and Geosciences 18, 899947.
Jordan, W., 1965. Regional environmental study of the early Mesozoic Nugget and
Navajo Sandstones. PhD dissertation, University of Wisconsin, Madison, 206 pp.
Lander, R.H., Walderhaug, O., 1999. Predicting porosity through simulating sandstone
compaction and quartz cementation. American Association of Petroleum Geologists
Bulletin 83, 433449.
Langmuir, D., 1997. Aqueous environmental geochemistry. Prentice Hall, Upper Saddle
River, New Jersey. 600 pp.
Larese-Casanova, P., Haderlein, S.B., Kappler, A., 2010. Biomineralization of lepidocrocite
and goethite by nitrate-reducing Fe(II)-oxidizing bacteria: Effect of pH, bicarbonate,
phosphate, and humic acids. Geochimica et Cosmochimica Acta 74, 37213734.
Lasaga, A.C., 1998. Kinetic Theory in the Earth Sciences. Princeton University Press,
Princeton, New Jersey.
Lindquist, S.J., 1983. Nugget Formation reservoir characteristics affecting production in
the overthrust belt of southwestern Wyoming. Journal of Petroleum Technology 35,
13551365.
Lindquist, S.J., 1988. Practical characterization of eolian reservoirs for development:
Nugget Sandstone, Utah-Wyoming thrust belt. Sedimentary Geology 56, 315339.
Liu, X., Millero, F.J., 1999. The solubility of iron hydroxide in sodium chloride solutions.
Geochimica et Cosmochimica Acta 63, 34873497.
Majzlan, J., Lang, B.E., Stevens, R., Navrotsky, A., Woodeld, B.F., Boerio-Goates, J., 2003a.
Thermodynamics of Fe oxides: Part I. Entropy at standard temperature and
pressure and head capacity of goethite (-FeOOH), lepidocrocite (-FeOOH), and
maghemite (-Fe2O3). American Mineralogist 88, 846854.
Majzlan, J., Grevel, K.-D., Navrotsky, A., 2003b. Thermodynamics of Fe oxides: Part II.
Enthalpies of formation and relative stability of goethite (-FeOOH), lepidocrocite (-FeOOH) and maghemite (-Fe2O 3). American Mineralogist 88,
855859.
Navrotsky, A., Mazeina, L., Majzlan, J., 2008. Size-driven structural and thermodynamic
complexity in iron oxides. Science 319, 16351638.
Net, L., 2003. Diagenesis and reservoir quality of the eolian Nugget/Navajo Sandstone
(Early Jurassic), Utah and Wyoming. M. S. Thesis, University of Texas, Austin, 183 pp.
Nuccio, V.F., Condon, S.M., 2000. Burial and thermal history of the Paradox Basin, Utah
and Colorado and petroleum potential of the Middle Pennsylvanian Paradox Basin.
U. S. Geological Survey, Bulletin 2000-O. 41 pp.
Oelkers, E.H., Helgeson, H.C., 1988. Calculation of the thermodynamic and transport
properties of aqueous species at high pressures and temperatures: aqueous tracer
diffusion coefcients of ions to 1000 C and 5 kb. Geochimica et Cosmochimica Acta
52, 6385.
Orm, J., Komatsu, G., Chan, M.A., Beitler, B., Parry, W.T., 2004. Geological features
indicative of processes related to hematite formation in Meridiani Planum and
Aram Chaos, Mars: A comparison with diagenetic hematite deposits in southern
Utah, USA. Icarus 171, 295316. doi:10.1016/j.icarus.2004.06.001.
Ortoleva, P.J., 1994. Geochemical Self-Organization. Oxford University Press, New
York. 411 pp.
Parry, W.T., Blamey, N.J.F., 2010. Fault uid composition from uid inclusion
measurements, Laramide age Uinta thrust fault, Utah. Chemical Geology 278,
105119.

68

W.T. Parry / Sedimentary Geology 233 (2011) 5368

Parry, W.T., Ballantyne, J.M., Bryant, N.L., 1980. Hydrothermal alteration enthalpy and
heat ow in the Roosevelt Hot Springs thermal area, Utah. Journal of Geophysical
Research 85, 25592566.
Parry, W.T., Chan, M.A., Beitler, B., 2004. Chemical bleaching indicates uid ow in
sandstone deformation bands. American Association of Petroleum Geologists
Bulletin 88, 175191.
Parry, W.T., Forster, C.B., Evans, J.P., Bowen, B.B., Chan, M.A., 2007. Geochemistry of CO2
sequestration in the Jurassic Navajo Sandstone, Colorado Plateau, Utah. Environmental Geosciences 14, 91109. doi:10.1306/eg.07120606004.
Parry, W.T., Chan, M.A., Nash, B.P., 2009. Diagenetic characteristics of the Jurassic
Navajo Sandstone in the Covenant oil eld, central Utah thrust belt. American
Association of Petroleum Geologists Bulletin 93, 10391061.
Pederson, J.L., Mackley, R.D., Eddleman, J.L., 2002. Colorado Plateau uplift and erosion
evaluated using GIS. GSA Today 12, 410.
Sahagian, D., Proussevitch, A., Carlson, W., 2002. Timing of Colorado Plateau uplift:
Initial constraints from vesicular basalt-derived paleoelevations. Geology 30,
807810.
Sanford, R.F., 1995. Ground-water ow and migration of hydrocarbons to the lower
Permian White Rim Sandstone, Tar Sand Triangle, Southeastern Utah. U. S.
Geological Survey Bulletin 2000-J. 25 pp.
Schwertman, U., Thalmann, H., 1976. The inuence of [Fe(II)], [Si] and pH on the
formation of lepidocrocite and ferrihydrite during oxidation of aqueous FeCl2
solutions. Clay Minerals 11, 189.
Sefton-Nash, E., Catling, D.C., 2008. Hematitic concretions at Meridiani Planum, Mars:
Their growth timescale and possible relationship with iron sulfates. Earth and
Planetary Science Letters 269, 365375.
Shaw, S., Pepper, S.E., Bryan, N.D., Livens, F.R., 2005. The kinetics and mechanisms of
goethite and hematite crystallization under alkaline conditions, and the presence of
phosphate. American Mineralogist 90, 18521860.
Shipton, Z.K., Evans, J.P., Robeson, K.R., Forster, C.B., Snelgrove, S., 2002. Structural
heterogeniety and permeability in faulted eolian sandstone: Implications for
subsurface modeling of faults. American Association of Petroleum Geologists
Bulletin 86, 863883.

Souza-Egipsy, V., Orm, J., Bowen, B.B., Chan, M.A., Komatsu, G., 2006. Ultrastructural
study of iron oxide precipitates; implications for the search for biosignatures in the
Meridiani hematite concretions, Mars. Astrobiology 6, 527545.
Spangler, L.E., Naftz, D.L., Peterman, Z.E., 1996. Hydrology, chemical quality, and
characterization of salinity in the Navajo aquifer in and near the Greater Aneth oil
eld, San Juan County, Utah. U. S. Geological Survey, Water-Resources Investigations Report 96-4155. 90 pp.
Tillman, L.E., 1989. Sedimentary facies and reservoir characteristics of the Nugget
Sandstone (Jurassic), Painter Reservoir eld, Uinta County, Wyoming. In: Coalson,
E.B., Kaplan, S.S., Keighin, C.W., Oglesby, C.A., Robinson, J.W. (Eds.), Petrogenesis
and petrophysics of selected sandstone reservoirs of the Rocky Mountain Region.
Rocky Mountain Association of Geologists, Denver, pp. 97108.
Velbel, M.A., 1993. Constancy of silicate-mineral weathering-rate ratios between
natural and experimental weathering: implications for hydrologic control of
differences in absolute rates. Chemical Geology 105, 8999.
Walker, T.R., 1967. Formation of red beds in modern and ancient deserts. Geological
Society of America Bulletin 78, 353368.
Walker, T.R., 1975. Red beds in the western interior of the United States. U. S. Geological
Survey Professional Paper 853, 4956.
Walker, T.R., 1979. A study of global sand seas: Red color in dune sand. In: McKee, E.D. (Ed.),
A study of global sand seas: U. S. Geological Survey Professional Paper, 105, pp. 6181.
White, A.F., Brantley, S.L., 2003. The effect of time on the weathering of silicate
minerals: why do weathering rates differ in the laboratory and eld. Chemical
Geology 202, 479506.
Wilkinson, M., 1991. "The rate of growth of sandstone-hosted calcite concretions": we
have discovered an error in the paper by M. Wilkinson and M. D. Dampier,
published in Geochimica et Cosmochimica Acta 54, 3391-3399. Geochimica et
Cosmochimica Acta 55, 1182.
Wilkinson, M., Dampier, M.D., 1990. The rate of growth of sandstone-hosted calcite
concretions. Geochimica et Cosmochimica Acta 54, 33913399.
Zhu, C., 2005. In situ feldspar dissolution rates in an aquifer. Geochimica et
Cosmochimica Acta 69, 14351453.

You might also like