Вы находитесь на странице: 1из 15

Research Paper THEMED ISSUE:  Geothermal Energy from Sedimentary Basins: Challenges, Potential, and Ways Forward

GEOSPHERE Heat flow and thermal modeling of the Appalachian


Basin, West Virginia
GEOSPHERE; v. 11, no. 5
Z.S. Frone, D.D. Blackwell, M.C. Richards, and M.J. Hornbach
Roy M. Huffington Department of Earth Sciences, Southern Methodist University, 3225 Daniel Avenue, Dallas, Texas 75223, USA
doi:10.1130/GES01155.1

7 figures; 1 table
ABSTRACT poor understanding of the heat flow in the basin and the nature of the thermal
CORRESPONDENCE:  zfrone1@gmail.edu properties in the crust below it. Both of these properties have implication for oil
Recent heat flow studies indicate that the Appalachian Basin in West Vir- and gas maturation and for geothermal energy utilization.
CITATION:  Frone, Z.S., Blackwell, D.D., Richards, ginia may represent an important location for high heat flow and future geo- In this paper 6 equilibrium temperature logs and 148 corrected bottom-hole
M.C., and Hornbach, M.J., 2015, Heat flow and ther-
mal modeling of the Appalachian Basin, West thermal energy development. Currently, however, only limited one-dimensional temperatures (BHT) are used to estimate basin temperatures and the surface
Virginia: Geosphere, v. 11, no. 5, p. 1279–1290, doi: (1-D) heat flow studies exist in this region, making it difficult to assess the po- heat flow along a transect of the West Virginia portion of the Appalachian Basin.
10.1130 /GES01155.1. tential for geothermal development. Here, we develop the first high resolution 2- Two temperature corrections are tested and compared against wells with
D basin model for a portion of West Virginia. The model uses 2-D finite differ- equilibrated temperature logs to evaluate the validity of the corrections.
Received 21 November 2014
ence heat conduction, basin cross sections, equilibrium temperature, and oil and Numerical modeling of a basin cross section is used to test the validity and
Revision received 29 June 2015
Accepted 6 July 2015 gas bottom-hole temperature data to quantify heat flow at the surface and at the potential causes for a previously observed heat flow anomaly, and to predict the
Published online 15 September 2015 base of the sedimentary basin. The temperature data show elevated tem-perature equilibrium temperature distribution with the basin. The results provide insight
gradients in the eastern portion of the basin. A 2-D advection-diffusion model, into the nature of the crust below the basin, constraints on the surface heat flow,
created using available lithologic and structural data, was designed to test and a map of current basin temperatures.
whether variations in crustal properties, structure, erosion, or fluid advec-tion
can account for the observed temperatures in the basin. Thermal properties were
populated using measured values as well as published averages. A linear heat BACKGROUND
flow vs. heat production relationship was used to determine heat flow at the
base of the model. The model constrains the heat flow at the base of the Geologic History
sedimentary basin to 49–55 mW/m 2. Analysis of modeling results suggests that
heat flow at the base of the sedimentary basin is nearly uniform. Variations in The Appalachian Basin is a foreland basin located in the eastern United
basin temperatures are most likely due to variations in sediment thermal prop- States (Fig. 1). The basin is ~480,400 km 2 in area and underlies portions of New
erties, complex structures, and/or localized fluid advection. York, Pennsylvania, eastern Ohio, western Maryland, West Virginia, eastern
Kentucky, western Virginia, Tennessee, northwestern Georgia, and Alabama
(Ryder, 2006). Sedimentary rocks contained in the basin range in age from early
INTRODUCTION Cambrian (540 Ma) to early Permian (280 Ma) (Fig. 2). Three major oro-genic
events affected the sedimentation and structure in the basin during the Paleozoic
The thermal regime of the lithosphere is important for understanding for- Era. From oldest to youngest these are the Taconic orogeny (Ordo-vician), the
mation and evolution of the Earth’s crust (Pollack, 1982; Nyblade and Pollack, Acadian orogeny (Devonian to Early Mississippian), and the Alle ghanian
1993; Jaupart and Mareschal, 1999). Heat flow provides insight into the ther-mal orogeny (Mississippian to Permian) (Ryder, 2006). The Rome Trough forms the
regime of the lithosphere on a million year time scale. Stable continental igneous deepest part of the basin (Fig. 3). The trough is interpreted to be a major graben
terrains have been shown to have an approximately linear relation-ship between formed by passive margin extension during the early Cambrian along the east
measured surface heat flow and crustal heat production (Birch et al., 1968; coast of the North American craton (Gao et al., 2000). It extends northeast from
Lachenbruch, 1968; Roy et al., 1968; Costain et al., 1986; Decker et al., 1988). eastern Kentucky through West Virginia and western Penn-sylvania. Grenville
This relationship implies that changes in surface heat flow within a stable crustal age (1.4 Ga) metasedimentary and igneous rocks underlie most of the basin
block are related primarily to lateral variation in average crustal heat production. (Shumaker and Wilson, 1996).
For permission to copy, contact Copyright There have been a limited number of heat flow studies in West Virginia (Joyner, The West Virginia portion of the basin contains sedimentary rocks deposited
Permissions, GSA, or editing@geosociety.org. 1960; Frone and Blackwell, 2010). This has resulted in a from the Cambrian through the Pennsylvanian (Fig. 2). Cambrian and Ordo-
© 2015 Geological Society of America

Frone et al.  | Heat flow and thermal modeling of the Appalachian Basin 1279
GEOSPHERE | Volume 11  | Number 5
Research Paper
45°N Lake

Huron
Lake Ontario
43°N Lake

Michigan
41°N

(WGS84)
Atlantic
Ocean
39°N

Latitude
37°N

35°N Map

Area
33°N 0 km 100 km 200 km 300 km 400 km

–89°W –87°W –85°W –83°W –81°W –79°W –77°W –75°W –73°W –71°W

Longitude (WGS84)
Figure 1. Outline of Appalachian Basin in the eastern United States. WGS84—World Geodetic System 1984.

vician lithologies are generally passive margin deposits consisting of marine of the Salina salt basin resulted in deposition of as much as 400 m of evaporate

sandstones and carbonates deposited in a shallow sea (Milici, 1996). Thicker minerals in the central portion of the basin (Milici, 1996). From the late Silurian
sections of the Rose Run Sandstone and Black River–Trenton carbonates within into the Early Devonian, carbonate shelf deposits dominated deposition. The
the Rome Trough suggest that the trough was actively subsiding during the beginning of the Acadian orogeny in the Late Devonian resulted from oblique
Cambrian and Ordovician (Patchen et al., 2006). The onset of Taconic orogeny convergence of Laurentia with Avalonian continental fragments (Ettensohn,
sedimentation is marked by the Knox unconformity (Milici, 1996). The uncon- 1985). Devonian shales (Marcellus Shale) were deposited on the margins of
formity is the result of Middle Ordovician uplift and erosion of the Early Ordo- the Catskill Delta (Ryder, 2006). Devonian black shales are overlain by Late
vician carbonate platform (Ryder, 2006). During the Taconic orogeny, the Blue Devonian–Middle Mississippian sandstones and interbedded sandstones and
Ridge crystalline basement rocks were thrust to the surface on the eastern mar- shales (Swezey, 2002). A period of carbonate platform deposition occurred in
gin of the Appalachian foreland. This resulted in erosion and widespread silici- the Middle–Late Mississippian. In the Late Mississippian the Alleghanian orog-
clastic sediment deposition within the basin from the Late Ordovician to the eny resulted from the closing of the Iapetus Ocean basin and the assemblage
early Silurian. Early Silurian sandstones and conglomerates were overlain by of Pangea. This event is marked by a shift from marine limestones and shale
marine shales (Milici, 1996). Deposition of siliceous units related to the Taconic sedimentation to coal-bearing siliciclastic units (Milici, 1996). Sandstone and
orogeny gave way to carbonate shelf and reef limestones and dolomites in the interbedded sand and shales units eroded from the Alleghanian orogeny domi
middle and late Silurian (Swezey, 2002). During the late Silurian the subsidence nate the Pennsylvanian and Permian Periods (Swezey, 2002).
GEOSPHERE | Volume 11  | Number 5 Frone et al.  | Heat flow and thermal modeling of the Appalachian Basin 1280
Research Paper

250 the thermal conductivity values are measured from the well core or cuttings. -

Permian
275 Ideally, the wellbore is cased from surface to total depth and has been shut-in or
not producing for an extended period of time to allow for near-field tem-perature
disturbances cause by drilling and production to dissipate. This type of
Alleghanian measurement can produce accurate results, but it can be difficult to obtain
Orogeny
access to multiple wells within a given area. The difficulty in acquiring detailed
300
equilibrium temperature logs has resulted in development of other methods of

Carboniferous
Miss.Penn.
Siliciclastics
estimating heat flow. The same types of data are needed (temperature gra-dient
325 (Sandstone, Shale, etc.) and thermal conductivity), but typically they are of lower quality. While continuous
Carbonates or multiple temperature measurements in a single well are rare, a single
350 (Limestone and Dolomite) temperature value in a well is very common. In petroleum wells a single
Evaporites maximum recorded temperature (or BHT) is typically measured as a part of the
Acadian
well logging process after drilling. This type of data is inherently noisy and results
Orogeny Igneous and Metamorphic in large uncertainties in the temperature gradient (Blackwell and Steele, 1989).
375 basement Despite the error associated with heat flow measurements de-rived from BHT,

Devonian
(Ma)
400 they are typically within 15% (Chapman et al., 1984). The large number of BHT
data points that typically exist in sedimentary basins (often hundreds to
thousands in oil and gas production regions) makes these data extremely
Age

valuable for broad scale [two-dimensional (2-D) and 3-D] heat flow studies in
basins (Andrews-Speed et al., 1984; Chapman et al., 1984; Deming
Silurian

425 and Chapman, 1988).


Taconic The first heat flow study in West Virginia resulted in three heat flow values for
Orogeny the state (Joyner, 1960). These values range from 50 to 53 mW/m 2 ± 25% for
Ordovicia

450 Figure 2. Generalized stratigraphic column


northern West Virginia. A more recent study of heat flow in West Virginia using
n

BHT data shows a complex heat flow pattern across the state with an apparent
for the Appalachian Basin (modified from
Ryder et al., 2009). Penn.—Pennsylvanian;
high heat flow anomaly (~75 mW/m2) in eastern West Virginia (Frone and
475
Miss.—Mississippian; pC—Precambrian. Blackwell, 2010). This study used a generalized 1-D basin model for deter-mining
500 thermal conductivity values for individual wells. The results indicated heat flow
Cambrian

values of 40–50 mW/m2 on the western edge of the state with values increasing
eastward, with the highest values (70–75 mW/m 2) generally east of the Rome
525 Trough. These results are significantly different from previous heat flow values
and maps (Joyner, 1960; Blackwell et al., 1990; Blackwell and Richards, 2004)
that show a heat flow of 50–55 mW/m 2 for practically the entire state. Continental
heat flow globally is 65 ± 1.6 mW/m 2 and 61 ± 1.2 mW/m 2 for Paleozoic
sedimentary and metamorphic rocks (Pollack et al., 1993).
pC

550

Heat Flow–Heat Production Relationship


Basin Heat Flow
The concept of a linear correlation between surface heat flow (Q) and
plutonic unit radioactivity (A) was introduced by Birch et al. (1968) to explain the
Heat flow is calculated from the product of the formation temperature gradient
relationship of heat flow to crustal radioactivity in granitic rocks in New England
and thermal conductivity. Standard measurement requires an equili brated
and the Adirondack Mountains of upstate New York. Similar linear re-lationships
borehole temperature log, with measurements typically taken every 1–10 m, and
were also found by Lachenbruch (1968) and Roy et al. (1968) using Equation 1.
core or cuttings from the same well interval where the tempera-tures were
measured (Blackwell and Steele, 1989; Blackwell and Spafford, 1987).
Temperature gradients are calculated from the temperature log and (1)
Q = Qo + bA,

GEOSPHERE | Volume 11  | Number 5 Frone et al.  | Heat flow and thermal modeling of the Appalachian Basin 1281
Research Paper

41°N
TROUGH
OH PA
ROME
A
WV-15
40°N
WV-6
WV-7 WV-14
MD Figure 3. Map of West Virginia showing
locations of temperature-depth data. Data
Latitude (WGS84) WV-11 WV-13 WV-3 WV-8 from Spicer (1964) are shown as yellow
triangles and bottom-hole temperatures
WV-12

39°N
WV-1 A′ (BHT) as points. Colored points are BHTs
used for the basin modeling. Blue points
WV-9 WV-4 WV-5 are BHTs within and west of the Rome
Trough, and red points are BHTs east of
the Rome Trough. Brown lines are major
WV-2 WV-10
T basement structural boundaries. A–A′ is
ALLEGHENY FRON the location of the cross section (Ryder et
al., 2009) used to build the lithology
38°N model. WGS84—World Geodetic System
1984.

KY 0 km 50 km 100 km 150 km 200 km


VA
37°N
–83°W –82°W –81°W –80°W –79°W –78°W

Longitude (WGS84)

where Q is surface heat flow (mW/m 2), Qo is reduced heat flow or intercept Equilibrium
(mW/m2), b is the thickness of the radioactive upper crust (km), and A is the heat
production (mW/m3) at the site of heat flow measurement. The Qo was originally A subset of 15 temperature logs located in West Virginia from a national
identified as the heat that originates from below the radiogenically enriched upper collection was used (Spicer, 1964). Most of the data date back to the 1920s and
crustal layer and includes the mantle and deep crustal contri-bution to heat flow 1930s and were collected using maximum reading thermometers (Van Orstrand,
(Roy et al., 1988). The combined data of New England and the central stable 1935). One of the wells in this collection was drilled in 1891 and at the time was
region of the United States, which included West Virginia, yield a slope b of 7.5 ± the second-deepest well in the world; Hallock (1897) provided some insight to the
0.2 km and an intercept (Qo) of 33 ± 1 mW/m 2; subse-quent studies (Costain et data collection and provided evidence that these data represent equilibrium
al., 1986; Decker et al., 1988) resulted in the same average. Roy et al. (1968, temperatures. In addition, comparison of wells in the El Durado field in Kansas
1972) suggested that the linear relationship may have broad applicability for from Spicer (1964) and from McKenna and Blackwell (2001) show consistent
continental heat flow and to the stable portion of the continents (Thakur, 2011). results between the maximum reading thermometer data set and modern electric
logging tools. The logs consist of temperature measure-ments collected at 250–
500 m intervals over the depth of the wells. Well depths range from 1200 to 2200
m, and are located primarily in the northern and west-ern portion of Kansas (Figs.
METHODS 3 and 4).

Temperature Data BHT

Two types of temperature data are used for this study: equilibrium tempera- Oil and gas BHT data were collected from various publications (Ameri-can
tures and BHTs. Both data sets are from published and/or public data sources, Association of Petroleum Geologists, 1994; Hendry et al., 1982) and from the
therefore no new temperature measurements were completed for this study. West Virginia Economic and Geological Survey online database of well

GEOSPHERE | Volume 11  | Number 5 Frone et al.  | Heat flow and thermal modeling of the Appalachian Basin 1282
Research Paper

1000 after drilling (4–10 h) could be as much as 25%. There are different methods that
500 can be used to make this correction (Horner, 1951; Förster et al., 1997; Harrison
et al., 1983). The Horner correction, while the most robust, requires multiple
0 temperatures collected over some time interval at the same depth. Multiple log
runs were more common in the past (pre-1990), but typically not all of the logs
–500 have a BHT recorded or not all of the logs have been reported to the state.
Therefore, we did not use this method. The Förster et al. (1997) and Harrison et
–1000
al. (1983) methods are purely empirical corrections based only on the drilled
depth of the well. In both correction methods a correction factor is calculated

(m)
–1500
from an empirically derived equation. Förster et al. (1997) used a liner correction
–2500 method, while Harrison et al. (1983) used a second-order poly-nomial function.

Elevation
–2000 Using the equilibrium temperature data from West Virginia, both methods were
Temperature-Depth Data tested and the Harrison et al. (1983) correction was found to produce a
temperature distribution more consistent with equilibrium data. In addition, BHTs
shallower than 1 km were shown to better represent in situ temperatures without
–3000 WV-1
WV-3 a correction and values deeper than 3900 m were given a constant correction of
–3500 WV-6 19.1 °C (the maximum correction of the Harrison poly-nomial). The final
WV-8 piecewise function correction is shown in Equation 2, where z is the measured
–4000 WV-14 depth in meters and DT is the correction (in °C). This value is then added to the
WV-15
reported BHT for the final corrected temperature.
–4500 Corrected BHTs-west
Corrected BHTs-east 0, if z < 1 km
−6 2
if 1 km ≤ z ≤ 3.9 km. (2)
–5000 T = −16.51+ 0.018z − 2.34 × 10 z ,
19.1, if z > 3.9km
0 25 50 75 100 125 150
Temperature (°C)

Figure 4. Temperature-depth data used in the cross-section model. Solid lines are equilibrium
temperature logs, points are bottom-hole temperature (BHT) data. BHT point colors are the
Model
same as in Figure 3. East and west defined by position relative to the eastern boundary of the
Rome Trough (Fig. 3). BHT data are corrected using the polynomial correction from Harrison To investigate the potential causes for the observed heat flow and tempera-
et al. (1983) as modified in Equation 2.
ture variations in northern West Virginia, a 2-D steady-state temperature model
was developed. The model was constructed using the detailed west-northwest
stratigraphic cross section by Ryder et al. (2009) (A–A′, Fig. 3) and simulating
files. Data were limited to a minimum depth of 600 m for the study in order to heat flow advection-diffusion and the basement heat production. The goal of the
minimize the effects of shallow groundwater flow and the effects of surface model is to match equilibrium and BHT data along the cross section by varying
temperature changes on drilling mud temperatures and thermal gradient cal- the values for Qo, b, and A in the lithosphere. The final model is 228.5 km long
culations (Blackwell et al., 2010). The final data set consisted of 1454 tempera- and 9.2 km deep at 100 m resolution and includes basic facies information, well
ture-depth pairs, the locations of which are shown in Figure 3. locations, faults, and other structures in the basin. The model was written in
BHT data are generally low quality compared to equilibrium measure-ments. MATLAB (www.mathworks.com) using the finite-differ-ence method.
The reason for the low-quality stems from the methodology used to collect the
data. Measurements are generally taken shortly after drilling has been completed The 2-D advection-diffusion equation is:
on a well. The circulation of drilling fluids from the surface during and after drilling
has a significant effect on the near field in situ tem-peratures. The effect of fluid ∂T ∂T ∂T ∂ ∂2T ∂ ∂2T
ρCp +u +v = k + k + A, (3)
circulation is generally a reduction of the tempera tures in the wellbore for the ∂t ∂x ∂z ∂x ∂x 2 ∂z ∂z 2
sections deeper than ~600–1000 m (Blackwell and Steele, 1989; Förster et al.,
1997; McKenna and Blackwell, 2001). This effect needs to be removed in order where T is temperature (°C), t is time (s), r is density (kg/m 3), Cp is constant
to calculate a true thermal gradient for the well. Chapman et al. (1984) found that pressure heat capacity (J kg –1 K–1), k is thermal conductivity (W m –1 K–1), A is heat
the correction for measurements made soon production (mW/m3), x and z are spatial dimensions (m), and u and v are

GEOSPHERE | Volume 11  | Number 5 Frone et al.  | Heat flow and thermal modeling of the Appalachian Basin 1283
Research Paper
advection velocities (m/s) in the x and z directions, respectively. Thermal con- additional model was tested where the crust was thinned based on the thick-

ductivity values for given lithologies were populated from data compiled by ness of overlying sedimentary formations. In this model, the heat flow enter-
Carter et al. (1998) and Gallardo and Blackwell (1999) for Paleozoic sedimentary ing the basin varied and results in lower heat flow and temperatures in the
rocks and from unpublished measurements on cuttings from the Appalachian deepest part of the basin (Rome Trough). The data do not show a decrease in
Basin. Density and heat capacity are held constant and determined by averag- temperatures in the Rome Trough area, so this model of the crust was not ex-
ing data for limestone, sandstone, and shale from Robertson (1988) (Table 1). amined further. The model used for the crust is a silicic upper crust and a mafic
Heat production values in the basement and the sedimentary rocks are lower crust (Drury, 1989; Christensen and Mooney, 1995). The terms mafic and
included in the model. Heat is produced in the crust by the natural decay of silicic in this case refer only to the relative amount of heat producing elements
primarily U, Th, and K (Beardsmore and Cull, 2001). Heat production values for found in each layer, with the mafic layer having lower and the silicic having
sediments were determined using a simple relationship established by Bücker higher heat production. Each model iteration was run to steady state with
and Rybach (1996) between heat generation and the natural gamma ray, GR varied input values for reduced heat flow (Qo), upper crust heat production
(API), response from well logs (Equation 4). (A), and initial upper crustal thickness (b). The heat flow at the sediment-base-
ment contact was constant across the model due to the constant values used
A = 0.0158(GR − 0.8). (4) for Qo, A, and b.
The boundary conditions were Dirichlet at the surface (T = 10 °C), Neumann
The model heat production values incorporate the digital natural gamma at the base, and the side boundaries were open. The basal Neumann boundary
logs used by Ryder et al. (2009) to construct the cross section. Formation top (constant heat flow) was varied according to the input values for Qo, b, and A.
data were used to determine the rock type represented by the natural gamma When steady-state temperatures were reached, calculated temperatures were
log. The heat production results from Equation 4 were averaged for each rock compared to measured temperatures. Residuals were computed using Equa-
type (Table 1). tion 5. Residuals from the steady-state temperature models were calculated for
Temperature values within a distance of 25 km from either side of the cross the equilibrium temperature wells and BHTs. A higher confidence was given
section were projected onto the 2-D cross section (red and blue dots in Fig. 3). to the temperature-depth (T-D) well residuals, as these temperatures are as-
This resulted in 143 BHT data points and 6 equilibrium temperature logs (Figs. sumed to be equilibrium measurements while BHT measurements were cor-
4 and 5). rected values that likely have an error of ±10 °C (Richards et al., 2012).
For the model, the crust below the sedimentary cover had a constant
Residual = ∑ (known temperature − model temperature)2 .
thickness that was defined by b in Equation 1. This is similar to the way the (5)
lithosphere below foreland basins is modeled as an elastically deformed plate
2
known temperature
(Flemings and Jordan, 1989; Garcia-Castellanos et al., 1997). This model as-
sumes constant heat flow and heat production from the basement rocks. An Using constraints from Thakur (2011) for Qo (33 ± 1 mW/m2) and b (7.5 ±
0.2 km) a range of models was tested with different values for A. The basement
A value was varied between 1.5 and 3.5 mW/m3, the range for granitoid rocks
TABLE 1. APPALACHIAN BASIN THERMAL PROPERTIES (Vilà et al., 2010).
Variable k A ρ1 Cp*
Thermal Heat Density Specific RESULTS
Lithology conductivity production heat capacity
(kg/m3)
(W/m K) (µW/m3) (J/kg)
The model results showed that for the constrained Qo and b values for
Sandstone 2.8 0.85 2500 900
the eastern United States from Thakur (2011) and basement heat production
Shale 1.8 1.9 2500 900
values between 2.0 and 3.0 µW/m3, the model reproduced the observed tem-
Black shale 1.4 1.9 2500 900
Limestone 2.5 0.5 2500 900 peratures within this section of the basin (Figs. 5A and 6). Heat flow at the
Dolomite 3 0.5 2500 900 sediment-basement contact ranged from 49 to 55 mW/m2. This heat flow con-
Evaporite 4.7 0.5 2500 900 straint can be further used to constrain future crustal studies in the region. In
Granite 2.7 Variable 2500 900 the model this value was controlled by the input values for Qo, b, and A, but
Note: Thermal conductivity values are from Carter et al. (1998) and unpublished data different distributions and values of heat production and reduced heat flow
for Paleozoic sedimentary rocks. can also result in the same heat flow at the base of the sedimentary basin. The
*Values for density and specific heat capacity are averages based on Robertson range in heat production values is consistent with granitoid rock types (Vilà
(1988). et al., 2010). The model shows increasing surface heat flow from west to east.

GEOSPHERE | Volume 11  | Number 5 Frone et al.  | Heat flow and thermal modeling of the Appalachian Basin 1284
Research Paper

A A′

Temperature Residual (°C)


Model Hotter Well Hotter
20 70
A 65

Heat Flow (mW/m )2


10
0 60

–10 55

–20
B 50

1000 WV-14 WV-8


WV-15 WV-6 WV-1WV-3
0

25°C C
25°

–1000 50°C 50°C 50°C

Figure 5. Thermal conductivity model and


modeled temperature results along cross-
–2000 75°C 75°C section A–A′. (A) Model tempera-ture
75°C
residuals (in °C) and model surface heat
flow, Q = 52 ± 3 mW/m2 at the sedi ment-
basement contact. (B) Model ther-mal
(m)

–3000 100°C conductivity values and calculated


0° temperature results for a sediment-base-
10 ment contact heat flow of 52 ± 3 mW/m2.
C Triangles are temperature values from
C equilibrium wells, circles are bottom-hole
Elevation

100°C temperature well measurements. V.E.—


vertical exaggeration.
–4000 125°C 125°
150°C

°C
125
–5000
°C

M 150°C
15 to 25
e
25 to 50
a –6000 50 to 75
s
150°C 75 to 100
u
r 100 to 125 175°C
e 125 to 165
d –7000
Thermal Conductivity W/m/K
T
e
m
p
e
r
a
t
u
r
e
s
175°C

200°C
175°C

1.4 1.8 2.5 2.6 2.7 2.8 2.9 3.0 4.7 160000 180000 200000 220000
–8000
0 20000 40000 60000 80000 100000 120000 140000
Distance Along D–D′ Section (m)

GEOSPHERE | Volume 11  | Number 5 Frone et al.  | Heat flow and thermal modeling of the Appalachian Basin
1285
Research Paper

TemperatureResidual(°C)
30 85

(mW/mFlowH
25 Model Heat Flow Profile 80
20 2010 Heat Flow Profile 75

eat2

)
15 Temperature Residuals 70
10 65
5 60
0 55
–5 50
–10 45
–15 40
–20 35
0 25000 50000 75000 100000 125000 150000 175000 200000 225000
Distance (m)
Figure 6. Model residuals and heat flow. Comparison of current model temperature residuals and heat flow profiles calculated
from numerical model with values from Frone and Blackwell (2010) map. Red dashed lines are temperature residuals are the
same as in Figure 5. The highest heat flow from the 2010 heat flow map corresponds to an area with the largest range of model
residuals.

The thicker sedimentary package occurring on the eastern side of the model 2010) or a Grenville-aged suture between Laurentia and an accreted terrain
results in more heat production and therefore higher surface heat flow. (Tohver et al., 2004; Mueller et al., 2008). Steltenpohl et al. (2010) suggested a
The surface heat flow calculated along the profile is compared with the heat right-lateral offset of ~220 km that is continuous from New York to Alabama.
flow from Frone and Blackwell (2010), who used primarily BHT data and a simpli- Both of these interpretations imply that a different basement lithology may
fied 1-D stratigraphic model to determine heat flow within the study area (Fig. 6). be present under eastern West Virginia. Heat flow results for southwestern
The region of the highest heat flow values from Frone and Blackwell (2010) cor- Pennsylvania from Shope (2012) show elevated heat flow values within the
respond with the largest temperature residual range from the 2-D model. Rome Trough, ~200 km north northeast of the modeled cross section, that is
Four scenarios are presented to attempt to explain the observed scatter suggested to be the result of elevated basement heat production. This result
in the measured temperatures between wells WV-3 and WV-8 (Figs. 5 and 6). is consistent with the interpretation of the magnetic data. This case was tested
by allowing the heat production in the basement to vary at the eastern bound-
Heat Production ary of the Rome Trough by 25%–200%. The results of this modeling show that
an increase in heat production of 25%–75% in the eastern basement block still
fits the observed data, but does not produce the higher temperatures observed.
In the first scenario, there is higher heat flow due to higher heat production
in the basement or a thicker upper crust beneath the eastern side of the basin.
From Equation 1, it is clear that if Qo is constant over large areas, then changes
in Q are related to changes in b, A, or a combination of the two. For example, Q
for a 10-km-thick upper crust, a uniform increase in heat production of 1 mW/m3
results in a surface heat flow increase of 10 mW/m 2 (Fig. 7). The same increase b A=1 µW/m3 A=2 µW/m3 A=1 µW/m3
in heat flow could also be generated from a constant heat production and an A=1 µW/m3
increase of the upper crustal thickness from 10 to 20 km (Fig. 7). From pre-
liminary receiver function analysis of the Earthscope Transportable Array over Qo
West Virginia, Pennsylvania, Ohio, Virginia, and Maryland, crustal thickness
within the Appalachian Basin is 45 ± 5 km (Crotwell and Owens, 2005; Trabant Figure 7. Graphical representation of the effect of changes in A (heat production) or b (thickness

et al., 2012). Therefore, in this study variations in heat flow are interpreted as of heat-producing upper crust) on Q (surface heat flow). Qo, reduced heat flow, is constant in
predominantly changes in basement heat production. both cases. The dashed line shows the theoretical heat flow if the heat production boundary is
at the surface; the solid line shows the general shape of the surface heat flow if the heat-pro-
Magnetic anomalies suggest that the eastern boundary of the Rome Trough ducing bodies are below the surface. The sharpness of the anomaly depends on the depth of the
is part of a large-scale strike-slip fault (King and Zietz, 1978; Steltenpohl et al., bodies; deeper bodies result in a smoother anomaly (Simmons, 1967).

GEOSPHERE | Volume 11  | Number 5 Frone et al.  | Heat flow and thermal modeling of the Appalachian Basin 1286
Research Paper

Higher heat production values can fit the higher observed BHT values; how-ever, cored anticlines (Ryder et al., 2009). Structures like these can affect tempera-
the models no longer fit the equilibrium temperature data in wells WV-3 and WV- tures in two ways. Isotherms can be either elevated or depressed at the peak of
8 and this does not explain the lower BHT values in the same area. For heat anticlines depending on the thermal conductivity of the deformed forma-tions
production variations in the basement to explain the higher observed BHT values (Guyod, 1946). Elevated isotherms are within the anticlines in this model (Fig. 5).
between these wells, the eastern basement block would need to be divided into The equilibrium well WV-8 is located on the crest of the Deer Park– Leadmine
subblocks. There is evidence for down to the east normal faulting in the eastern anticline and shows elevated temperatures compared to the other equilibrium
basement block to the south (Ryder et al., 2008); however, the offset of these wells (Fig. 5). A few of the BHT values in this region between WV-3 and WV-8
faults decreases to the north and is thought to terminate ~25 km north of the show the largest deviations from the modeled values.
modeled cross section (Shumaker and Wilson, 1996). Higher heat production The second temperature effect that could be caused by the structure is lo-
values in smaller subblocks were tested to attempt to explain the elevated BHT calized fluid flow along thrust faults within the anticlines. Bonini (2007) showed
values. Results show that an increase in heat production of 50%– 70% from the that for thrust fault–related folds, a reduction in lithostatic pressure can cause
eastern margin of the Rome Trough to the normal fault shown by Shumaker and fluid overpressure and migration along faults. The reduction in lithostatic pres-
Wilson (1996) provides a better fit to the elevated BHT values; however, this sure is caused by surface erosion. The migration of fluids from depth along faults
model also produces higher than observed temperatures in WV-3 and WV-8 and results in elevated temperatures along the faults. The magnitude of the
does not explain the low BHT values. Thus first-order changes in heat production temperature effect in the rock surrounding the fault is dependent upon the
in the basement do not provide a simple explanation for the variation in BHT temperature of the fluid and the time scale of the flow. Similarly, migration of
values between WV-3 and WV-8. fluids from the surface downward may result in depressed temperatures. The
scale of the temperature effect is dependent upon the fluid volume and velocity.
The migration of fluids along faults in the anticlines is likely not uni-form along the
Lateral Groundwater Flow strike of the structure due to heterogeneities in the subsurface. Advection of
fluids along fault pathways is not included in this model. This is an unlikely
The second scenario is that intrabasin fluid flow could result in the observed process in the Appalachian Basin for two reasons. First, slip along a fault plane
temperatures. This case would imply that the thermal anomaly is caused by can result in the formation of phyllosilicate minerals, which in turn decrease the
advective heat transport rather than conduction. A large-scale example of ad- permeability of the fault (Cavailhes et al., 2013). Second, the lack of recorded
vective heat transport is the heat flow anomaly in South Dakota and Nebraska seismicity in the region suggests that faults have not been reactivated, resulting
(Gosnold, 1990). That anomaly is caused by gravitationally driven groundwater in sealed faults that are likely to inhibit fluid migration.
flow, laterally transporting heat within the sediments. It has been shown that
velocities of 10–8 m/s were needed to produce the observed heat flow anom-aly.
Darcy velocities for the Appalachian Basin from Gupta and Bair (1997) are used Erosion
to model the effects of advection on the basin temperatures. Using their
maximum velocity value of 3.05 × 10 –12 m/s, the maximum temperature change The fourth scenario for the observed temperature scatter is related to ero-
within the basin is ~1 °C compared to pure conduction. The temperature effect of sion or sedimentation rates with in the basin. Erosion of the upper surface of a
advection at this velocity is small and the change at any one well is within the basin results in an uplift of isotherms in the subsurface while sedimentation
error of the measurement. Advection may still cause localized temperature depresses isotherms, resulting in a higher or lower heat flow, respectively
effects; however, basin-scale advection as currently understood cannot repro- (Jaeger, 1965). This results in a high or low temperature gradient. This gradient
duce the observed temperatures. will persist if the rate of erosion or deposition is high enough to outpace the
dominant heat transport method. The degree to which erosion or sedimenta-tion
affects the heat flow is controlled by the thermal diffusivity and the rate of erosion
Structural Complexity or sedimentation.
From fission track studies (Roden, 1991; Blackmer et al., 1994; Boettcher
The third scenario of the anomalous BHT values is related to the struc-tures and Milliken, 1994) and radiogenic isotope studies (Matmon et al., 2003), ero-
in the basin. The anomalous values are generally located between the Etam and sion rates for the Appalachian Basin from 265 to 30 Ma are between 10 and 28
Deer Park–Leadmine anticlines. The high-amplitude anticlines are the result of m/m.y. From the Miocene to the present, erosion rates increased to 50–100
thin-skinned deformation during the Late Mississippian–Permian Alleghanian m/m.y., possibly due to changes in mantle dynamics (Miller et al., 2013; Black-
orogeny (Wilson and Shumaker, 1992; Kulander and Ryder, 2005). The anticlines mer et al., 1994; Boettcher and Milliken, 1994; Galloway et al., 2011). In order to
were formed by compression and deformation along décolle-ment horizons. The constrain the maximum change in temperatures that could be caused by erosion,
results of this thrusting are high-amplitude thrust fault– a rate of 100 m/m.y. for 20 Ma was used.

GEOSPHERE | Volume 11  | Number 5 Frone et al.  | Heat flow and thermal modeling of the Appalachian Basin 1287
Research Paper

In this study a numerical solution was used to estimate the effect of ero-sion were affected by local advection caused by fluid migration along thrust faults
on basin temperatures. The solution instantaneously erodes 100 m/m.y. of rock. within anticlines, and erosion. Of these potential causes, complex structure and
The solution shows a decreasing effect with depth, with a maximum temperature 3-D effects best explain the variability in the BHT values and the calculated heat
increase of 6.8% at the surface and an average temperature in-crease of 6.7% in flow values.
the depth range 1–6 km above a steady-state background temperature
distribution. This is for a 1-D case, which is preliminarily correct for the study
area, so that the temperature increases affect all wells in the study more or less ACKNOWLEDGMENTS
equally (Blackmer et al., 1994). For this portion, and likely all, of the Appalachian We thank Associate Editor Anna Crowell, Paul Morgan, and two anonymous reviewers whose
Basin, erosion has an effect on subsurface temperatures and heat flow; however, questions and comments helped to improve this manuscript. The Roy M. Huffington Department of
Earth Sciences and the Southern Methodist University Geothermal Laboratory provided fund-ing for
it does not appear to be a viable solution to the variable measured temperatures. this research.

REFERENCES CITED
CONCLUSIONS American Association of Petroleum Geologists, 1994, DP-AAPG DataRom (CSDE, COSUNA,
GSNA): American Association of Petroleum Geologists Multimedia CD-ROM, 569 p.
Andrews-Speed, C.P., Oxburgh, E.R., and Cooper, B.A., 1984, Temperatures and depth-depen-dent
Previous heat flow studies in the Appalachian Basin have had limited avail-
heat flow in western North Sea: American Association of Petroleum Geologists Bulle-tin, v. 68, p.
able borehole data (Blackwell and Richards, 2004; Joyner, 1960). Use of BHT 1764–1781.
data can provide a large data set that, when checked against equilibrium data, Beardsmore, G.R., and Cull, J.P., 2001, Crustal heat flow: A guide to measurement and model-
can be a useful source of temperature data. Incorporating both equilibrium and ling: Cambridge, Cambridge University Press, 324 p.
Birch, F., Roy, R.F., and Decker, E.R., 1968, Heat flow and thermal history in New England and New
BHT data with heat flow and heat production provides a more detailed analysis of York, in Zen, E., ed., Studies of Appalachian geology: Northern and maritime: New York, Wiley, p.
heat flow and basin temperatures in the northern West Virginia portion of the 437–451.
Appalachian Basin. Blackmer, G.C., Omar, G.I., and Gold, D.P., 1994, Post-Alleghanian unroofing history of the Appa
lachian Basin, Pennsylvania, from apatite fission track analysis and thermal models: Tec tonics,
This study shows that the thermal regime in this portion of the Appalachian v. 13, p. 1259–1276, doi:10.1029/94TC01507.
Basin is dominated by conductive heat flow. A uniform thickness model of the Blackwell, D.D., and Richards, M., 2004, Geothermal map of North America: American Associa-tion of
crust following the Q-A relationship can explain the heat flow and heat pro- Petroleum Geologists, scale 1:6,500,000.
Blackwell, D.D., and Spafford, R.E., 1987, Experimental methods in continental heat flow, in
duction in the crust below the basin. Results show that for Qo = 33 ± 1 mW/m 2 Celotta, R., et al., eds., Geophysics—Field measurements: Methods in Experimental Physics
and b = 7.5 ± 0.5 km the heat production in the basement is between 2 and 3 Volume 24, Part B: Orlando, Florida, Academic Press, p. 189–226, doi:10.1016/S0076-695X
µW/m3 and the heat flow at the base of the basin is between 49 and 55 mW/m 2. (08)60599-2.
Blackwell, D.D., and Steele, J.L., 1989, Thermal conductivity of sedimentary rocks: Measure-ment
The heat flow value at the base of the basin had the strongest influence on the
and significance, in Naeser, N.D., et al., eds., Thermal history of sedimentary basins: Methods and
modeled temperatures. Therefore, the values for Qo, b, and A could vary outside case histories: New York, Springer, p. 12–36, doi:10.1007/978-1-4612-3492-0_2 Blackwell, D.D.,
of the ranges used from Thakur (2011) as long as the heat flow at the top of the Steele, J.L., and Carter, L.S., 1990, Heat flow patterns of the North American continent: A discussion
of the DNAG geothermal map of North America, in Slemmons, D.B., et al., eds., Neotectonics of
basement is within the range of 49–55 mW/m 2. North America: Boulder, Colorado, Geological Society of Amer-
Models using the Q-A relationship for heat flow within the crust show that ica, Decade Map Volume 1, p. 423–436.
most of the measured temperatures within the basin can be matched with a Blackwell, D.D., Richards, M., Batir, J., Frone, Z., and Park, J., 2010, New geothermal resource map
of the northeastern U.S. and technique for mapping temperature at depth: Geothermal
simple conductive heat flow model with uniform basement properties. Varia-tion
Resources Council Transactions, v. 34, p. 313–318.
in basement heat production values to 10% does not significantly affect the Boettcher, S.S., and Milliken, K.L., 1994, Mesozoic–Cenozoic unroofing of the southern Appala-chian
results of the model. The modeled surface heat flow values between 55 and 63 Basin: Apatite fission track evidence from middle Pennsylvanian sandstones: Journal of Geology,
v. 102, p. 655–668, doi:10.1086/629710.
mW/m2 and are within the expected range for Paleozoic age sedimen-tary rocks
Bonini, M., 2007, Interrelations of mud volcanism, fluid venting, and thrust-anticline folding: Ex-
based on the Pollack et al. (1993) analysis. amples from the external Northern Apennines (Emilia-Romagna, Italy): Journal of Geophys-ical
Heat flow calculations from Frone and Blackwell (2010) using BHT data show Research, v. 112, B08413, doi:10.1029/2006JB004859.
Bücker, C., and Rybach, L., 1996, A simple method to determine heat production from gamma-ray
that the heat flow east of the Rome Trough is ~40% higher than within and west
logs: Marine and Petroleum Geology, v. 13, p. 373–375, doi:10.1016/0264-8172(95)00089-5.
of the trough. Numerical models of a 2-D cross section of the ba-sin are Carter, L.S., Kelley, S.A., Blackwell, D.D., and Naeser, N.D., 1998, Heat flow and thermal history of
consistent with a heat flow increase from west to east; however, the predicted the Anadarko Basin, Oklahoma: American Association of Petroleum Geologists Bulletin, v. 82, p.
291–316.
increase is closer to 15% and is more variable. Thus, the heat flow anomaly that
Cavailhes, T., Soliva, R., Labaume, P., Wibberley, C., Sizun, J.P., Gout, C., Charpentier, D., Chau-vet,
was previously reported appears to be overestimated. The val-ues may have A., Scalabrino, B., and Buatier, M., 2013, Phyllosilicates formation in fault rocks: Implica-tions for
been overestimated due to variation in basement heat produc-tion, large-scale dormant fault-sealing potential and fault strength in the upper crust: Geophysical Research
Letters, v. 40, p. 4272–4278, doi:10.1002/grl.50829.
basin advection, the inclusion of BHT measurements that

GEOSPHERE | Volume 11  | Number 5 Frone et al.  | Heat flow and thermal modeling of the Appalachian Basin 1288
Research Paper

Chapman, D.S., Keho, T.H., Bauer, M.S., and Picard, M.D., 1984, Heat flow in the Uinta Basin King, E.R., and Zietz, I., 1978, The New York–Alabama Lineament: Geophysical evidence for a major
determined from bottom-hole temperature (BHT) data: Geophysics, v. 49, p. 453–466, doi: crustal break in the basement beneath the Appalachian Basin: Geology, v. 6, p. 312– 318,
10.1190/1.1441680. doi:10.1130/0091-7613(1978)6<312:TNYLGE>2.0.CO;2.
Christensen, N.I., and Mooney, W.D., 1995, Seismic velocity structure and composition of the Kulander, C.S., and Ryder, R.T., 2005, Regional seismic lines across the Rome Trough and Allegheny
continental crust: A global view: Journal of Geophysical Research, v. 100, no. B6, p. 9761– 9788, Plateau of northern West Virginia, western Maryland, and southwestern Pennsylvania: U.S.
doi:10.1029/95JB00259. Geological Survey Map I-2791, 9 p.
Costain, J.K., Speer, J.A., Glover, L., Perry, L., Dashevsky, S., and McKinney, M., 1986, Heat flow in Lachenbruch, A.H., 1968, Preliminary geothermal model of the Sierra Nevada: Journal of Geo-
the Piedmont and Atlantic Coastal Plain of the southeastern United States: Journal of physical Research, v. 73, p. 6977–6989, doi:10.1029/JB073i022p06977.
Geophysical Research, v. 91, no. B2, p. 2123–2135, doi:10.1029/JB091iB02p02123. Matmon, A., Bierman, P.R., Larsen, J., Southworth, S., Pavich, M., Finkel, R., and Caffee, M., 2003,
Crotwell, H.P., and Owens, T.J., 2005, Automated receiver function processing: Seismological Erosion of an ancient mountain range, the Great Smoky Mountains, North Carolina and
Research Letters, v. 76, p. 702–709, doi:10.1785/gssrl.76.6.702. Tennessee: American Journal of Science, v. 303, p. 817–855, doi:10.2475/ajs.303.9.817.
Decker, E.R., Heasler, H.P., Buelow, K.L., Baker, K.H., and Hallin, J.S., 1988, Significance of past and McKenna, J.R., and Blackwell, D.D., 2001, Thermal regime of a large midcontinent oil field (Butler
recent heat flow and radioactivity studies in the southern Rocky Mountains region: Geological Country, Kansas) from high resolution temperature logs, in Johnson, K.S., and Merriam, D.F.,
Society of America Bulletin, v. 100, p. 1851–1885, doi:10.1130/0016-7606(1988)100 eds., Petroleum systems of sedimentary basins in the southern Midcontinent: 2000 Symposium
<1851:SOPARH>2.3.CO;2. Proceedings, Oklahoma Geological Survey and U.S. Department of Energy, p. 151–162.
Deming, D., and Chapman, D.S., 1988, Heat flow in the Utah-Wyoming thrust belt from analysis of
bottom-hole temperature data measured in oil and gas wells: Journal of Geophysical Research, Milici, R.C., 1996, Stratigraphic history of the Appalachian Basin, in Roen, J.B., and Walker, B.J.,
v. 93, no. B11, p. 13,657–13,672, doi:10.1029/JB093iB11p13657. eds., The atlas of major Appalachian gas plays: West Virginia Geological and Economic Survey
Drury, M.J., 1989, The heat flow-heat generation relationship: Implications for the nature of con- Publication V-25, p. 4–7.
tinental crust: Tectonophysics, v. 164, p. 93–106, doi:10.1016/0040-1951(89)90003-6. Miller, S.R., Sak, P.B., Kirby, E., and Bierman, P.R., 2013, Neogene rejuvenation of central Appa-
Ettensohn, F.R., 1985, The Catskill Delta complex and the Acadian orogeny: A model, in Wood-row, lachian topography: Evidence for differential rock uplift from stream profiles and erosion rates:
D.L., and Sevon, W.D., eds., The Catskill Delta: Geological Society of America Special Paper Earth and Planetary Science Letters, v. 369, p. 1–12, doi:10.1016/j.epsl.2013.04.007.
201, p. 39–50, doi:10.1130/SPE201-p39. Mueller, P.A., Kamenov, G.D., Heatherington, A.L., and Richards, J., 2008, Crustal evolution in the
Flemings, P.B., and Jordan, T.E., 1989, A synthetic stratigraphic model of foreland basin de-velopment: Southern Appalachian orogen: Evidence from Hf isotopes in detrital zircons: Journal of Geology,
Journal of Geophysical Research, v. 94, no. B4, p. 3851–3866, doi:10.1029 /JB094iB04p03851. v. 116, p. 414–422, doi:10.1086/589311.
Förster, A., Merriam, D.F., and Davis, J.C., 1997, Spatial analysis of temperature (BHT/DST): Data Nyblade, A.A., and Pollack, H.N., 1993, A global analysis of heat flow from Precambrian terrains:
and consequences for heat flow determination in sedimentary basins: Geologische Rund-schau, Implications for the thermal structure of Archean and Proterozoic lithosphere: Journal of
v. 86, p. 252–261, doi:10.1007/s005310050138. Geophysical Research, v. 98, no. B7, p. 12,207–12,218, doi:10.1029/93JB00521.
Frone, Z., and Blackwell, D.D., 2010, Geothermal map of the northeast United States and the West Patchen, D.G., et al., 2006, A geologic play book for Trenton-Black River Appalachian Basin Ex-
Virginia thermal anomaly: Geothermal Resources Council Transactions, v. 34, p. 339–344. ploration: Department of Energy DE-FC26-03NT41856, http://www.wvgs.wvnet.edu/www/tbr
Gallardo, J., and Blackwell, D.D., 1999, Thermal structure of the Anadarko Basin: American Asso- /project_reports.asp (accessed April 2015).
ciation of Petroleum Geologists Bulletin, v. 83, p. 333–361. Pollack, H.N., 1982, The heat flow from the continents: Annual Review of Earth and Planetary
Galloway, W.E., Whiteaker, T.L., and Ganey-Curry, P., 2011, History of Cenozoic North American Sciences, v. 10, p. 459, doi:10.1146/annurev.ea.10.050182.002331.
drainage basin evolution, sediment yield, and accumulation in the Gulf of Mexico Basin: Pollack, H.N., Hurter, S.J., and Johnson, J.R., 1993, Heat flow from the Earth’s interior: Analysis
Geosphere, v. 7, p. 938–973, doi:10.1130/GES00647.1. of the global data set: Reviews of Geophysics, v. 31, p. 267–280, doi:10.1029/93RG01249.
Gao, D., Shumaker, R.C., and Wilson, T.H., 2000, Along-axis segmentation and growth history of the Richards, M., Blackwell, D.D., Williams, M., Frone, Z., Dingwall, R., Batir, J., and Chickering, C.,
Rome Trough in the central Appalachian Basin: American Association of Petroleum Geologists 2012, Proposed reliability code for heat flow sites: Geothermal Resources Council Transac-tions,
Bulletin, v. 84, p. 75–99. v. 36, p. 211–217.
Garcia-Castellanos, D., Fernàndez, M., and Torne, M., 1997, Numerical modeling of foreland basin Robertson, E.C., 1988, Thermal properties of rocks: U.S. Geological Survey Open-File Report 88–
formation: A program relating thrusting, flexure, sediment geometry and lithosphere rheol ogy: 441, 106 p.
Computers & Geosciences, v. 23, p. 993–1003, doi:10.1016/S0098-3004(97)00057-5. Roden, M.K., 1991, Apatite fission-track thermochronology of the Southern Appalachian Basin:
Gosnold, W.D., 1990, Heat flow in the Great Plains of the United States: Journal of Geophysical Maryland, West Virginia, and Virginia: Journal of Geology, v. 99, p. 41–53, doi:10.1086 /629472.
Research, v. 95, no. B1, p. 353–374, doi:10.1029/JB095iB01p00353.
Gupta, N., and Bair, E.S., 1997, Variable-density flow in the midcontinent basins and arches region of Roy, R.F., Blackwell, D.D., and Birch, F., 1968, Heat generation of plutonic rocks and continental heat
the United States: Water Resources Research, v. 33, p. 1785–1802, doi:10.1029/97WR01199. flow provinces: Earth and Planetary Science Letters, v. 5, p. 1–12, doi:10.1016/S0012
Guyod, H.C., 1946, Temperature well-logging heat conduction: Oil Weekly, v. 123, BM-IC-7414. -821X(68)80002-0.
Hallock, W.B., 1897, Subterranean temperatures at Wheeling W. Va., and Pittsburgh, Pa., in Moses, Roy, R.F., Blackwell, D.D., and Decker, E.R., 1972, Continental heat flow, in Robertson, E.C., eds.,
A.J., et al., eds., Columbia University School of Mines Quarterly, v. 18, p. 148–153. Harrison, The nature of the solid Earth: New York, McGraw-Hill, p. 506–543.
W.E., Luza, K.V., Prater, M.L., and Cheung, P.K., 1983, Geothermal resource assessment Ryder, R.T., 2006, Appalachian Basin Province (067): U.S. Geological Survey, 86 p., http://
in Oklahoma: Oklahoma Geological Survey Special Publication 83–1, 46 p. certmapper.cr.usgs.gov/data/noga95/prov67/text/prov67.pdf (accessed April 2015).
Hendry, R., Hilfiker, K., Hodge, D., Morgan, P., Swanberg, C., and Shannon, S.S., 1982, Geother-mal Ryder, R.T., Swezey, C.S., Crangle, R.D., and Trippi, M.H., 2008, Geologic cross section E-E’ through
investigations in West Virginia: Los Alamos National Laboratory LA-9558-HDR, 57 p. the Appalachian Basin from the Findlay Arch, Wood County, Ohio to the Valley and Ridge
Horner, D.R., 1951, Pressure build-up in wells: The Hague, Third World Petroleum Congress, WPC- Province, Pendleton County, West Virginia: U.S. Geological Survey Scientific Investi-gation Map
4135, 19 p. 2985, 48 p.
Jaeger, J.C., 1965, Application of the theory of heat conduction to geothermal measurements, in Lee, Ryder, R.T., Crangle, R.D., Trippi, M.H., Swezey, C.S., Lentz, E.E., Rowan, E.L., and Hope, R.S.,
H.K., ed., Terrestrial heat flow: American Geophysical Union Geophysical Monograph 8, p. 7–23, 2009, Geologic cross section D-D’ through the Appalachian Basin from the Findlay Arch,
doi:10.1029/GM008p0007. Sandusky County, Ohio, to the Valley and Ridge Province, Hardy County, West Virginia: U.S.
Jaupart, C., and Mareschal, J.C., 1999, The thermal structure and thickness of continental roots: Geological Survey Scientific Investigations Map 3067, 52 p.
Lithos, v. 48, p. 93–114, doi:10.1016/S0024-4937(99)00023-7. Shope, E.N., 2012, A detailed approach to low-grade geothermal resources in the Appalachian Basin
Joyner, W.B., 1960, Heat flow in Pennsylvania and West Virginia: Geophysics, v. 25, p. 1229–1241, of New York and Pennsylvania: Heterogeneities within the geologic model and their effect on
doi:10.1190/1.1438811. geothermal resource assessment [M.S. thesis]: Ithaca, New York, Cornell Univer-sity, 181 p.

GEOSPHERE | Volume 11  | Number 5 Frone et al.  | Heat flow and thermal modeling of the Appalachian Basin 1289
Research Paper

Shumaker, R.C., and Wilson, T.H., 1996, Basement structure of the Appalachian foreland in West Tohver, E., Bettencourt, J.S., Tosdal, R., Mezger, K., Leite, W.B., and Payolla, B.L., 2004, Terrane
Virginia: Its style and effect on sedimentation, in van der Pluijm, B.A., and Catacosinos, P.A., transfer during the Grenville orogeny: Tracing the Amazonian ancestry of southern Appa-lachian
eds., Basement and basins of eastern North America: Geological Society of America Special basement through Pb and Nd isotopes: Earth and Planetary Science Letters, v. 228, p. 161–176,
Paper 308, p. 139–155, doi:10.1130/0-8137-2308-6.139. doi:10.1016/j.epsl.2004.09.029.
Simmons, G., 1967, Interpretation of heat flow anomalies 1. Contrasts in heat production: Re-views of Trabant, C., Hutko, A.R., Bahavar, M., Karstens, R., Ahern, T., and Aster, R., 2012, Data products at
Geophysics, v. 5, p. 43–52, doi:10.1029/RG005i001p00043.
the IRIS DMC: Stepping-stones for research and other applications: Seismological Research
Spicer, H.C., 1964, A compilation of deep Earth temperature data, U.S.A. 1910–1945: U.S. Geologi
Letters, v. 83, p. 846–854, doi:10.1785/0220120032.
cal Survey Open-File Report 64-147, 3 sheets.
Van Orstrand, C.E., 1935, Normal geothermal gradient in United States: American Association of
Steltenpohl, M.G., Zietz, I., Horton, J.W., Jr., and Daniels, D.L., 2010, New York–Alabama Linea-
ment: A buried right-slip fault bordering the Appalachians and Midcontinent North America: Petroleum Geologists Bulletin, v. 19, p. 78–115.
Geology, v. 38, p. 571–574, doi:10.1130/G30978.1. Vilà, M., Fernández, M., and Jiménez-Munt, I., 2010, Radiogenic heat production variability of some
Swezey, C.S., 2002, Regional stratigraphy and petroleum systems of the Appalachian Basin, North common lithological groups and its significance to lithospheric thermal modeling: Tectonophysics,
America: U.S. Geological Survey Geologic Investigations Series Map I-2768, 1 p. v. 490, p. 152–164, doi:10.1016/j.tecto.2010.05.003.
Thakur, M., 2011, Stabilization temperature of the Archean cratons: Implication for thermal re-gime, Wilson, T.H., and Shumaker, R.C., 1992, Broad Top thrust sheet: An extensive blind thrust in the
secular variation of surface heat flow and lithospheric thickness in Precambrian ter-rains [Ph.D. central Appalachians (1): American Association of Petroleum Geologists Bulletin, v. 76, p. 1310–
thesis]: Dallas, Texas, Southern Methodist University, 214 p. 1323.

GEOSPHERE | Volume 11  | Number 5 Frone et al.  | Heat flow and thermal modeling of the Appalachian Basin 1290

Вам также может понравиться