Mordenskyetal 2016aTRS

Download as pdf or txt
Download as pdf or txt
You are on page 1of 53

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/326988532

Characterization of the Oriskany and Berea sandstones: evaluating


biogeochemical reactions of potential sandstone–hydraulic fracturing fluid
interaction

Technical Report · November 2016


DOI: 10.13140/RG.2.2.34929.79201

CITATIONS READS

3 183

5 authors, including:

Stanley Paul Mordensky Circe Verba


U. S. Geological Survey U.S. Department of Energy
35 PUBLICATIONS 213 CITATIONS 41 PUBLICATIONS 571 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Magmatic volatiles from Cascade shield volcanoes View project

Carbon Storage - Oil Gas Shale View project

All content following this page was uploaded by Stanley Paul Mordensky on 13 August 2018.

The user has requested enhancement of the downloaded file.


Characterization of the Oriskany and
Berea Sandstones: Evaluating
Biogeochemical Reactions of Potential
Sandstone–Hydraulic Fracturing Fluid
Interaction
22 November 2016

Office of Fossil Energy


NETL-TRS-13-2016
Disclaimer
This report was prepared as an account of work sponsored by an agency of the
United States Government. Neither the United States Government nor any agency
thereof, nor any of their employees, makes any warranty, express or implied, or
assumes any legal liability or responsibility for the accuracy, completeness, or
usefulness of any information, apparatus, product, or process disclosed, or
represents that its use would not infringe privately owned rights. Reference
therein to any specific commercial product, process, or service by trade name,
trademark, manufacturer, or otherwise does not necessarily constitute or imply its
endorsement, recommendation, or favoring by the United States Government or
any agency thereof. The views and opinions of authors expressed therein do not
necessarily state or reflect those of the United States Government or any agency
thereof.

Cover Illustration: The cover depicts the top of the inflow sample of Berea sandstone
after guar gum flow-though. The Berea sandstone is coated by a thin (<10 µm thick)
layer of guar gum.

Suggested Citation: Mordensky, S. P.; Rabjohns, K.; Harris, A.; Lieuallen, A. E.; Verba,
C. Characterization of the Oriskany and Berea Sandstones: Evaluating Biogeochemical
Reactions of Potential Sandstone–Hydraulic Fracturing Fluid Interaction; NETL-TRS-
13-2016; NETL Technical Report Series; U.S. Department of Energy, National Energy
Technology Laboratory: Albany, OR, 2016; p 48.

An electronic version of this report can be found at:


http://www.netl.doe.gov/research/on-site-research/publications/featured-technical-
reports
https://edx.netl.doe.gov/ucr
Characterization of the Oriskany and Berea Sandstones: Evaluating
Biogeochemical Reactions of Potential Sandstone–Hydraulic Fracturing Fluid
Interaction

Stanley P. Mordensky1,2, Kelley Rabjohns1,2, Aubrey Harris1,3,


A. Erin Lieuallen1,2, Circe Verba2

1 Oak Ridge Institute for Science and Education, MC-100-44, Oak Ridge, TN 37831
2 U.S. Department of Energy, National Energy Technology Laboratory, 1450 Queen Avenue
SW, Albany, OR 97321
3 U.S. Department of Energy, National Energy Technology Laboratory, 3610 Collins Ferry
Road, Morgantown, WV 26507

NETL-TRS-13-2016

22 November 2016

NETL Contacts:
Daniel Soeder, Principal Investigator
Alexandra Hakala, Technical Portfolio Lead
Cynthia Powell, Executive Director, Research & Innovation Center
This page intentionally left blank.
Characterization of the Oriskany and Berea Sandstones

Table of Contents
EXECUTIVE SUMMARY ...........................................................................................................1
1. INTRODUCTION....................................................................................................................2
1.1MINERAL CHARACTERIZATION ..............................................................................2
1.2CHARACTERIZATION OF THE ORISKANY SANDSTONE, THE BEREA
SANDSTONE, AND THE MARCELLUS SHALE .......................................................4
1.3 HYDRAULIC FRACTURING FLUID CHEMICAL SUMMARY ...............................7
1.4 MINERAL-HYDRAULIC FRACTURING FLUID CHEMICAL RESEARCH .........14
2. METHODS .............................................................................................................................18
3. OBSERVATIONS AND DISCUSSION...............................................................................19
3.1 BEREA SANDSTONE ANALYSIS .............................................................................19
3.2 BEREA-GUAR GUM FLOW-THROUGH ANALYSIS .............................................20
4. CONCLUSION ......................................................................................................................30
5. REFERENCES .......................................................................................................................31

I
Characterization of the Oriskany and Berea Sandstones

This page intentionally left blank.

II
Characterization of the Oriskany and Berea Sandstones

List of Figures
Figure 1: Stratigraphic column (modified from Midwest Regional Carbon Sequestration
Partnership, 2015). .................................................................................................................. 4
Figure 2: Oriskany sandstone outcrop in a U.S. Silica Quarry, Berkeley Springs, WV. Photo
provided by Daniel Soeder and his research team with the National Energy Technology
Laboratory, U.S. Department of Energy. ................................................................................ 5
Figure 3: Berea sandstone core from Cleveland Quarries, Vermillion, Ohio. The Berea core in
this figure appears similar to the Berea core analyzed in this study. Photo provided by
Daniel Soeder and research team with the National Energy Technology Laboratory, U.S.
Department of Energy. ............................................................................................................ 6
Figure 4: Skeletal chemical structure of guar gum. This figure depicts guar gum in skeletal
formula format to communicate the complex molecule as a simple depiction. In a skeletal
formula, carbon atoms are expressed as the end of line segments, and carbon-hydrogen
bonds are assumed to bring carbon’s valence to four. ............................................................ 8
Figure 5: General chemical structure of polyacrylamide. ............................................................... 9
Figure 6: Chemical structure of ethylene glycol. .......................................................................... 10
Figure 7: Chemical structure of poly(diallyldimethylammonium chloride). ................................ 11
Figure 8: Chemical structure of glutaraldehyde............................................................................ 12
Figure 9: Chemical structure of isopropanol. ............................................................................... 13
Figure 10: Aerobic and anaerobic biodegradation of glutaraldehyde. .......................................... 16
Figure 11: Composition of the Berea sandstone (log scale of the relative weight percent). ........ 19
Figure 12: SEM-BSE image of the angular porosity of the Berea sandstone. .............................. 20
Figure 13: SEM-BSE images of post-guar gum inflow and outflow surfaces. ............................ 21
Figure 14: Porosity changes after guar gum flow-through experiment. Unaltered Berea sandstone
has similar porosity throughout its entirety (blue). The modal abundance of guar gum is
higher closer to the inflow surface resulting in lower porosity; as the abundance decreases,
the effective porosity increases across the core (red). .......................................................... 21
Figure 15: SEM-BSE image coupled with EDS of the Berea sandstone inflow section with a dark
coating of guar gum. ............................................................................................................. 23
Figure 16: SEM-BSE image and EDS analysis of the guar gum at the granular level. The guar
gum does not appear to adhere as a thin coating across all minerals’ surfaces. Instead, the
guar gum prefers to stay in contiguous volumes within the Berea sandstone with areas of
unblocked pore space and uncovered minerals. .................................................................... 24
Figure 17: SEM-EDS elemental maps of Berea sandstone inflow surface. The detection of non-
carbon elements through the guar gum coating suggests the guar gum layer is thin (<10
µm). Regions of higher carbon concentrations (guar gum) appear to be accompanied by
aluminum, potassium, and lower silica concentrations, suggesting guar gum preferentially
adheres to clay minerals. Each frame is of the same field of view. ...................................... 25
Figure 18: SEM-BSE image of guar gum filling pore spaces. This image is of a core cut ~6 cm
from the inflow surface. ........................................................................................................ 26
Figure 19: SEM-BSE image of organic carbon and nitrogen. These organic particles (depicted by
darker objects in the images) were less than 6 cm from the inflow surface. ........................ 27
Figure 20: SEM-BSE images and maps coupled with EDS analysis identify organic particles in
Berea outflow sample. .......................................................................................................... 28

III
Characterization of the Oriskany and Berea Sandstones

List of Tables
Table 1: Anaerobic decomposition of ethylene glycol (Veltman et al., 1998) ............................. 11

IV
Characterization of the Oriskany and Berea Sandstones

Acronyms, Abbreviations, and Symbols


Term Description
1,5 Pentanediol HOCH2CH2CH2CH2CH2OH
5-Hydroxypentanal C5H10O2
Acetate CH3COO
Anhydrite CaSO4
Apatite Ca10(PO4)6(OH,F,Cl)2
Barite BaSO4
Bicarbonate HCO3−
BSE Back-scattered electrons
Calcite CaCO3
Carbogen HC
Celestite SrSO4
Chlorite (Mg,Fe2+,Mn,Al)12((Si,Al)8O20)(OH)16
Chromite FeCr2O4
Dolomite CaMg(CO3)2
EDS Energy dispersive x-ray spectroscopy
Ethanol CH3CH2OH
Ethylene glycol HO–CH2CH2–OH
FE-SEM Field emission scanning electron microscope
Feldspar (K,Na,Ca)(AlSi)4O8
Garnet (Mg,Fe,Mn,Ca)3(Al,Fe3+)2(SiO4)3
Glutaraldehyde C5H8O2
Glutaric acid C3H6(COOH)2
Guar gum C12H24O12
Gypsum CaSO4•2H2O
Halite NaCl
Illite (K1.5-1.0Al4(Si6.5-7Al1.5-1.0O20)(OH)4)
Isopropanol C3H8O
Kaolinite Al2Si2O5(OH)4
Kieserite MgSO4•H2O
Magnesite MgCO3
Methane CH4

V
Characterization of the Oriskany and Berea Sandstones

Acronyms, Abbreviations, Symbols (cont.)


Term Description
MNA Monitored natural attenuation
Polyacrylamide CH2CHCONH2
PolyDADMAC Poly(diallyldimethylammonium chloride), (C8H16NCl)n
Pyrite FeS2
Quartz SiO2
Rutile TiO2
SEM Scanning electron microscope or scanning electron microscopy
Siderite FeCO3
Smectite (Ca,Na).2-.4(Al,Mg,Fe)2(Si,Al)4O10(OH)2·nH2O
Strontianite SrCO3
Tourmaline (Na,Ca)(Mg,Fe,Mn,Li,Al)3(Al,Mg,Fe3+)6[Si6O18](BO3)3(O,OH)3(OH,F)
Zircon ZrSiO4

VI
Characterization of the Oriskany and Berea Sandstones

Acknowledgments
This work was completed as part of National Energy Technology Laboratory (NETL) research
for the U.S. Department of Energy’s (DOE) Complementary Research Program under Section
999 of the Energy Policy Act of 2005. The authors wish to acknowledge Ray Boswell (NETL
Strategic Center for Natural Gas and Oil) and Elena Melchert (DOE Office of Fossil Energy) for
programmatic guidance, direction, and support.
The authors would like to thank Keith Collins for helping debug the in-house SEM, and to
everyone on Daniel Soeder’s research team for their involvement with this research, including
Daniel Soeder and Hank Edenborn for their reviews.

VII
Characterization of the Oriskany and Berea Sandstones

This page intentionally left blank.

VIII
Characterization of the Oriskany and Berea Sandstones

EXECUTIVE SUMMARY
The Marcellus shale, located in the mid-Atlantic Appalachian Basin, has been identified as a
source for natural gas and targeted for hydraulic fracturing recovery methods. Hydraulic
fracturing is a technique used by the oil and gas industry to access petroleum reserves in geologic
formations that cannot be accessed with conventional drilling techniques (Capo et al., 2014).
This unconventional technique fractures rock formations that have low permeability by pumping
pressurized hydraulic fracturing fluids into the subsurface. Although the major components of
hydraulic fracturing fluid are water and sand, chemicals, such as recalcitrant biocides and
polyacrylamide, are also used (Frac Focus, 2015).
There is domestic concern that the chemicals could reach groundwater or surface water during
transport, storage, or the fracturing process (Chapman et al., 2012). In the event of a surface spill,
understanding the natural attenuation of the chemicals in hydraulic fracturing fluid, as well as the
physical and chemical properties of the aquifers surrounding the spill site, will help mitigate
potential dangers to drinking water. However, reports on the degradation pathways of these
chemicals are limited in existing literature.
The Appalachian Basin Marcellus shale and its surrounding sandstones host diverse
mineralogical suites. During the hydraulic fracturing process, the hydraulic fracturing fluids
come into contact with variable mineral compositions. The reactions between the fracturing fluid
chemicals and the minerals are very diverse. This report: 1) describes common minerals (e.g.
quartz, clay, pyrite, and carbonates) present in the Marcellus shale, as well as the Oriskany and
Berea sandstones, which are located stratigraphically below and above the Marcellus shale; 2)
summarizes the existing literature of the degradation pathways for common hydraulic fracturing
fluid chemicals [polyacrylamide, ethylene glycol, poly(diallyldimethylammonium chloride),
glutaraldehyde, guar gum, and isopropanol]; 3) reviews the known research about the
interactions between several hydraulic fracturing chemicals [e.g. polyacrylamide, ethylene
glycol, poly(diallyldimethylammonium chloride), and glutaraldehyde] with the minerals (quartz,
clay, pyrite, and carbonates) common to the lithologies of the Marcellus shale and its
surrounding sandstones;1 and 4) characterizes the Berea sandstone and analyzes the physical and
chemical effects of flowing guar gum through a Berea sandstone core.

1
The existing literature for some mineral-hydraulic fracturing fluid pairs was limited.

1
Characterization of the Oriskany and Berea Sandstones

1. INTRODUCTION

1.1 MINERAL CHARACTERIZATION


The Earth consists of thousands of minerals, but only a few constitute the majority of its crust.
This section discusses commonly found minerals in the Berea and Oriskany sandstones and
Marcellus shale. Understanding the mineralogy of a formation is important for recognizing how
minerals influence natural processes and the impact that hydraulic fracturing may have on
formations. Having a fundamental understanding of minerals and their potential reactions with
their surroundings is paramount to understanding fracturing fluids’ natural attenuation processes
and the physical, chemical, and biological processes that reduce the concentration of fracturing
fluids in a natural, subsurface environment (U.S. Geological Survey, 2015). The specific
characterizations of the formations are discussed in greater detail in Section 2.

Quartz
Quartz (SiO2) is the second most abundant mineral in the Earth’s crust and is found in a wide
range of igneous, metamorphic, sedimentary, and hydrothermal settings (Nesse, 2012). The
hardness and lack of cleavage of quartz make the mineral resistant to additional weathering
(Deer et al., 1992). Both the Berea and Oriskany sandstones are primarily composed of quartz.

Clays
Clay is composed of several different phyllosilicate minerals, which are also known as sheet
silicates. Four of the more common types are kaolinite, chlorite, illite, and smectite. The atomic
layering of some phyllosilicates electrostatically attracts liquid H2O, allowing some
phyllosilicates (smectite and illite, but not kaolinite or chlorite) to adsorb higher volumes of
water than other mineral classes, promoting flocculation. Because of their structure,
phyllosilicates contain variable amounts of alkali metals, alkaline earths, iron, and magnesium
(Dana, 2008).
Chlorite [(Mg,Fe2+,Mn,Al)12((Si,Al)8O20)(OH)16], is a common clay phyllosilicate, which forms
primarily in low- to moderate-grade metamorphic rocks. Chlorite is heavily associated with slate,
a common secondary lithology of the Marcellus shale. As the metamorphic grade of a rock
increases, the rock’s chlorite content decreases. Chlorite does not express exceptional water
adsorption tendencies (Dana, 2008).
Kaolinite [Al2Si2O5(OH)4] is another common phyllosilicate clay mineral that forms where its
parent materials, Al-rich minerals such as feldspar and muscovite mica, were deposited under
non-alkaline conditions, like that of the Appalachian Basin. Kaolinite is also formed at shallow
depths and low temperatures in mud rocks, clastic sediments, and coal (Deer et al., 1992).
Kaolinite does not express exceptional water adsorption tendencies (Dana, 2008).
Illite [K1.5-1.0Al4(Si6.5-7Al1.5-1.0O20)(OH)4], another common clay of the Appalachian Basin, is
frequently produced by burial metamorphism from Al-rich minerals, including kaolinite and
smectite (Deer et al., 1992). Illite expresses high water adsorption tendencies (Dana, 2008).
Although a common clay mineral worldwide, smectite is precluded from being a common
mineral in the Appalachian Basin due to the area’s thermal maturity (Deer et al., 1992; Rowan,
2006). The most common mineral in the smectite sub-group is montmorillonite
[(Ca,Na)0.3(Al,Mg)2Si4O10·nH2O]. Much like kaolinite, montmorillonite is derived from Al-rich

2
Characterization of the Oriskany and Berea Sandstones

minerals. Unlike kaolinite, montmorillonite varies in water content (Deer et al., 1992; Perkins,
2011). Smectite expresses high water adsorption tendencies (Dana, 2008).

Pyrite
Pyrite (FeS2), an iron sulfide, is the most common sulfide mineral and is a reagent for the
production of sulfuric acid (H2SO4), which is a common source of high stream acidity in regions
where mining is prevalent, such as the Appalachian Basin. Oxygenated water interacts with
pyrite to oxidize the iron, producing sulfuric acid as expressed in the following equation (Nesse,
2012):

2FeS2(s) + 7O2 + 2H2O → 2Fe2+(aq) + 4SO2−4(aq) + 4H+ (1)

Pyrite is a common mineral in igneous, hydrothermal, and metamorphic rocks. Euxinic


conditions, a common depositional environment of black shale, are also conducive for sulfide
deposition (Deer et al., 1992). In shale and coal, pyrite is often present as fine-grained crystals or
tiny, raspberry-shaped clusters called framboids. Pyrite is also produced by sulfate-reducing
bacteria, converting Fe3+ into Fe2+. In addition, pyrite may contain trace Ni2+ or Co2+ owing to
metal substitution (Nesse, 2012). Some hydraulic fracturing fluids contain chemical compounds
which could interact with pyrite through hydrolysis (Edwards et al., 2014).

Carbonates
Carbonates are formed by CO32- bonding with a cation. The more commonly occurring carbonate
cations are Ca2+, Mg2+, Mn2+, Fe2+, Fe3+, Zn2+, Ba2+, Sr2+, and Pb2+ (Nesse, 2012). Oil shale
classification is based on carbonate composition (Yen and Chilingarian, 1976). Carbonate
minerals tell of their depositional history and are discussed below.
Calcium carbonate (CaCO3) is the most commonly occurring carbonate and naturally crystallizes
as several polymorphs—calcite, aragonite, and vaterite. Calcite forms between clastic sediments
as cement and in other environments as a biomineral, an evaporite, a hydrothermal deposit, and a
metamorphic product. Aragonite and vaterite share the same chemical composition as calcite but
express different crystal atomic lattice structures and stoichiometry due to their higher pressures
of crystallization and the presence of trace enrichments (e.g. Ba, Mg2+, SO42-, Sr) (Berner, 1975;
Walter, 1986; Nesse, 2012). Being the most common composition of carbonate minerals,
calcium carbonates are prolific in the Appalachian Mountains.
Magnesite (MgCO3) is common in Mg-rich metamorphic rocks, but is also frequently found
mixed with CaCO3 minerals in evaporite deposits and as a hydrothermal mineral. Unlike CaCO 3
minerals, magnesite is not a common biomineral but is common to the Appalachian Mountains
(Nesse, 2012).
Dolomite [CaMg(CO3)2] is another common carbonate in the Appalachian Basin. Dolomite can
be considered as the interlayering of calcite and magnesite at a 1:1 ratio. Fe 2+ often replaces
small amounts of Mg2+. Dolomite is primarily a sedimentary rock, but metamorphic and
hydrothermal occurrences are not uncommon (Deer et al., 1992). Several Appalachian
limestones and marbles are composed predominantly of calcite and dolomite (Nesse, 2012). Deer
et al. (1992) suggests that some dolomites are the product of Mg2+ released during the

3
Characterization of the Oriskany and Berea Sandstones

breakdown of smectite to illite. Dolomite precipitation is favored in solutions with low


Ca2+/Mg2+ and Ca2+/CO32- ratios and high temperatures (Deer et al., 1992).
Siderite (FeCO3) is a frequent trace carbonate in metamorphic rocks of sedimentary protoliths.
Although Zielinski and McIver (1982) report the Marcellus shale to contain ~0% siderite,
specific localities and members of the Marcellus have higher siderite concentrations; Soeder
(2014) observed abundant siderite concretions up to tenths of a centimeter in size in the Oatka
Creek Member of the Marcellus shale. Siderite is rarely found in pure form and, instead, contains
low concentrations of Mn2+ and Mg2+ (Deer et al., 1992).

1.2 CHARACTERIZATION OF THE ORISKANY SANDSTONE, THE BEREA


SANDSTONE, AND THE MARCELLUS SHALE
The Oriskany and Berea sandstone formations are located below and above the Appalachian
Basin Devonian shales, which include the Marcellus shale (Figure 1), and are important
stratigraphic units to understand in the context of hydraulic fracturing because of their proximity
to the Marcellus shale. The Oriskany and Berea sandstones could be exposed to hydraulic
fracturing fluids or provide means for hydraulic fracturing fluids to reach aquifers. Surface spills
also pose a contamination hazard. Understanding the natural attenuation of the chemicals in
hydraulic fracturing fluid and the physical and chemical properties of the aquifers surrounding a
site of interest would help mitigate potential danger to human populations and the natural
environment.

Figure 1: Stratigraphic column (modified from Midwest Regional Carbon Sequestration


Partnership, 2015).

4
Characterization of the Oriskany and Berea Sandstones

Oriskany Sandstone
The Oriskany sandstone (Figure 2) is a thoroughly studied geologic formation because of its
proximity to the Marcellus shale and its use as a source for sand glass (Heinrich, 1981). The
Oriskany sandstone interbeds with limestone, chert, and shale lithologies. Limestones are
sedimentary rocks composed of calcium carbonate—specifically calcite and aragonite. Cherts are
fine-grained sedimentary rocks that are silica rich but also incorporate surrounding minerals and
elements into their structure. Shales are very fine mudstones composed mainly of clay minerals
with some organic matter, iron, sulfide, and heavy mineral substitution. Common minerals found
in these formations include quartz, calcium carbonates, kaolinite, illite, and chlorite (Dana,
2008).
Porosity throughout the Oriskany varies, averaging 4–7%, which is relatively low compared to
the ~20% porosity of the majority of sandstones (Kostelnik and Carter, 2009). The Oriskany
sandstone ranges from a fine- to medium- to coarse-grained sandstone with a cementing clay.
The grains also tend to appear relatively weathered, rounded, and well sorted.

Figure 2: Oriskany sandstone outcrop in a U.S. Silica Quarry, Berkeley Springs, WV. Photo
provided by Daniel Soeder and his research team with the National Energy Technology
Laboratory, U.S. Department of Energy.

Berea Sandstone
The Berea sandstone (Figure 3) is also a thoroughly studied geologic formation because of its
proximity to Appalachian Basin Devonian shales, its use as a test material for the petroleum
industry, and its role as a resistant building material. Common minerals in the Berea sandstone
are similar to the minerals found in the Oriskany, with quartz being the major mineral. Other
minor components in the Berea sandstone include feldspar [(K,Na,Ca)(Al,Si) 4O8] and chert, as
well as heavy minerals, such as tourmaline 2, zircon (ZrSiO4), rutile (TiO2), chromite (FeCr2O4),

2
(Na,Ca)(Mg,Fe,Mn,Li,Al)3(Al,Mg,Fe3+)6[Si6O18](BO3)3(O,OH)3(OH,F)

5
Characterization of the Oriskany and Berea Sandstones

apatite [Ca10(PO4)6(OH,F,Cl)2], garnet [(Mg,Fe,Mn,Ca)3(Al,Fe3+)2(SiO4)3], and barite (BaSO4)


(Pepper et al., 1954; Deer et al., 1992).
The Berea is fine-grained, clay-cemented quartz sandstone that is valuable for industrial
purposes. While the chemical composition of the Berea sandstone is similar to the Oriskany
sandstone, the Berea sandstone has higher porosity and permeability on average than the
Oriskany sandstone. Average porosity and permeability values for the Berea are 21% and 165
mD (Jagucki and Darner, 2001). Groundwater in the Berea has a residence of ~2,500 years, an
average pH of 7–8, and a dissolved oxygen concentration of 0–2 mg/L (Jagucki and Darner,
2001). In the field, the Berea seems coarser grained than it actually is because aggregates of fine
grains are cemented together and appear as coarse grains. Field descriptions of the Berea vary
from a medium-grained sandstone to a fine-grained pebbly sandstone to a coarse-grained pebbly
siltstone with a cementing clay. The Berea sandstone encloses lenses of coarse pebble-like lithics
composed of red shale, gray shale, black shale, and fine-grained steel-grey siltstone (Pepper et
al., 1954).

Figure 3: Berea sandstone core from Cleveland Quarries, Vermillion, Ohio. The Berea core
in this figure appears similar to the Berea core analyzed in this study. Photo provided by
Daniel Soeder and research team with the National Energy Technology Laboratory, U.S.
Department of Energy.

Marcellus Formation
The Marcellus Formation is a Devonian black shale in the Lower Hamilton group composed of
organic, terrigenous sediments that accumulated in a trough that was formed from the
impingement of Laurentia during the Acadian Orogeny that would later become the Appalachian
Basin (Woodrow, 1985; Blauch et al., 2009; Enomoto et al., 2011). The Marcellus shale formed
during sea-level fluctuations between shallowing and deepening conditions, sometimes under
euxinic, H2S-rich conditions (Sageman et al., 2003). Changes in sea level and climate resulted in
saline evaporates interbedding with organic terrigenous sediments (Werne et al., 2002; Enomoto
et al., 2011).

6
Characterization of the Oriskany and Berea Sandstones

Shale is naturally heterogeneous. The Marcellus shale is no exception and hosts distinct
lithofacies. One of the most important of the lithofacies is the argillaceous facies, which is
dominated by quartz and clay minerals with minor amounts of halite (NaCl), calcite, mica, and
pyrite. Boyce and Carr (2009) report high quartz contents (60%), relatively low illite-muscovite
clay contents (30%), and pyrite contents between 5–10%. The other important lithofacies, the
calcareous-argillaceous facies, are rich with calcite and contain only minor amounts of quartz.
The trace minerals are barite, strontiantite (SrCO 3), celestite (SrSO4), siderite (FeCO3), gypsum
(CaSO4·2H2O), anhydrite (CaSO4), and kieserite (MgSO4·H2O). The Marcellus also contains
other minor lithofacies like the Purcell limestone, which is a fine grained, thin-bedded layer
found within the Marcellus (Boyce and Carr, 2009). Other mineral features of the Marcellus
include a saturation of halite and dolomitization from the saline evaporite formation (Bruner and
Smosna, 2011).
The fluids involved with shale formation have an important effect on the mineralization during
diagenesis. As Marcellus sediments were deposited and lithified, salts and fluids from the
surrounding brine became incorporated into the rock structure. Once these pore fluids were
mobilized, salt dissolution resulted in subsequent autochthonous crystallization within the rock’s
pore space (Blauch et al., 2009). Dresel and Rose (2010) found that oil field brines from the
Devonian shales in Pennsylvania varied in concentration from 9.99 to 343 g/L total dissolved
solids.
Organic-rich sediments heated in the Earth's crust produce kerogen, an insoluble organic
mixture. Heating kerogen creates hydrocarbon chains, which vary in length and complexity
depending on the temperatures to which the kerogen is exposed. Less heat equates to longer,
more complex hydrocarbon chains (oil), but as heating increases, the chains become
progressively shorter and simpler, producing wet gas and dry gas (Selley, 1997; McCarthy et al.,
2011). The thermal maturity of the Marcellus shale makes wet and dry common hydrocarbon
deposits (Repetski et al., 2008; Bruner and Smosna, 2011). Total organic carbon varies within
the Marcellus Formation. Average values range from 2–5%, but can be as high as 12–20%
(Nyahay et al., 2007; Laughrey et al., 2011; Wang and Carr, 2013). Ziemkiewicz et al. (2013)
reported dissolved oxygen concentrations in Marcellus Formation groundwater between 3.24–
13.12 mg/L and pH values of 6.32–8.75. As with the Berea and Oriskany sandstones, porosity
also varies throughout the Marcellus. However, the porosity of shales is significantly lower than
that of most sandstones. Laughrey et al. (2011) reported porosity between 2.44–16.6% of bulk
volume and 2–4% organic porosity. They measured permeability of 0.047 to 0. 217 µD.
Comparatively, Soeder (1988) found Marcellus porosity of 9.28% and of 19.613 µD. Estimates
of porosity by the U.S. Department of Environmental Conservation (Leff, 2011) for Devonian
shales in the Appalachian Basin are a tighter range of 1–3%.

1.3 HYDRAULIC FRACTURING FLUID CHEMICAL SUMMARY


Organic compounds in hydraulic fracturing fluids will degrade naturally through both
microbiological and geochemical processes in groundwater, a process referred to as natural
attenuation. Natural attenuation is utilized as a common remediation technique for groundwater
contamination, and is referred to as monitored natural attenuation (MNA). For MNA to be
effective, the breakdown processes and rates for the various chemicals must be understood. If
MNA is shown to be too slow to prevent the contaminant from reaching the accessible
environment, cleanup can be accelerated by the use of enhanced biodegradation, reactive

7
Characterization of the Oriskany and Berea Sandstones

barriers, pump-and-treat, or other processes. This section reviews existing research for the
natural attenuation of common hydraulic fracturing fluids.

Guar Gum

HO HO

HO

HO

O HO

O HO OH
O
O HO O
OH
OH

Figure 4: Skeletal chemical structure of guar gum. This figure depicts guar gum in skeletal
formula format to communicate the complex molecule as a simple depiction. In a skeletal
formula, carbon atoms are expressed as the end of line segments, and carbon-hydrogen
bonds are assumed to bring carbon’s valence to four.

Guar gum (C12H24O12) (Figure 4) is a thickening agent used for carrying proppants into hydraulic
fractures. A proppant is a solid material (usually sand) in hydraulic fracturing fluid used to prop
open the fractures created by hydraulic pressure, maintaining open flow paths after the pressure
has been released (Frac Focus, 2015). Most of the previous research related to guar gum focuses
on guar gum breakdown in the digestive tract, because guar gum is used as a thickening agent in
many food products such as ice cream. It has not been thoroughly studied because it is non-toxic
to humans, but it still has the potential to chemically and physically interact with microbes and
other chemicals. Insoluble residues like protein and cellulose are associated with guar gum and
can reduce the flow of gas out of the well after the guar gum degrades. Solutions below pH 5 and
above pH 7 can cause hydrolysis of guar gum. Fermentation of guar gum produces short-chain
fatty acids and gases such as carbon dioxide, hydrogen, and methane (Mali et al., 2012).
Salyers et al. (1977 and 1978) studied the degradation of guar gum by anaerobic
microorganisms. They found that Bacteroides ovatus, Bacteroides ‘0061-1,’ and Bacteroides
vulgatus ferment guar gum. Additionally, they found that the enzymes responsible for the
degradation of the guar gum were extracellular and inducible. Extracellular enzymes are attached
on the outside of the cell but are still part of the organism. Complex and bulky substrates are able
to react with extracellular enzymes without having to move into the interior of the cell. Inducible
enzymes are adaptable and can be turned on or produced when the substrate is present. The end
products of fermentation will depend on the microorganism that is biodegrading the guar gum,
but the most common end products are acetate, butyrate, propionate, and lactate. Hartemink et al.
(1999) discovered that B. ovatus produces mainly acetate and propionate; Ruminococcus albus
produces acetate and a small amount of butyrate; Bifidobacterium adolescentis produces acetate

8
Characterization of the Oriskany and Berea Sandstones

and lactate; Clostridium butyricum produces butyrate; Clostridium coccoides produces succinate;
and Ruminococcus productus produces acetate.

Polyacrylamide

Figure 5: General chemical structure of polyacrylamide.

Polyacrylamide (–CH2CHCONH2–) (Figure 5) is a synthetic, industrial polymer that decreases


the friction of the hydraulic fracturing fluids and is sometimes referred to as “slickwater” (Frac
Focus, 2015). While polyacrylamide itself is non-toxic, acrylamide monomers created during
breakdown are toxic to peripheral nerves. This is not unusual in the degradation process; many
organic chemicals produce daughter products or “metabolites” that are more toxic than the
original compound. A complete understanding of the breakdown path from start to finish is
required to assess if MNA is a viable remediation option for any particular compound.
Acrylamide monomers are produced from abiotic degradation of polyacrylamide by heat or
ultraviolet irradiation, so the degradation of polyacrylamide is an important chemical pathway to
study. Wen et al. (2010) isolated polyacrylamide-metabolizing bacteria, Bacillus cereus and
Bacillus flexu, and identified polyacrylamide as their sole source of carbon. The addition of
glucose to the substrate at low concentration (below 200 mg/L) increased the rate of
biodegradation and can be co-metabolized with polyacrylamide. Conventional carbon sources
(i.e. glucose) at lower concentrations stimulate growth which leads to more enzymes to
breakdown toxic compounds; however the microbes will prefer to metabolize the additional
organic carbon over the toxic compounds which leads to a decrease in biodegradation (Wen, et
al., 2010). The highest rate of biodegradation that Wen et al. (2010) observed was roughly 70%
over 96 hours of cultivation at low initial concentrations of polyacrylamide. However, with an
initial concentration of 500 mg/L, B. cereus degrades roughly 50% over 96 hours, and B. flexu
only degrades roughly 30% over 96 hours.
In addition to Wen et al. (2010), other studies have focused on other organisms’ effects on the
degradation of polyacrylamide. Stahl et al. (2000) studied the white-rot fungi, Phanerochaete
chrysosporium, which degrades lignin along with a variety of environmental toxins, including
polyacrylamide. The enzymes responsible for extracellular degradation catalyze polyacrylamide
in manganese and iron redox reactions. Adding sawdust and soil microbes to the white-rot
fungus optimized biodegradation conditions. Solubilization of the polyacrylamide was the rate-
limiting step in this biodegradation experiment (Stahl et al., 2000). Nakamiya and Kinoshita
(1995) isolated bacteria (Enterobacter agglomerans and Azomonas macrocytogenes) that
metabolized polyacrylamide using a substrate of 10 mg/mL of polyacrylamide as their sole
carbon and nitrogen source. E. agglomerans is an anaerobic bacterium, while A. macrocytogenes
is an aerobic bacterium. Both bacteria only degraded a proportion of initial polyacrylamide at

9
Characterization of the Oriskany and Berea Sandstones

10 g/L (utilized to available polyacrylamide). However, when comparing real commercial


polyacrylamide to the lab-synthesized polyacrylamide, the growth of the bacteria was delayed,
and the degradation ratio was almost half of the rates seen with the lab polyacrylamide.
Kay-Shoemake et al. (1998) found microbes utilizing polyacrylamides as a nitrogen source.
Wang and Lee (2001) isolated denitrifying bacteria, Pseudomonas stutzeri, that use acrylamide
as a substrate for metabolism and observed the degradation of acrylamide under aerobic and
anaerobic conditions. P. stutzeri produced ammonia and acrylic acid as intermediates. Under
aerobic conditions, P. stutzeri metabolized acrylamide at different efficiencies, depending on the
concentration. Initial concentrations up to 148.9 mg/L of acrylamide were 91% removed after 5.9
hours and 100% removed after 10 hours. Initial concentrations up to 440.1 mg/L of acrylamide
were 53% removed after 6 hours and 80% removed after 14.1 hours. Ammonia and acrylic acid
were intermediate products of acrylamide biodegradation and act as carbon and nitrogen sources
under aerobic conditions. Under anaerobic conditions, P. stutzeri metabolized 100% of 148.9
mg/L acrylamide with 147.3 mg/L nitrate after 15.8 hours. Wang and Lee (2001) found that,
“With the existence of adequate electron acceptors, the P. stutzeri could utilize both acrylamide
and acrylic acid as the substrates undergoing denitrification, and acrylamide could be removed
completely from the wastewater.”

Ethylene Glycol

Figure 6: Chemical structure of ethylene glycol.

Ethylene glycol (HO–CH2CH2–OH) (Figure 6) compounds are used in addition to pH stabilizers


to inhibit CO2 corrosion of the drilling infrastructure (Frac Focus, 2015). Ethylene glycol
prevents the formation of hydrates, which can form when the pH stabilizers increase the pH of
the solution to inhibit corrosion. The anaerobic degradation of ethylene glycol, using nitrate as an
electron acceptor, produces methane and carbon dioxide as end products. Intermediate products
include acetaldehyde, ethanol, acetate, and bicarbonate (Dwyer, 1986 Schramm and Schink,
1991; Huang et al., 2005). The degradation pathway of ethylene glycol under anaerobic
methanogenic conditions is shown in Table 1 (Veltman et al., 1998).
Gruden et al. (2001) observed glycolic acid and oxalic acid as intermediate products and carbon
dioxide as the final end product of anaerobic degradation of ethylene glycol. Dwyer et al. (1986)
isolated two anaerobic bacteria from methanogenic cultures that metabolized polyethylene
glycols. One of the bacteria, Desulfovibrio desulfuricans, explicitly metabolized ethylene glycol.
The other bacteria, Bacteroides (strain PG1) metabolized diethylene glycol and polymers of
polyethylene glycol but could not metabolize ethylene glycol. Both bacteria used polyethylene
glycol as the sole carbon source, and both grew well in coculture with sulfate and methanogenic

10
Characterization of the Oriskany and Berea Sandstones

bacteria acting as an electron acceptor producing acetate, ethanol, and hydrogen as end products
with acetaldehyde as an intermediate product. The substrates that did not support growth of the
bacteria and fermentation of polyethylene glycols were glycolate, glyoxylate, glycerol,
ethanolamine, polyoxyethylene sorbitan monolaurate, polyoxyethylene 23-lauryl ether, and
polypropylene glycols. Additionally, de Ory et al. (1998) found that the bacteria, Acetobacter
aceti, metabolizes ethanol to acetate, and after all of the ethanol is converted, they metabolize the
available CO2.

Table 1: Anaerobic decomposition of ethylene glycol (Veltman et al., 1998)

Reaction Chemical Equation ΔG (Kcal)


Ethylene Glycol  Acetate + HOCH2CH2OH  0.5 CH3COO− + 0.5
-21.7
Ethanol HC + 0.5 CH3CH2OH + 0.5 H2O
CH3CH2OH + H2O  CH3COO− + HC
Ethanol  Acetate 2.3
+ 2 H2
HOCH2CH2OH  CH3COO− + HC +
Ethylene Glycol  Acetate -20.6
H2
Acetate  Methane CH3COO− + H2O  HCO3− + CH4 -7.4
Hydrogen  Methane 4 H2 + HC + HCO3−  CH4 + 3H2O -32.4
HOCH2CH2OH + 0.25 H2O  1.25
Ethylene Glycol  Methane -36.1
CH4 + 0.75 HC + 0.75HCO3-

Poly(Diallyldimethylammonium Chloride)

H 2C C H2

HC CH

H 2C C H2
+ –
N Cl
H3C C H3

Figure 7: Chemical structure of poly(diallyldimethylammonium chloride).

Poly(diallyldimethylammonium chloride) (polyDADMAC) (C8H16NCl)n (Figure 7) is a polymer


used for flocculation in hydraulic fracturing processes (Morrison and Boyd, 1987).
PolyDADMAC falls within the larger chemical category of quaternary ammonium compounds.
Morrison and Boyd (1987) studied quaternary ammonium compounds’ reactions to different pH
conditions and found quaternary ammonium compounds to be largely unaffected by pH changes
as ammonium compounds have no protons to give and are unaffected by OH -. However, at high
pH extremes, a quaternary ion may form a hydroxide that can undergo a Hoffman elimination
reaction.

11
Characterization of the Oriskany and Berea Sandstones

John (2008) used a novel approach—gel permeation chromatography—to study


polyDADMAC’s reactivity under varied temperature, pH, and microorganism conditions.
PolyDADMAC appears to be a stable polymer resistant to a range of varying environmental
conditions. John (2008) subjected polyDADMAC to a range of temperatures from ambient to
80°C. Between ambient temperatures and 70°C, no structural changes to the polymer were
observed. Between 70°C and 80°C, gel permeation chromatography displayed evidence of
scission between the long and short chains, but new product formation was not evident.
However, in the same study, John (2008) adjusted the polyDADMAC solution (0.5% m/v) pH by
adding phosphoric acid (10% v/v) and sodium hydroxide (0.1 N). The pH conditions were held
for 1 hour. Analysis revealed polyDADMAC to be resistant to pH conditions ranging 2–12.
Acidic and moderately basic solutions displayed no evidence of new products. Highly basic
solutions (> pH 12) displayed some evidence of new product, but concentrations were too low
for identification.
John (2008) inoculated two aliquots of 0.02% polyDADMAC solution—one with 200 μL of
Candida albicaris (yeast) and the other with a 25 mL aliquot with Bacillus subtilis (bacteria).
These two solutions were incubated at 30°C for a minimum of 48 hours. Analyses from the C.
albicaris and B. subtilis cultures indicated some biodegradation, but John (2008) does not
specify a concentration or composition of the new product. John (2008) also attempted similar
tests with Escherichia coli and coliforms, but was unsuccessful due to the death of the cultures.

Glutaraldehyde

Figure 8: Chemical structure of glutaraldehyde.

Glutaraldehyde (C5H8O2) (Figure 8) is the most common electrophilic biocide used during
hydraulic fracturing and is frequently used as a disinfectant in the medical field (Kahrilas et al.,
2015; Frac Focus, 2015). Although the literature regarding glutaraldehyde’s interaction with
minerals is virtually non-existent, glutaraldehyde’s degradation processes are well studied.
Glutaraldehyde functions by damaging cellular walls of the organisms to which it is exposed
(Kahrilas et al., 2015).
Leung (2001a) summarizes numerous glutaraldehyde toxicity studies. The median effective
concentration for glutaraldehyde (EC50) was found to be 25 mg/L for sewage bacteria, 17 mg/L
for Polyseed® (Polybac Corp.) bacteria, and 0.78 mg/L for oysters (Kahrilas et al., 2015). The
No Observed Effect Concentration was found to be 5 mg/L, above which glutaraldehyde
inhibited microbial behavior over prolonged exposure. Emmanuel et al. (2005) studied the toxic
effects of glutaraldehyde when in solution with three commonly used hospital surfactants
(cetyltrimethylammonium bromide, sodium dodecyl sulfonate, and Triton X-100) on Vibrio
fischeri and Daphnia, but mixing did not increase impact. Glutaraldehyde appears resilient to
abiotic degradation. A 50-mL, 2-mg/L sample of glutaraldehyde displayed no degradation in the

12
Characterization of the Oriskany and Berea Sandstones

dark at 20°C after 28 days. However, 50°C produced ~8% degradation after 14 days (Leung,
2001a).
Glutaraldehyde is not strongly susceptible to hydrolysis in acidic or neutral pH environments, but
it is unstable in alkaline environments. At 25°C, 50% degradation took 508 days in a pH 2
environment, 102 days in a pH 7 environment, and 46 days in a pH 9 environment. Leung
(2001a) identified the degradation product as a cyclized dimer of glutaraldehyde (3-formyl-6-
hydroxy-2-cyclohexene-1-propanal). Glutaraldehyde can be deactivated by two methods. In the
first, concentrations up to 5% glutaraldehyde can be deactivated by adding 2–3 molar parts
sodium bisulfate (NaHSO3) to the solution. Deactivation is complete after 10 minutes. In the
second method, NaOH, KOH, or dibasic ammonium phosphate is added to a pH 12 solution with
glutaraldehyde. The glutaraldehyde becomes deactivated in ~8 hours and produces a cyclicized
dimer (3-formyl-6-hydroxy-2-cyclohexene-1-propanal). As the acidity increases, deactivation
time increases (Leung, 2001b).

Isopropanol

Figure 9: Chemical structure of isopropanol.

Isopropanol (C3H8O) (Figure 9) inhibits corrosion and increases viscosity of the fluid in
hydraulic fracturing (Frac Focus, 2015). It is widely used for industrial purposes and can be
oxidized to acetone, which can be metabolized by bacteria. Biodegradation has been observed
under anaerobic conditions (Ueyama et al., 1971; Fox and Ketha, 1996), at low concentrations
(Murrays et al., 1980; Kanemitu et al., 1980), and at high concentrations (Inoue and Horikoshi,
1989; Moriya and Horikoshi, 1993). Isopropanol can be toxic to some microorganisms but can
be the main carbon source for denitrifying bacteria (Hwang et al., 1995). Quinlan et al. (1999)
described the isopropanol-degrading bacteria’s substrate preferences, specific growth rate, and
solvent-tolerance. Siegel and Kamen (1950) studied the anaerobic microbe Rhodopseudomonas
gelatinosa that metabolized isopropanol, acetone, n-butanol, n-hexanol, ethylmethyl ketone,
acetic acid, propionic acid, butyric acid, and acetoacetic acid. Acetone was produced as an
intermediate product of isopropanol degradation, accumulated, and metabolized by R. gelatinosa.
The growth rate of R. gelatinosa increased with the addition of 0.3% Difco peptone and 200
mg/L of dehydrated yeast.
Bustard et al. (2001) studied the biodegradation of isopropanol by a consortium of bacteria in an
aerobic three-phase fixed bed bioreactor. Bustard et al. (2001) claimed that a consortium of a
mixed population of microorganisms is more successful at biodegradation than an isolated
microorganism. The consortium was cultivated by enriching a sample from an oil waste sump
with hydrocarbons and isopropanol. This allowed the isopropanol-biodegrading microorganisms
to thrive. The consortium degraded 96% of the initial concentration of isopropanol (0.785 g/L)

13
Characterization of the Oriskany and Berea Sandstones

after 80 hours of treatment. The rate of biodegradation fluctuated between 3.9 x 10 -3 g/L/h and
300 x 10-3 g/L/h with an average degradation rate of 80 x 10-3 g/L/h. While Bustard et al. (2001)
did not isolate and identify the specific bacteria in the consortium, they did observe that the
degradation pathway is similar to the degradation mechanisms of Xanothobacter sp. and
Rhodococcus rhodachrous.

1.4 MINERAL-HYDRAULIC FRACTURING FLUID CHEMICAL RESEARCH


Hydraulic fracturing fluids emplaced at depth are introduced to the mineralogies of the rock units
targeted for hydraulic fracturing. Loss of hydraulic fracturing fluids at the surface can also lead
to groundwater infiltration and affect shallow aquifers. Understanding the reactions and
interaction between the hydraulic fracturing fluids and the mineralogy is key to the hydraulic
fracturing fluids’ natural attenuation in a subsurface environment. The literature surrounding
these interactions does not cover every hydraulic fracturing fluid and commonly occurring
mineral. This report summarizes the literature available that describes these interactions.

Polyacrylamide
Polyacrylamide—Quartz
Quartz-polyacrylamide research focuses on the adsorption of polyacrylamide onto the crystal’s
surface. However, whether non-ionic polyacrylamide adsorbs to quartz crystals has not yet been
determined (Samoshina et al., 2003). Lecourtier et al. (1990) found evidence supporting non-
ionic polyacrylamide adsorption to be temperature-based. Bjelopavlic et al. (2000) postulated
that the surface charge of quartz depends on too many parameters to allow for consistent non-
ionic polyacrylamide adsorption. Guevellou et al. (1995) reported more adsorption with anionic
than non-ionic acrylamide. Samishona et al. (2003) found cationic polyacrylamide to adsorb well
(.8-1.8 Γ mg/m2) on quartz crystals, whereas anionic polyacrylamide adsorbed, but to a lesser
extent (0.1–0.3 Γ mg/m2). Guevellou et al. (1995) found aluminate—a compound containing
oxygen, aluminum, and other metals—to precipitate out of a cationic polyacrylamide-sand-
kaolinite-aluminate mixture after 5–10 days. Precipitation did not occur when an anionic
polyacrylamide was used in place of the cationic polyacrylamide (Guevellou et al., 1995).
Polyacrylamide—Clay
Kaolinite adsorbs polyacrylamide by a factor of three more than quartz. Graveling et al. (1990)
postulates that kaolinite’s greater adsorption characteristic is a product of the mineral’s basal
planes. Tekin et al. (2005) suggested kaolinite adsorption of cationic polyacrylamide increases
with pH, temperature, and ionic strength. The increased adsorption ability of kaolinite allows
polyacrylamide to serve as an electrostatic stabilizer for kaolinite (Tekin et al., 2005).
Polyacrylamide causes clays to increase in volume, strength, pore space, and permeability
(Wallace and Wallace, 1986). When serving as a flocculent, cationic polyacrylamides create a
larger floc size than anionic polyacrylamides. However, increasing the charge for cationic and
anionic polyacrylamides by between 10–35% decreases the floc size (Nasser and James, 2007).
Polyacrylamide—Carbonates
Peng and Garnier (2012) demonstrated that the floc size of calcium carbonates increases as a
function of the cationic polyacrylamide charge density. Smaller polyacrylamide chains express a
higher charge density. When NaCl was added to the mixture, this relationship was reversed.

14
Characterization of the Oriskany and Berea Sandstones

Wang et al. (2006) precipitated aragonite from a Ca(OH)2 and CO2/N2 gas mixture in a
polyacrylamide substrate as expressed in the following equation.

Ca(OH)2 + CO2 → CaCO3 + H2O (2)

Polyacrylamide—Pyrite
The literature on pyrite-polyacrylamide interaction is sparse, which is surprising given pyrite’s
common occurrence in coal extraction and other industrial processes. Boulton et al. (2001)
developed a method using low-weight polyacrylamides to separate sphalerite (ZnS), a zinc ore,
from pyrite.

Ethylene Glycol
Ethylene Glycol—Clay
Clays do not appear chemically active to ethylene glycol at surface temperatures and pressures.
Hosterman and Whitlow (1983) found no evidence of chemical change in clays that were mixed
with ethylene glycol and raised to a temperature of 350°C for 30 minutes. An XRD sample
preparation manual advises to apply ethylene glycol on illite and smectite to induce swelling for
a better analysis (Poppe et al., 2002). This practice suggests ethylene glycol and clays are not
chemically active on a short-term timescale. Kaolinite and chlorite do not swell from ethylene
glycol contact (Hosterman and Whitlow, 1983).
Ethylene Glycol—Carbonates
Ethylene glycol’s role in preventing calcium carbonate buildup in industrial applications has
helped foster ethylene glycol-carbonate research. Flaten (2009) studied the effect of ethylene
concentration on calcium carbonate precipitation. Solutions with high ethylene glycol
concentrations produced vaterite, while higher temperature solutions of the same composition
produced aragonite. Flaten (2009) postulated that ethylene glycol extends the life of metastable
polymorphs (e.g. vaterite) by delaying the growth of more stable polymorphs (e.g. calcite).
Flaten et al. (2010) studied the relationship between the induction time of calcium carbonate
precipitation and ethylene glycol. At 90 weight percent ethylene glycol and 25°C, CaCO 3
required an induction time of 12 days. Increasing the ethylene glycol content increased the
required induction time. Maintaining constant ethylene glycol concentrations while lowering
temperatures also increased the calcium carbonate induction time. Similarly, increasing the
ethylene glycol concentration increased the nucleation rate but decreased the crystal growth rate
(Konopacka-Konopacka-Łyskawa and Lackowski, 2011). These effects are likely due to CO 2’s
decreased solubility in ethylene glycol (Hayduk and Malik, 1971).
Ethylene Glycol—Pyrite
Ethylene glycol-pyrite reactions do not receive the same attention as ethylene glycol-calcium
carbonate reactions. Mahajan et al. (2007) added ethylene glycol to hydrogen peroxide and
chalcopyrite to stabilize the hydrogen peroxide without altering the chalcopyrite morphology.
This resulted in increased Cu2+ dissolution and leaching efficiency. In another study, Wang et al.
(2011) used ethylene glycol (after homogenized iron chloride hexahydrate and
dimethyformamide solution) to assist with the precipitation of nanoflakes-built pyrite

15
Characterization of the Oriskany and Berea Sandstones

microspheres and argued that this method could be used to precipitate other chalcogenides, such
as PbS and CuxS.

Glutaraldehyde
Glutaraldehyde—Sediments
Leung (2001b) studied the biodegradation of glutaraldehyde when added to water and sediments
from the Sacramento River. Under aerobic degradation, glutaraldehyde had a half-life of
10.6 hours and was completely metabolized after 48 hours. Glutaraldehyde degraded to an
intermediate-stage glutaric acid before being completely metabolized to CO 2 (Figure 25). Under
anaerobic degradation, glutaraldehyde had a half-life of 7.7 hours and completely metabolized
after 72 hours. Glutaraldehyde degraded to an intermediate-stage 5-hydroxypentanal before
being completely metabolized to 1,5-pentanediol (Figure 10).

Aerobic Biodegradation

Glutaraldehyde Glutaric Acid


Anaerobic Biodegradation

Glutaraldehyde 5-Hydroxypentanal 1,5-Pentanediol


Figure 10: Aerobic and anaerobic biodegradation of glutaraldehyde.

Glutaraldehyde autopolymerizes via aldol condensation into α,β-unsaturated polymers that are
not classified as toxic (Kahrilas et al., 2015). Landrum et al. (2003) observed that higher
temperatures (5°C vs. 15°C vs. 25°C), higher concentrations of microorganisms, and lower
concentrations of glutaraldehyde accelerated the rate of biotic glutaraldehyde degradation, but
they found no difference in abiotic degradation rates between water-only and water-sediment
samples. At and below 4°C, biotic degradation appeared to effectively cease. Landrum et al.
(2003) found glutaraldehyde’s abiotic degradation over 28 days to be inconsequential at
temperatures below 25°C and did not run samples above 25°C.

Poly(diallyldimethylammonium chloride)
This literature review found no information involving specific quartz, clay, pyrite, or carbonate
interaction.

16
Characterization of the Oriskany and Berea Sandstones

Natural Attenuation Summary


The natural attenuation literature regarding hydraulic fracturing fluids suggest that understanding
the mineral suites, bacterial presence, and physical conditions of a spill site are just as important
as knowing which chemicals are present at the spill. The degradation pathways of hydraulic
fracturing fluids are complex and highly dependent on the pH, redox state, mineral compositions,
and bacterial species in the soil and shallow aquifers that may be affected by a spill. Additional
research on specific degradation pathways involving different organic chemicals and fluid
compositions would aid in increasing the understanding of potential environmental effects of
hydraulic fracturing in the Appalachian Basin. This study of the interaction between guar gum
and the Berea sandstone is one such attempt to observe how hydraulic fracturing fluid chemicals
may interact with the lithologies of the Appalachian Basin.

17
Characterization of the Oriskany and Berea Sandstones

2. METHODS
The experiment in this study isolates reactions caused by one additive in hydraulic fracturing
fluid: guar gum. Guar gum was chosen for the initial natural attention study because it is
extremely common in hydraulic fracturing operations and may be a strong contributor to
microbial reactions. Guar gum is used as a gelling agent to suspend proppant transported to the
fractures at depth. This organic chemical is used in food products and is easily degraded by
biological processes.
The guar gum was mixed with tap water to a concentration of 3 mg/mL before introduction to
the Berea core from Cleveland Quarries, Vermillion, Ohio as described in Section 1.2.2 (13 cm
long, 5 cm in diameter). The Berea core, enveloped in a sleeve, was placed under confining
pressure of 500 psi and connected to an Isco syringe-pump, which forced fluid through at a
constant rate of 0.04 mL/min. The confining pressure on the sleeve forced the guar gum to flow
through the core rather than around the outside. Effluent was collected at the downstream end.
The experiment was performed until the pressure needed to force the guar gum through the
Berea core reached 20 psi. This pressure likely increased due to guar gum filling pore spaces and
preventing aqueous flow. Over the duration of the test, approximately 360 ml of water/guar gum
solution passed through the Berea core. It took approximately 40 hours from the time that the
guar gum was introduced to the system until it was observed on the downstream end of the Berea
core. Samples were collected to measure the concentration of guar gum in the effluent and
extrapolate the amount of guar gum retained in the core. After shutting down the experiment, the
guar gum-exposed Berea core was allowed to sit for 36 days before being removed from its
sleeve and core-holder apparatus and then cut into 3-cm thick pucks, using a diamond saw
wetted with 0.25 ounces of Struers cooling fluid per 1.5 gallons of tap water. After cutting, the
pucks were rinsed with de-ionized water. These samples were placed in zip-lock plastic bags and
shipped for analysis. Upon arrival, initial samples were impregnated with an epoxy, polished
down to a 3-µm diamond grit, and coated with palladium. However, the carbon-rich epoxy
interfered with carbon detection limits during analysis. A second set of samples were prepared
with no epoxy to avoid altering the guar gum microstructure.
To analyze the geochemical effects of guar gum on the Berea sandstone core, a field emission
scanning electron microscope (FE-SEM) FEI Inspect F was used to examine the sandstone
before and after interaction with guar gum. The inflow and outflow ends of the sample were
compared. A SEM back-scattered electron (BSE) detector allowed identification of
compositional boundaries, specifically cemented grains and organics. A sample’s composition
determines the degree to which electrons will be backscattered because BSE rates correlate with
atomic number; heavier elements produce stronger backscattering, which appears brighter in
images. In addition, energy dispersive X-ray spectroscopy (EDS) permitted elemental X-ray
mapping, elemental distribution, and spot analysis of the sample. SEM has inherent limitations
when characterizing sandstones, which are particularly susceptible to charging due to clay grains
and Fe-rich inclusions. For analysis, an accelerating voltage ranging 12–18 keV was used with a
spot size 4–5. Elemental intensities were determined by internal standards that were manually set
into Oxford-INCA software interfaced with the SEM and optimized with copper. Porosity
changes were completed qualitatively by comparing before and after images in area covered by
guar gum.

18
Characterization of the Oriskany and Berea Sandstones

3. OBSERVATIONS AND DISCUSSION

3.1 BEREA SANDSTONE ANALYSIS


The Berea sandstone is composed of ~95% angular quartz with trace amounts of clays, feldspar,
and heavy minerals. The quartz grains express conchoidal fracturing and are dissolute along the
sharp edges of the grains and the pore-grain interface. Most of the non-quartz minerals are clays
(e.g. kaolinite and illite) and feldspars. Cementation and recrystallization appear along the edges
of the Berea sandstone pores. The main components of Berea sandstone are oxygen, silica,
alumina, and carbon (Figures 11). Trace amounts of other elements, such as potassium, calcium,
iron, sulfur, chlorine, magnesium, and sodium, have also been discovered (Figure 11). Figure 12
shows that the pore spaces, which vary in size (< 20–100µm), have extensive connectivity and
are mostly angular. The pores are continuous and close in proximity.

1
Log Concentration (wt. %)

0.1

0.01

0.001

0.0001
C Al Si S Ca Fe O Na K Cl Mg
Elements

Figure 11: Composition of the Berea sandstone (log scale of the relative weight percent).

19
Characterization of the Oriskany and Berea Sandstones

Figure 12: SEM-BSE image of the angular porosity of the Berea sandstone.

3.2 BEREA-GUAR GUM FLOW-THROUGH ANALYSIS


The remainder of this report focuses on the geochemical and microstructural changes of the
Berea sandstone once guar gum has been injected. A general trend is that the guar gum is higher
in concentration near the injection site and decreases rapidly (<6 cm) toward the outflow. As a
result, guar gum covers a significant portion (>50%) of the inflow surface, whereas the guar gum
is less prevalent (<1% surface coverage) on the outflow surface. Figure 13 contrasts the inflow
and outflow surfaces of the Berea sandstone sample of this experiment. The Berea’s pores on and
near the inflow surface are blocked by the guar gum, which decreases the sandstone’s effective
porosity and permeability; however, the farther the distance away from the inflow surface, the
more porosity and permeability return to unaltered levels. Figure 14 depicts the original porosity
of pristine Berea sandstone compared to Berea sandstone post-flow-through. The higher
concentration of guar gum on the inflow surface compared to on the outflow surface is likely a
result of guar gum’s partial insolubility and suggests that the guar gum may prevent further flow
due to plugging of the pores, which inconsistent with guar gum rheological studies (Weaver et
al., 2003; Kawamura, 2008; Gastone et al., 2014).

20
Characterization of the Oriskany and Berea Sandstones

Figure 13: SEM-BSE images of post-guar gum inflow and outflow surfaces.

Figure 14: Porosity changes after guar gum flow-through experiment. Unaltered Berea
sandstone has similar porosity throughout its entirety (blue). The modal abundance of guar
gum is higher closer to the inflow surface resulting in lower porosity; as the abundance
decreases, the effective porosity increases across the core (red).

21
Characterization of the Oriskany and Berea Sandstones

Inflow Surface: Berea-Guar Gum Flow-Through Analysis


SEM-BSE images (Figures 15, 16, 17) show that the guar gum is not uniformly distributed
across the post-flow Berea sample, but it appears to stay in contiguous volumes. The quartz
within the Berea sandstone itself did not appear to change chemically; however, the injection of
the viscous guar gum into the Berea sandstone resulted in pores becoming blocked. There was
evidence of a guar gum residue that filled the majority of the inflow section (Figures 15, 16),
causing the carbon content of the post-flow sample to be several times greater than that of the
pre-flow sample. Kolb and Kamaria (1971) compared effects of guar gum gels on sandstones and
limestones and found that guar gum has a greater effect on limestone porosity than sandstone
porosity. However, Kolb and Kamaria (1971) do not examine if the limestone’s lower porosity or
different chemical composition (CaCO3 vs. SiO2) could be the cause of guar gum’s unequal
effects and conclude that additional research is needed to understand why limestone is more
affected than sandstone.
Guar gum’s presence appears to be accompanied by higher aluminum and potassium
concentrations along with slightly lower silica contents in SEM-EDS elemental analysis and
mapping of the inflow surface (Figures 15, 16, 17). The higher observed aluminum and
potassium concentrations are likely the effect of the SEM’s interaction volume including some
clay minerals because the guar gum and quartz contain no aluminum or potassium; therefore, the
correlation between carbon, aluminum, potassium, and silica contents suggests that the guar gum
adheres preferentially to clay minerals. However, the Berea sandstone does not contain enough
clay minerals to account for the percentage of the area covered by the guar gum on the inflow
sample. As seen in SEM images, some of the grains have thin clay coatings that could help
promote the guar gum to adhere to the surface, however the surface roughness of the angular
quartz and detrital quartz within the pores may promote trapping of the guar gum.
SEM-EDS analysis also identified iron species, likely iron hydroxide or iron oxide that appear as
lighter grains in SEM-BSE images (Figure 15), although the iron species are not a large modal
component of the Berea sandstone. However, iron appears to be homogenously dispersed in the
guar gum coating on the inflow surface (Figure 17). Guar Gum Spectra 1 in Figure 16 also
contained trace concentrations of nickel, a common iron substitute. Some of the iron observed in
the SEM-EDS analysis may be a product of secondary fluorescence from iron-oxide. Conversely,
the trace concentrations of iron may be a product of using tap water to dissolve the guar gum for
this experiment. Future iterations will use de-ionized water during the flow-through process to
avoid this potential complication.
Abdulrahman and Alsewailem (2011) found that adding ~20 weight percent NaCl to a guar gum
solution substantially increased viscosity. Although the sodium observed (<1 weight percent) is
four orders of magnitude lower than in the Abdulrahman and Alsewailem (2011) study, the role
NaCl plays in determining guar gum solution viscosity underlines the importance of using de-
ionized water in future studies to avoid sodium contamination.

22
Characterization of the Oriskany and Berea Sandstones

C wt. O wt. Na wt. Mg Al wt. Si wt. P wt. Cl wt. K wt. Ca wt. Ti wt. Fe wt.
Spectra
% % % wt. % % % % % % % % %
Guar
23.29 53.06 BDL BDL 23.02 0.34 0.08 BDL BDL 0.02 0.13 0.07
Gum 1
Guar
Gum 2
48.95 24.57 0.06 0.04 1.07 0.47 0.02 BDL 0.08 BDL 9.34 15.40
w/ Fe-
Oxide
Quartz 6.21 53.78 BDL BDL BDL 40.01 BDL BDL BDL BDL BDL BDL
Fe-
5.58 49.68 0.28 0.36 2.60 4.27 0.37 0.24 0.56 0.10 0.02 35.94
Oxide

Figure 15: SEM-BSE image coupled with EDS of the Berea sandstone inflow section with a
dark coating of guar gum.

23
Characterization of the Oriskany and Berea Sandstones

C wt. O wt. Al wt. Si wt. P wt. Cl wt. K wt. Ca wt. Ti wt. Fe wt. Ni wt.
Spectra
% % % % % % % % % % %
Guar
71.91 15.32 0.52 9.01 0.08 0.20 0.04 0.12 2.24 0.19 0.28
Gum 1
Guar
Gum 2
27.38 34.39 1.00 9.44 BDL BDL BDL BDL 4.31 23.49 BDL
w/ Fe-
Oxide
Guar
67.40 26.78 0.69 1.63 BDL 0.21 BDL BDL 2.98 0.30 BDL
Gum 3
Guar
60.02 37.99 BDL 1.99 BDL BDL BDL BDL BDL BDL BDL
Gum 4

Figure 16: SEM-BSE image and EDS analysis of the guar gum at the granular level. The
guar gum does not appear to adhere as a thin coating across all minerals’ surfaces. Instead,
the guar gum prefers to stay in contiguous volumes within the Berea sandstone with areas of
unblocked pore space and uncovered minerals.

24
Characterization of the Oriskany and Berea Sandstones

Figure 17: SEM-EDS elemental maps of Berea sandstone inflow surface. The detection of
non-carbon elements through the guar gum coating suggests the guar gum layer is thin (<10
µm). Regions of higher carbon concentrations (guar gum) appear to be accompanied by
aluminum, potassium, and lower silica concentrations, suggesting guar gum preferentially
adheres to clay minerals. Each frame is of the same field of view.

Internal and Outflow: Berea-Guar Gum Flow-Through Analysis


Guar gum was mostly deposited in the pore space of the Berea sandstone within ~3 cm from the
core’s inflow surface, and the volume of the guar gum adhering to the sandstone appears to
decrease with increasing distance from the inflow surface (Figure 14, 18). In Figure 18, the
internal section of the Berea sandstone does not appear to be as filled with guar gum as the
inflow sample surface. The guar gum invades some but not all pore spaces. This phenomenon
may be attributed to the pores’ interconnectivity and the guar gum’s limited solubility in solution
(Weaver et al., 2003; Kawamura, 2008). The insoluble particulates could block the inflow
surface and restrict flow. As the guar gum moves through the sandstone, its abundance decreases
with respect to that of the surrounding sandstone, suggesting limited transport of the material.

25
Characterization of the Oriskany and Berea Sandstones

As the guar gum abundance decreases, organic particles of carbon and nitrogen become
observable (Figure 19). These individual particles do not seem to be dependent on pore space.
The carbon-nitrogen particles also appear on the outflow surface of the Berea core. Figure 20
gives the elemental composition of the carbon-nitrogen particles along with the composition for
the surrounding Berea core outflow surface. Figure 20 also provides an elemental map displaying
the composition of the organic particles surrounding Berea sandstone.

Figure 18: SEM-BSE image of guar gum filling pore spaces. This image is of a core cut
~6 cm from the inflow surface.

26
Characterization of the Oriskany and Berea Sandstones

Figure 19: SEM-BSE image of organic carbon and nitrogen. These organic particles
(depicted by darker objects in the images) were less than 6 cm from the inflow surface.

27
Characterization of the Oriskany and Berea Sandstones

C wt. N wt. O wt. Na Al wt. Si wt. S wt. Cl wt. K wt. Fe wt.


Spectra % % % wt. % % % % % % %
Unknown 1 48.92 21.10 23.34 1.67 0.09 2.89 1.42 BDL BDL 0.58
Unknown 2 46.13 15.31 31.81 1.63 0.12 3.61 1.38 BDL BDL BDL
Quartz 2.64 BDL 58.50 BDL BDL 38.58 BDL BDL BDL BDL

Figure 20: SEM-BSE images and maps coupled with EDS analysis identify organic particles
in Berea outflow sample.

28
Characterization of the Oriskany and Berea Sandstones

Elemental maps depict the strong contrast of the compositions of the organic-rich carbon
particles, against the Berea sandstone. The carbon-nitrogen particles appear relatively depleted of
oxygen. The traces of sulfur, sodium, and aluminum in the carbon-nitrogen material suggest that
some sort of sulfur, sodium, and aluminum species are present. The silica, which has greater
concentrations than the sulfur, sodium, and aluminum, is likely a product of the SEM beam
interaction volume including quartz, suggesting these less abundant elements could also be
products of an interaction volume larger than the phase being analyzed.
The carbon-nitrogen particles may be a product of microbial activity fixing nitrogen from the air
to organic products, or a contaminant from the Struers cooling fluid used to cut the samples.
Previous microbial research does not indicate that microbes consume atmospheric nitrogen when
in contact with guar gum (Salyers et al., 1977; Salyers et al., 1978; Hartemink et al., 1999).
Mariotti et al. (2001) studied whether guar gum could affect the absorption of nitrogen in the
digestive tracks of larger organisms and concluded that guar gum had no noticeable effect.
Contamination from the Struers cooling fluid is a reasonable cause for observation of these
particles. The C:N ratio appears to be roughly 2:1, and the lack of oxygen in the particles is
consistent with the HCCN chemical composition of the Struers cooling fluid.
Fully understanding the complete geochemical implications of guar gum’s interaction with the
Berea sandstone requires additional research. This initial study shows that residual guar gum
remains in contact with a sandstone after the fluid. Future research should utilize only different
de-ionized water cutting techniques to remove the possibility of exposing samples to potential
contaminants.

29
Characterization of the Oriskany and Berea Sandstones

4. CONCLUSION
The Marcellus shale of the Appalachian Basin is undergoing hydraulic fracturing to promote
natural gas production from the shale. Hydraulic fracturing fluid emplaces several chemicals into
the deep subsurface to extract shale-gas. The degradation pathways and reactive potential of
these chemicals with substrate mineralogy has not been widely reported in the existing literature.
Should a spill of these hydraulic fracturing fluids occur, the necessary environmental
remediation methods and residence time of these chemicals in the environment need to be known
in order to implement the appropriate environmental management strategies.
This report describes minerals in the Oriskany and Berea sandstones and the Marcellus shale
(e.g. quartz, feldspar, calcium carbonates, halite, kaolinite, illite, chlorite, mica, pyrite,
tourmaline, zircon, siderite, gypsum, and anhydrite). This study reviewed common hydraulic
fracturing fluid organic chemical additives and summarized their degradation pathways and
reactive potential with substrate mineralogy of several of the more common hydraulic fracturing
organic chemical additives (polyacrylamide, ethylene glycol, poly[diallyldimethylammonium
chloride], glutaraldehyde, and isopropanol). The range of degradation variables is as diverse as
the number of hydraulic fracturing fluids used in the industry. Every spill will need to be
addressed on a case-by-case basis to account for fracturing fluids specific to that spill.
Berea sandstone exposed to guar gum was characterized and analyzed in a water solution. The
first several centimeters of the Berea sandstone was partially coated and filled in by a thin layer
of guar gum; however, the guar gum residue did not penetrate deeply into the sample.

30
Characterization of the Oriskany and Berea Sandstones

5. REFERENCES
Abdulrahman, A. A.; Alsewailem, F. D. Adsorption of Guar, Xanthan and Xanthan-Guar
Mixtures on High Salinity, High Temperature Reservoirs, Offshore Mediterranean
Conference and Exhibition, Ravenna, Italy, March 23–25, 2011; Offshore Mediterranean
Conference: Ravenna, 2011.
Berner, R. A. The role of magnesium in the crystal growth of calcite and aragonite from sea
water. Geochimica Et Cosmochimica Acta 1975, 39, 489–504.
Bjelopavlic, M.; Singh, P. K.; El-Shall, H.; Moudgil, B. M. Role of Surface Molecular
Architecture and Energetics of Hydrogen Bonding Sites in Adsorption of Polymers and
Surfactants. Journal of Colloid and Interface Science 2000, 226, 159–169.
Blauch, M. E.; Meyers, R. R.; Lipinski, B. A.; Houston, N. A. Marcellus Shale Post-Frac
Flowback Waters—Where is All That Salt Coming From and What are the Implications?
Society of Petroleum Engineers Eastern Regional Meeting, Charleston, WV, Sept 23–25,
2009; Society of Petroleum Engineers: Charleston, WV.
Boulton, A.; Fornasiero, D.; Ralston, J. Selective Depression of Pyrite with Polyacrylamide
Polymers. International Journal of Mineral Processing 2000, 61, 13–22.
Boyce, M. L.; Carr, T. R. Lithostratigraphy and Petrophysics of the Devonian Marcellus Interval
in West Virginia and Southerwestern Pennsylvania Abstract, American Association of
Petroleum Geologists Student Expo, Sept 21–22, 2009.
Bruner, K. R.; Smosna, R. A Comparative Study of the Mississippian Barnett Shale, Forth Worth
Basin, and Devonian Marcellus Shale, Appalachian Basin; DOE/NETL-2011/1478;
National Energy Technology Laboratory: Morgantown, WV, 2011.
Bustard, M.; Meeyo, V.; Wright, P. C. Biodegradation of Isopropanol in a Three-Phase Fixed
Bed Bioreactor: Start Up and Acclimation Using a Previously Enriched Microbial
Culture. Environmental Technology 2001, 22, 1193–1201.
Capo, R. C.; Stewart, B. W.; Rowan, E. L; Kolesar, C. A.; Wall, A. J.; Chapman, E. C.;
Hammack, R. W.; Schroeder, K. T. The Strontium Isotopic Evolution of Marcellus
Formation Produced Waters, Southwestern Pennsylvania. International Journal of Coal
Geology 2014, 126, 57–63.
Chapman, E. C.; Capo, R. C.; Stewart, B. W.; Kirby, C. S.; Hammack, R. W.; Schroeder, K. T.;
Edenborn, H. M. Geochemical and Strontium Isotope Characterization of Produced
Waters from Marcellus Shale Natural Gas Extraction. Environmental Science and
Technology 2012, 46, 3545–3553.
Dana, J. D. Manual of Science, 23rd ed.; Klein, C., Dutrow, B., Eds.; John Wiley and Sons: New
Jersey, 2008.
de Ory, I.; Romero, L. E.; Cantero, D. Modeling the Kinetics of Growth of the Acebactor Aceti
in Discontinuous Culture: Influence of the Temperature of Operation. Applied
Microbiology Biotechnology 1998, 49, 189–193.
Deer, W. A.; Howie, R. A.; Zussman, J. An Introduction to the Rock Forming Minerals, 2nd ed.;
Pearson Education Limited: Harlow, 1992.

31
Characterization of the Oriskany and Berea Sandstones

Dresel, P. E.; Rose, A. W. Chemistry and origin of oil and gas well brines in western
Pennsylvania; Open-File Report OFOG 10–01.0; Pennsylvania Geological Survey, 4th
series, 2010.
Dwyer, D. F.; Tiedje, J. M. Metabolism of Polyethylene Glycol by Two Anaerobic Bacteria,
Desulfovibrio Desulfuricans and a Baceroides sp. Applied and Environmental
Microbiology 1986, 52, 852–856.
Edwards, N.; Lowry, G.; Karamalidis, A. K. Hydrolysis of Hydraulic Fracturing Fluid Organic
Additives and Their Interaction with Pyrite, National Groundwater Association
Workshop, Pittsburgh, PA, Nov 13–14, 2014; National Groundwater Association:
Westerville, OH, 2014.
Emmanuel, E.; Hanna, K.; Bazin, C.; Keck, G.; Clement, B.; Perrodin, Y. Fate of Glutaraldehyde
in Hospital Wastewater and Combined Effects of Glutaraldehyde on Surfactants on
Aquatic Organisms. Environmental International 2005, 31, 399–406.
Enomoto, C. B.; Coleman, J. L.; Haynes, J. T.; Whitmeyer, S. J.; McDowell, R. R.; Lewis, E. J.
Geology of the Devonian Marcellus Shale—Valley and Ridge Province, Virginia and
West Virginia— A Field Trip Guidebook, The American Association of Petroleum
Geologists Eastern Section Meeting, Sept 28–29, 2011; American Association of
Petroleum Geologists.
Flaten, E. M. Polymorphism and Morphology of Calcium Carbonate Precipitated in Mixed
Solvents of Ethylene Glycol and Water. Journal of Crystal Growth 2009, 311, 3533–
3538.
Flaten, E. M.; Seiersten, M.; Andreassen, J. P. Induction Time Studies of Calcium Carbonate in
Ethylene Glycol and Water. Chemical Engineering Research and Design 2010, 88, 1659–
1668.
Fox, P.; Ketha, S. Anaerobic Treatment of High-Sulphate Wastewater and Substrate Interactions
with Isopropanol. Journal of Environmental Engineering 1996, 122, 989–994.
Frac Focus. What Chemicals are Used. https://fracfocus.org/chemical-use/what-chemicals-are-
used (accessed Feb 20th, 2015).
Gastone, F.; Tosco, T.; Sethi, R. Guar Gum Solutions for Improved Delivery of Iron Particles in
Porous Media (Part 1): Porous Medium Rheology and Guar Gum-Induced Clogging.
Journal of Contaminant Hydrology 2014, 166, 23–33.
Graveling, G. J.; Ragnarsdottir, V. K.; Allen, G. C.; Eastman, J.; Brady, P. V.; Basley, S. D.
Controls on Polyacrylamide Adsorption to Quartz, Kaolinite, and Feldspar. Geochimicia
et Cosmochimica Acta 1990, 61, 3515–3523.
Gruden, C. L.; Dow, S. M.; Hernandez, M. T. Fate and Toxicity of Aircraft Deicing Fluid
Additives through Anaerobic Digestion. Water Environment Research 2001, 73, 72–79.
Guevellou, Y.; Noik, C.; Lecourtier, J.; Defives, D. Polyacrylamidc Adsorption onto Dissolving
Minerals at Basic pH. Colloids and Surfaces A: Physiochemical and Engineering
Prospects 1995, 100, 173–185.
Hartemink, R.; Schoustra, S. E.; Rombouts, F. M. Degradation of Guar Gum by Intestinal
Bacteria. Bioscience Microflora 1999, 18, 17–25.

32
Characterization of the Oriskany and Berea Sandstones

Hayduk, W.; Malik, V. K. Density, Viscosity, and Carbon Dioxide Solubility and Diffusivity in
Aqueous Ethlyene Glycol Solutions. Journal of Chemical Engineering 1971, 16, 143–
146.
Heinrich, E. W. M. Geologic Types of Glass-sand Deposits and some North American
Representatives. The Geological Society of America Bulletin 1981, 92, 611–613.
Hosterman, J. W.; Whitlow, S. I.; Clay Mineralogy of Devonian Shales in the Appalachian
Basin. Geological Survey Professional Paper 1298: U.S. Geological Survey, 1983.
Huang, Y. L.; Li, Q. B.; Deng, X.; Lu, Y. H.; Liao, X. K.; Hong, M. Y.; Aerobic and Anaerobic
Biodegradation of Polyethlyene Glycols Using Sludge Microbes. Process Biochemistry,
2005, 40, 207–211.
Hwang, Y.; Sakuma, H.; Tanaka, T. Denitrification with Isopropanol as a Carbon Source in a
Biofilm System. Water Science Technology 1995, 30, 69–81.
Inoue, A.; Horikoshi, K. A. Pseudomonas Thrives in High Concentrations of Toleuene. Nature
1989, 388, 264–266.
Jagucki, M.; Darner, R. A. Ground-Water Quality in Geauga County, Ohio—Review of Previous
Studies, Status in 1999, a Comparison of 1986 and 1999 Data. Report 01-416: Water–
Resources Investigations, 2001.
John, W. Synthesis, Properties, and Analysis of Polydadmac for Water Purification. Ph.D.
Dissertation, Stellenbosch University, Stellenbosch, South Africa, 2008.
Kahrilas, A.; Blotevogel, J.; Stewart, P. S.; Borch, T. Biocides in Hydraulic Fracturing Fluids: A
Critical Review of Their Usage, Mobility, Degradation, and Toxicity. Environmental
Science and Technology 2015, 49, 16–32.
Kanemitu, H.; Fukuda, M.; Yano, K. Plasmid-Borne Biodegradation of Toluene and
Ethylbenzene in a Pseodomonas. Journal of Fermentation Technology 1980, 57, 175–
181.
Kay-Shoemake, J. L.; Watwood, M. E.; Lentz, R. D.; Sojka, R. E. Polyacrylamide as an Organic
Nitrogen Source for Soil Microorganisms with Potential Effects on Inorganic Soil
Nitrogen in Agricultural Soil. Soil Biol. Biochemistry 1998, 30, 1045–1052.
Kawamura, Y. Guar Gum Technical and Chemical Assessment. Joint FAO/WHO Expert
Committee on Food Additives; Rome, Italy, 2008.
Kolb, A. K.; Kamaria, P. Laboratory Investigation of Damage from Guar Gum Base Gels,
Society of Petroleum Engineers Eastern Regional Meeting, Nov 4–5, 1971; Society of
Petroleum Engineers: Charleston, WV, 1971.
Konopacka-Łyskawa, D.; Lackowski, M. Influence of Ethylene Glycol on CaCO 3 Particles
Formation via Carbonation in the Gas–Slurry System. Journal of Crystal Growth 2011,
321, 136–141.
Kostelnik, K.; Carter, K. M. Unraveling the Stratigraphy of the Oriskany Sandstone: A Necessity
in Assessing Its Site-Specific Carbon Sequestration Potential. Environmental
Geosciences 2009, 16, 187–200.

33
Characterization of the Oriskany and Berea Sandstones

Landrum, P. F.; Sano, L.; Mapili, M. A.; Garcia, E.; Krueger, A. M.; Russell, A. M. Degradation
of Chemical Biocides with Application to Ballast Water Treatment; GLERL-123; NOAA
Technical Memorandum, 2003.
Laughrey, C. D.; Ruble, T. E.; Lemmens, H.; Kostelnik, J.; Butcher, A. R.; Walker, G. Black
Shale Diagenesis: Insights from Integrated High-Definition Analyses of Post-Mature
Marcellus Presentation, American Association of Petroleum Geologists Annual
Convention, Houston, TX, June 23, 2011; American Association of Petroleum
Geologists.
Lecourtier, J.; Lee, L. T.; Chauveteau, G. Adsorption of Polyacrylamides on Siliceous Minerals,
Colloids and Surfaces 1990, 47, 219–231.
Leff, E. Supplemental Generic Environmental Impact Statement on the Oil, Gas, and Mining
Regulatory Program; New York State Department of Environmental Conservation:
Albany, NY, 2011.
Leung, H. W. Ecotoxicology of Glutaraldehyde: Review of Environmental Fate and Effects
Studies. Ecotoxicology and Environmental Safety 2001a, 49, 26–39.
Leung, H. W. Aerobic and Anaerobic Metabolism of Glutaraldehyde in a River Water–Sediment
System. Archives of Environmental Contamination and Toxicology 2001b, 41, 267–273.
Mahajan, V.; Misra, M.; Zhong, K.; Fuerstenau, M. C. Enhanced Leaching of Copper from
Chalcopyrite in Hydrogen Peroxide–Glycol System. Minerals Engineering 2007, 20,
670–674.
Mali, Y. N.; Pawar, S. P.; Gujarathi, N. A.; Rane, B. R.; Bakliwal, S. R.; Applications of Natural
Polymers in Sustained Release Drug Deliversy System: A Review. Pharma Science
Monitor 2012, 3, 3314–3335.
Mariotti, F.; Pueyo, M. E.; Tome, D.; Benamouzig, R.; Mahe, S. Guar Gum Does Not Impair the
Absorption and Utilization of Dietary Nitrogen but Affects Early Endogoneous Urea
Kinetics in Humans. American Journal of Clinic Nutrition 2001, 74, 487–493.
McCarthy, K.; Rojas, K.; Niemann, M.; Palmowski, D.; Peters, K.; Stankiewicz, A. Basic
Petroleum Geochemistry for Sour Rock Evaluation. Oil Field Review 2011, 23, 32–43.
Midwest Regional Carbon Sequestration Partnership. MRCSP Regional Geologic Cross Section
and Correlation Chart. http://www.mrcsp.org/ (accessed Feb 17, 2015).
Moriya, K; Horikoshi, K. Isolation of a Benzene-Tolerant Bacterium and Its Hydrocarbon
Degradation. Journal of Fermentation Bioengineering 1993, 76, 397–399.
Morrison, R. T.; Boyd, R. N. Organic Chemistry, 5th ed.; Allyn Bacon: New York, 1987.
Murrays, J.; Roberta, A.; Hall, A.; Griffini, M. Microbial Metabolism by a Pure Strain of
Psuedomonas sp. Journal of General Microbiology 1980, 120, 89–94.
Nakamiya, K.; Kinoshita, S. Isolation of Polyacrylamide-Degrading Bacteria. Journal of
Fermentation 1995, 80, 418–420.
Nasser, M. S.; James, A. E. Effect of Polyacrylamide Polymers on Flox Size and Rheological
Behavior of Kaolinite Suspensions. Colloids and Surfaces A: Physiochemical and
Engineering Aspects 2007, 301, 311–322.

34
Characterization of the Oriskany and Berea Sandstones

Nesse, W. D. Introduction to Mineralogy, 2nd ed.; Oxford University Press: Oxford, New York:
2012.
Nyahay, R.; Leone, J.; Smith, L. B.; Martin, J. P.; Jarvie, D. J. Update on Regional Assessment of
Gas Potential in the Devonian Marcellus and Ordovician Utica Shales of New York,
American Association of Petroleum Geologists Eastern Section Meeting, Lexington, KY,
Sept 16–18, 2007; American Association of Petroleum Geologists Reservoir
Characterization Group, New York.
Peng, P.; Garnier, G. Effect of Cationic Polyacrylamide on Precipitated Calcium Carbonate
Flocculation: Kinetics, Charge Density, and Ionic Strength. Colloids and Surfaces A:
Physiochemical and Engineering Aspects 2012, 408, 32–39.
Pepper, J. F.; De Witt, W. J.; Demareset, D. F. Geology of the Bedford Shale and Berea
Sandstone in the Appalachian Basin: A Study of the Stratigraphy, Sedimentation, and
Paleogeography of Rocks of Bedford and Berea Age in Ohio and Adjacent States; U.S.
Department of the Interior and the U.S. Geological Survey: 1954.
Perkins, D. Mineralogy, 3rd ed.; Folchetti, N., Ed.; Pearson Education: Upper Saddle River, NJ,
2011.
Poppe, L. J.; Paskevich, V. F.; Hathaway, J. C.; Blackwood, D. S. USGS Coastal and Marine
Geology Program; Open-File Report 01-041; U.S. Geoloical Survey: 2002.
Quinlan, C.; Strevett, K.; Ketcham, M. VOC Elimination in a Compost Biofilter Using
Previously Acclimated Bacterial Inoculum. Journal of the Air Waste Management
Association 1999, 49, 544–553.
Repetski, J. E.; Ryder, R. T.; Weary, D. J.; Harris, A. G.; Trippi, M. H. Thermal Maturity
Patterns (CAI and %R0) in Upper Ordovician and Devonian Rocks of the Appalachian
Basin: A Major Revision of USGS Map I-917-E Using New Subsurface Collections;
Scientific Investigations Map 3006; U.S. Geological Survey, Reston, VA, 2008.
Rowan, E. L. Burial and Thermal History of the Central Appalachian Basin, Based on Three 2-
D Models of Ohio, Pennsylvania, and West Virginia; Open-File Report 2006-1019; U.S.
Geological Survey, Reston, VA, 2006.
Sageman, B. B.; Murphy, A. E.; Werne, J. P.; Ver Straeten, C. A.; Hollander, D. J.; Lyons, T. W.
(2003). A Tale of Shales: The Relative Roles of Production, Decomposition, and Dilution
in the Accumulation of Organic-Rich Strata, Middle-Upper Devonian, Appalachian
Basin. Chemical Geology 2003, 195, 229–273.
Salyers, A. A.; Palmer, J. K.; Wilkens, T. D. Degradation of Polysaccarides by Intestinal
Bacterial Enzymes. The American Journal of Clinical Nutrition 1978, 31, 319–322.
Salyers, A. A.; Vercellotti, J. R.; West, S. H.; Wilkins, T. D. Fermentation of Mucin and Plant
Polysaccarides by Strains of Bacteroides from the Human Colon. Applied and
Environmental Microbiology 1977, 319–322.
Samoshina, Y.; Diaz, A.; Becker, Y.; Nylander, T.; Lindman, B. Adsorption of Cationic,
Anionic, and Hydrophobically Modified Polyacrylamides on Silica Surfaces. Colloids
and Surfaces A: Physiochem. Eng. Aspects 2003, 231, 195–205.

35
Characterization of the Oriskany and Berea Sandstones

Schramm, E.; Schink, B. Ether-Cleaving Enzyme and Diol Dehydratase Involved in Anaerobic
Polyethylene Glycol Degradation by a New Homoacetogenic Bacterium. Biodegradation,
1991, 2, 71–79.
Selley, R. C. Elements of Petroleum Geology, 2nd ed.; Academic Press: San Diego, CA, 1997.
Siegel, J. M.; Kamen, M. D. Studies on the Metabolism of Photosynthetic Bacteria. VI.
Metabolism of Isopropanol by a New Strain of Rhodopseudomanas Gelatinosa. Journal
of Bacteriology 1950, 4, 286–293.
Soeder, D. L. Porosity and Permeability of Eastern Devonian Gas Shale. Society of Petroleum
Engineers Formation Evaluation 1988, 3, 116–124.
Soeder, D. J.; Enomoto, C. B.; Chermak, J. A. The Devonian Marcellus Shale and Millboro
Shale. The Geological Society of America Field Guide 2014, 35, 129–160.
Stahl, J. D.; Cameron, M. D.; Haselbach, J.; Aust, S. D. Biodegradation of Superabsorbent
Polymers in Soil. Environmental Science and Pollution 2000, 7, 83–88.
Tekin, N.; Demirbas, O.; Alkan, M. Adsorption of Cationic Polyacrylamide onto Kaolinite.
Microporous and Mesoporous Materials 2005, 85, 340–350.
Ueyama, H.; Yamauti, Y.; Tuki, N.; Fukibama, T. Studies on the Fermentation of
Petrochemicals. I. Taxonomic studies on Arthobacter sp. isolated from soil. Journal of
Fermentation Technology 1971, 49, 581–586.
U.S. Geological Survey. Natural Attenuation.
http://toxics.usgs.gov/definitions/natural_attenuation.html (accessed Feb 17, 2015).
Veltman, S.; Schoenberg, T.; Switzenbaum, M. S. Alcohol and Acid Formation During the
Anaerobic Decomposition of Propylene Glycol under Methanogenic Conditions.
Biodegradation 1998, 9, 113–118.
Wallace, G. A.; Wallace, A. Control of Soil Erosion by Polymeric Soil Conditioners. Soil
Science 1986, 141, 363–367.
Walter, L. M. Relative efficiency of carbonate dissolution and precipitation during diagenesis: a
progress report on the role of solution chemistry. Roles of organic matter in mineral
diagenesis. Society of Economic Paleontologists and Mineralogists. Special Publication
1986, 38, 1–12.
Wang, C. C.; Lee, C. M. Denitifrication with Acrylamide by Pure Culture of Bacteria Isolated
from Acrylonitrile-Butadiene-Styene Resin Manufactured Wastewater Treatment System.
Chemosphere 2001, 44, 1047–1053.
Wang, C.; Zhao, J.; Zhao, X.; Bala, H.; Wang, Z. Synthesis of Nanosized Calcium Carbonate
(Aragonite) via Polyacrylamide Inducing Process. Powder Technology 2006, 163, 134–
138.
Wang, D.; Wu, M.; Wang, Q.; Wang, T.; Chen, J. Controlled growth of uniform nanoflakes-built
pyrite FeS2 microspheres and their electrochemical properties. Ionics 2011, 17, 163–167.
Wang, G.; Carr, T. R. Organic-Rich Marcellus Shale Lithofacies Modeling and Distribution
Pattern Analysis in the Appalachian Basin. American Association of Petroleum
Geologists Bulletin 2013, 97, 2173–2205.

36
Characterization of the Oriskany and Berea Sandstones

Weaver, J.; Gdanski, R.; Karcher, A. Guar Gum Degradation: A Kinetic Study, International
Symposium on Oilfield Chemistry, Houston, Texas, Feb 5–7, 2003; Richardson: Society
of Petroleum Engineers.
Wen, Q.; Chen, Z.; Zhao, Y.; Zhang, H.; Feng, Y. Biodegradation of Polyacrylamide by Bacteria
Isolated from Activated Sludge and Oil-Contaminated Soil. Journal of Hazardous
Materials 2010, 175, 955–959.
Werne, J. P.; Sageman, B. B.; Lyons, T. W.; Hollander, D. J. An Integrated Assessment of a
“Type Euxinic” Deposit: Evidence for Multiple Controls on Black Shale Deposition in
the Middle Devonian Oatka Creek Formation. American Journal of Science 2002, 302,
110–143.
Woodrow, D. L. Paleogeography, Paleoclimate, and Sedimentary Processes of the Late
Devonian Catskill Delta. Geological Society of America Special Paper 1985, 201, 51–64.
Yen, T. F.; Chilingarian, G.V. Developments in Petroleum Science: Oil Shale; American
Elsevier Publishing Company, Inc.: New York, 1976.
Zielinksi, R. E.; McIver, R. D. Resource Exploration Assessment of the Oil and Gas Potential in
Devonian Gas Shales in the Appalachian Basin; MLM-MU-82-61-0002/DOE/DP/0053-
1125; U.S. Department of Energy, Morgantown, WV, 1981.
Ziemkiewicz, P.; Hause, J.; Gutta, B.; Fillhart, J.; Mack, B.; O'Neil, M. Final Report Water
Quality Literature Review and Field Monitoring of Active Shale Gas Wells Phase I For
“Assessing Environmental Impacts of Horizontal Gas Well Drilling Operations”; West
Virginia Department of Environmental Protection, Charleston, WV, 2013.

37
Characterization of the Oriskany and Berea Sandstones

This page intentionally left blank.

38
View publication stats

Sean Plasynski Elena Melchert


Executive Director Director
Technology Development & Integration Division of Upstream Oil and Gas
Center Research
National Energy Technology Laboratory U.S. Department of Energy
U.S. Department of Energy
Cynthia Powell
Jared Ciferno Executive Director
Associate Director Research & Innovation Center
Oil and Gas National Energy Technology Laboratory
Technology Development & Integration U.S. Department of Energy
Center
National Energy Technology Laboratory
U.S. Department of Energy

NETL Technical Report Series

You might also like