Sei sulla pagina 1di 144

INTRODUCTION

Soil vapor extraction (SVE) is perhaps the most successful method of remediating

the unsaturated zone, with thousands of contaminated sites benefiting from this

technology. However, SVE is not without limitations, as the performance of this

technology can be hampered by commonly occurring site conditions. Applications of

SVE are limited to the recovery of volatile organic compounds from regions where air

flow can be induced at a rate of at least one pore volume per day, according to Gierke

[2000]. Where airflow is rapid, the costs associated with treating contaminated soils can

be as low as a few dollars per cubic meter. Low permeability, or clay-rich, soils will

inhibit fluid flow and reduce the systems ability to flush pore spaces, and in some cases,

increase costs to more than a thousand dollars per cubic meter [Gierke 2000].

Some depositional environments create silts or clays with thin interbedded sand

layers [Prothero 1977]. An SVE well intersecting a permeable sand layer will markedly

increase the airflow from low permeability formations. A similar response is associated

with SVE wells intersecting naturally occurring fractures, which result in significant

increases in subsurface airflow.

In regions where naturally occurring sand layers or fractures are absent, several

methods of inducing fractures in the subsurface have been developed. Hydraulic and

pneumatic fracturing methods crack the ground by injecting a pressurized fluid, whereas

blasting creates fractures using explosives. Hydraulic fracturing is also capable of

injecting granular material into the fractures in order to keep them from closing
2

subsequent to the fracturing process. By injecting sand into the fracture, a high

permeability layer may be created in the matrix. Previous field studies have shown that a

sand filled hydraulic fracture will increase the discharge from a vapor extraction well

and the radius of pressure influence during gas pumping tests by an order of magnitude in

glacial tills [Wolf and Murdoch 1993]. However, the details associated with how to

predict pressure and flow and the vicinity of an SVE well intersecting a sand-filled

hydraulic fracture are poorly understood.

The objective of this study is to characterize the effects of sand-filled hydraulic

fractures on the flow of air to a vapor extraction well. This characterization will be done

using both field experiments and theoretical analysis to compare the performance of

conventional SVE wells to that of wells intersecting fractures. Comparing only well

discharge is of little value because the pressure on each well may differ. The

comparison will be based on the specific gas capacity of wells (a scalar measure of well

performance), and the pattern and distribution of suction and flow in the vicinity of wells.

Field experiments consist of creating sand-filled hydraulic fractures and performing gas

pump tests at two sites underlain by saprolite in western South Carolina. A theoretical

analysis is described using two models developed for this study. One model is based on

an analytical technique to solve for the pressure distributions in the vicinity of a

conventional extraction well enveloped in a low permeability well skin. This model is

inverted to estimate the vertical and horizontal permeabilities of the subsurface using

field observations of pressure in the vicinity of wells. The model is also used to estimate

skin factors, which account for pressure losses due to a low permeability sheath around

the well. The other model employs a steady state, finite difference method to solve for
3

pressure distributions in the vicinity of wells in heterogeneous, anisotropic materials.

This model considers the fracture as a thin horizontal, heterogeneity in a heterogeneous

medium in an attempt to represent field conditions more accurately than would a

homogeneous model. The ability of the numerical model to predict subsurface pressure

is tested using the field data observed during gas pumping tests.

INDUCED FRACTURES

Fractures created in the vicinity of wells have been used by the oil and gas

industry for more than 50 years to improve the yield of wells [Hubbert and Willis 1957].

In the past ten years similar technologies of induced fracturing have been applied at

environmental sites to increase the recovery of wells [Murdoch 1994]. The recovery

rates of wells completed in low permeability material, such as glacial tills or clays, are

markedly better after induced fractures have been created. However, the fracturing

process appears to increase yields in most wells where the permeability of the fracture is

significantly more than that of the formation [Murdoch 1994].

Induced fractures are created when a stress is applied that exceeds the strength of

material enveloping a well bore. Once failure of the material has initiated then the

fracture will propagate, if the driving stresses persist. The pressure of the fluid required

to drive fracture propagation is a function of the geologic characteristics of the formation

and the depth where fractures are being created [Murdoch 1994].

Pressures are measured during the fracturing process with gauges or transducers

at the top of the casing, where the fluid is being injected. The data are recorded and
4

presented in terms of driving pressure, which is the confining stress of the formation

subtracted from the injection pressure, as a function of time [Murdoch 1993]. The

pressure record is characterized by a steady increase in pressure as the starter slot inflates,

followed by a decrease in the slope of the pressure curve, indicating the onset of fracture

propagation a point known as the critical driving pressure. At this point the fracture has

initiated and propagation can be maintained at a lower pressure. Fracture propagation

continues until the injection is halted, the injection fluid leaks off into the formation, or

the fracture intersects a barrier [Murdoch 1994]. In sediment, the details of the pressure

record strongly depend on the water content of the material and the length of the starter

slot, which may be a weak point in the medium or a notch cut specifically for fracturing

applications. Murdoch [1993] demonstrated the effects of water content in laboratory

experiments by obtaining a series of pressure records for various slot lengths and water

contents in the Center Hill Clay, a colluvial silty-clay in the vicinity of Cincinnati, Ohio.

The driving pressure at the onset of propagation diminished significantly as the length of

the initial slot increases. The effect of water content on fracture propagation is related to

a term called fracture toughness, which is the magnitude of the critical stress intensity

factor, KIC. Fracture toughness is commonly regarded as a material property

characterizing the resistance of a material to mode I, or dilational, fracturing. As the

water content of a material increases, there is a decrease in fracture toughness [Murdoch

1993].

The forms of induced fractures depend mainly on site conditions and techniques

associated with the fracturing process. Conditions at the site affecting the form of

fractures are formation heterogeneities, permeability, subsurface boring, loading at the


5

ground surface, and the state of stress in the subsurface [Murdoch 1994]. The state of

stress is an important control on the dip of a hydraulic fracture, because as fractures

develop, the aperture opens in the direction of minimal compressive stress. At depths

typical of natural resource recovery (100s of meters), the greatest compressive stress is

vertical, due to the weight of the overburden [Gidley 1989]. However, as depth

decreases, the compressive stresses of the overburden are reduced, allowing for

horizontal stresses to dominate. Therefore, fractures at shallow depths tend to propagate

horizontally, with the aperture opening vertically, whereas deep fractures tend to

propagate vertically, with the aperture opening horizontally, in each case, opening in the

direction of least compressive stress [Hubbert and Willis 1957; Gidley1989]. Factors

associated with fracturing techniques consist of fluid type, rate or pressure of injection,

and borehole configuration. Induced fractures created for environmental applications are

generally at depths less than 100 meters for vadose zone or ground water contamination.

The resulting fracture form can range from a flat-lying circular disc to a steeply dipping,

elongate feature [Murdoch and Slack 2002]. The steeply dipping fractures tend to climb

and intersect the ground surface after a relatively small amount of fluid has been injected,

thus halting fracture propagation. Flat-lying fractures are considered more functional

than steeply dipping fractures because they can grow to significant sizes without climbing

and intersecting the ground surface [Murdoch 1994].

Accepted methods of creating induced fractures include hydraulic, pneumatic, and

explosive techniques. Both hydraulic and pneumatic fracturing involve the injection of

pressurized fluid to propagate cracks in the vicinity of the well. Air is typically the fluid

used for pneumatic fracturing, which has the benefit of requiring relatively simple
6

equipment compared to hydraulic fracturing [Murdoch 1994]. The injected fluid for

hydraulic fracturing is either water or gelled water. The equipment required to inject

gelled water, such as guar gum gel, is more specialized, but has the capabilities of

suspending large quantities of solids in the injected fluid [Frank and Barkley 1994].

Solids carried into the subsurface by the injection fluid are known as proppants [Murdoch

1994], and serve a variety of remediation purposes. Silica sand is often used to hold

fractures open, thereby providing a high permeability layer. Granules of reactive

compounds have also been injected to form in-situ permeable barriers in the subsurface.

Other applications include injected nutrient sources to improve bioremediation and

graphite for electro-osmosis [Murdoch and Slack 2002].

Hydraulic Fractures

The hydraulic fracturing technique for stimulating well performance was

developed by the petroleum industry in the late-1940s, and within 10 years of its

introduction, hundreds of thousands of individual treatments had been performed

[Hubbert and Willis 1957]. The typical hydraulic fracturing procedure consists of

combining materials to make a suitable fracturing fluid, which is then pumped into

borehole completed at a certain depth [Gidley 1989]. For the creation of fractures in the

study reported on herein, the locations and depths of the fractures are determined by the

placement of the injection well. The well casing and a retractable, conical tip connected

to drill rod (lance) are pushed to target depth [Murdoch 1995]. The casing and inner drill

rod are fastened together on the upper end of the assembly with a drive head. Individual

sections of casing and rod are threaded together and pushed to the desired depth. The soil
7

is exposed at the bottom of the injection well by pulling out the lance [Murdoch 1995],

which opens the wells to the formation at the desired depth (e.g., 3.0 m). A horizontal

radial notch is cut at the bottom of the casing with a specialized high-pressure jetting tool

(Fig. 1). The jetting tool consists of a 3,000 psi pressure washer, and a notching head

attached to a wand constructed of small diameter (half-inch) pipe. The notching head has

a water jet oriented normal to the axis of the well, so rotating the wand during cutting

creates a disc-shaped notch in the exposed formation [Murdoch and Slack 2002].

Direct Removal of Lance


Push
High Pressure
Cap Water

Notching
Inner Steel Wand
Rod Casing
Jetting Tool
Lance
Tip

Step 1. Step 2. Step 3.

Figure 1. Method of installing injection well and creating notch in soil. Step 1. Drive the casing and
the inner rod to depth. Step 2. Remove Rod. Step 3. Cut notch with jetting tool based on [Murdoch
1995].

An injection head is secured to the casing at the ground surface once the well is

installed and the notch is cut. The injection head includes a transducer to measure

pressure during the fracturing event. The hydraulic fractures are created by the injection

of a slurry into the well. Before the proppants are introduced into the slurry, a volume, or

pad of gel is pumped into the well to initiate a fracture and begin propagation. Guar gum
8

gel is a common fluid used to create hydraulic fractures. Guar gum, which is derived

from the guar bean, is mixed with water to form short-chain polymers, creating a fluid

similar in consistency to mineral oil, known as unlinked gel [Murdoch 1994]. A cross-

linker is added to the mixture, which causes the polymers to link and form a thick gel.

The linked gel is capable of suspending high concentrations of solid material (up to 70%)

and transport it into the fracture. The polymers are broken down by an enzyme, which is

added to the gel. The breakdown of the polymers allows the fluid to leak off into the

formation or be recovered through the injection well [Murdoch 1994].

The idealized shape of shallow hydraulic fractures has been determined from

mapping in excavations and intrepretations from borings in the vicinity of fractures

[Murdoch and Slack 2002]. The typical hydraulic fracture at shallow depths resembles

an ellipsoid, and is slightly asymmetric with respect to the injection well, with a gentle

dip, averaging about 10º [Murdoch and Slack 2002]. In some cases, the fractures are

nearly vertical, and climb rapidly to intersect the ground surface (or vent) in the vicinity

of the borehole [Murdoch and Slack 2002]. The major axis of a typical shallow hydraulic

fracture is 3 times the depth of fracture initiation, and is approximately 20% larger than

the minor axis [Murdoch and Slack 2002].

The overlying material is uplifted, or displaced during the creation of hydraulic

fractures to produce a dome. It appears that the displacement observed at the ground

surface parallels the aperture of the fracture [Murdoch and Slack 2002]. Once the

fracturing process is discontinued, the uplifted dome subsides as the fluid leaks-off into

the formation. The fracture is prevented from closing completely by proppants, which in

the case of fractures created for this study is sand, injected with the fluid. The amount of
9

closure is dependent on the ratio between the bulk volume of the sand and the total

volume of the slurry, which for the fractures used in this study is about 0.3. Therefore,

the average aperture of the sand-filled fracture is about one-third the maximum uplift

recorded during the fracturing process [Murdoch and Slack 2000].

The equipment required for hydraulic fracturing with gel consists of tanks for gel

mixing and storage, a slurry mixer, and an injection pump. The gel is hydrated in 500 gal

tanks, and is then pumped to the mixer. The slurry mixer is designed to continuously

blend unlinked gel, sand, cross-linker, and enzyme additive [Murdoch 1994]. The mixing

system consists of a sand hopper with an auger at the base to feed sand to a mixing

trough. Unlinked gel, cross-linker, and enzyme are pumped into the mixing trough,

where they are blended with the sand to form a slurry. The slurry in the mixing trough is

funneled into a positive displacement pump, which discharges through high pressure hose

to the well injection head (Fig. 2).

Positive
Screw Displacement Pump
Conveyor
Sand Hopper
To Injection
Mixer Well

P P P

Gel Storage

Breaker Linker
Storage Storage

Figure 2. Schematic of above ground hydraulic fracturing process and equipment


10

Pneumatic Fractures

The development of the pneumatic fracturing process began in 1988 by

researchers of the Hazardous Substance Management Research Center (HSMRC) at the

New Jersey Institute of Technology to enhance in-situ remediation of low permeability

materials [Schuring and Chan 1992]. Pneumatic Fracturing Extraction (PFE) is

fundamentally similar to the process of hydraulic fracturing [Schuring and Chan 1992].

Fractures are created by injecting fluid into a well, except in the case of pneumatic

fractures, the fluid is air [Murdoch 1994]. The process is carried out in open boreholes

when the formation stability allows, or in a casing with specialized screens for boreholes

requiring a completed well. The typical pneumatic fracturing method involves injecting

controlled bursts, up to 30 s in duration, of high-pressure air, typically less than 500 psig

[Schuring and Chan 1992]. Fracturing in the well bore is initiated in isolated intervals

(typically 30-45 cm) by the use of inflatable, rubber packers above and below the target

zone (Fig 3). The packers helps to focus the pressure and initiate fractures at weak

points within the packed interval [Liskowitz et al 1994].

Less is known about the forms of pneumatic fractures, due to difficulties

associated with excavation and mapping of fractures without proppants. However,

pnuematic fractures still show some evidence of horizontal propagation at shallow depths

[Liskowitz 1994]. Surface displacement recorded during fracturing events show the

development of a dome-shaped feature at the ground surface. The presence of flat-lying

beds has been shown to contribute to creation of horizontal pnuematic fractures.


11

The components of a pneumatic fracturing system consist of a pressurized air

supply, which is typically 8 to 12 cylinders of compressed air connected in series, a

pneumatic control manifold with regulators, and a downhole fracturing device [Frank

and Barkley 1994]. The downhole fracturing device has two inflatable packers, an air

delivery pipe, and a tip that evenly disperses the air throughout the isolated interval (Fig.

3).

Flow Manifold and


Regulators

Compressed Gas

Inflatable
Packers

20 -30
Fracture
Interval

Open Borehole

Figure 3. Schematic of above and below ground equipment for pnuematic fracturing.
Two inflatable packers isolate the fracture interval, which is pressurized by pnuematic
source. Idealized pnuematic fracture also shown. Figure based on [Schuring and Chan1992].
12

SOIL VAPOR EXTRACTION

Soil vapor extraction (SVE) is an in-situ process used for the removal of volatile

contaminants from the unsaturated zone [Shan et al 1992]. The use of vapor extraction

for contaminant removal has been adapted from gas extraction at landfill sites, but has

become an accepted, cost-effective technique for the removal of volatile, or semi-volatile

organic compounds (VOCs) from the vadose zone [Pederson et al 1991]. The

contaminants typically occur at or near the ground surface and are released from leaky

underground storage tanks (UST), pipelines, or accidental spills [Shan et al 1992; Fan

1987] . A significant fraction of the released contaminants can be retained in the vadose

zone due to capillary attraction, volatilization, and sorption [Hutzler 1987]. The vapors

within the interstitial spaces in the soil are expected to be in equilibrium with the liquid

phase [Pederson et al 1991]. SVE involves pumping VOCs out of the soil by applying

suction to one or more extraction wells. Convective airflow is induced by pressure

gradients created in the subsurface [Fan 1987]. As airflow occurs, the liquid phase will

evaporate to maintain the vapor-liquid equilibrium. Clean air is drawn in from the

ground surface as the contaminated air is removed (Fig. 4) [Falta et al 1993].


13

Water

Clean
Air
Contaminated
Soil Gas

Liquid
Contaminant

Figure 4. Grain-scale view of SVE process. Clean air is drawn into contaminated zone
under induced vacuum. Displaces the contaminated soil gas, causing volatilization of
contaminant trapped in water and soil pores based on[Gierke 2000].

The pattern of flow in the subsurface can have a significant effect on the ability of

SVE to remove contaminants by advection [Fan 1987]. The contaminant distribution in

the vadose zone can be quite complex, so it is often necessary to use enhanced techniques

for adjusting or controlling the flow pattern in the subsurface [Fan 1987]. These

techniques include capping the ground surface with an impermeable barrier, such as

concrete, or installing vent wells to allow the passage of air from the ground surface

[Hutzler 1987]. Idealized scenarios show how the pattern of flow is affected by such

techniques (Fig. 5).


14

a. b. c.
Impermeable
Barrier

Vent
Well

Figure 5. Idealized flow paths for conventional and enhanced SVE systems. a. flow path to an
extraction well for an unaltered system b. flow path to a well with an impermeable barrier at the
ground surface. c. flow path to an extraction well with a vent well

The performance of an SVE system depends on the physical and chemical

properties of the contaminant and on conditions at the site. Vapor removal is a function

of the volatility, soil adsorption, and initial distribution of the contaminant, as well as the

rate and pattern of airflow in the subsurface [Pederson et al 1991]. Volatility, which is

described by the vapor pressure and Henry s Law constant, is perhaps the most important

contaminant characteristic [Pederson et al 1991]. Once released into the environment,

the composition of the contaminant will change over time. The more volatile, soluble

compounds are removed initially, leaving a mixture of less-volatile, less-soluble

compounds [Pederson et al 1991]. Isolated globules of liquid contaminant may weather

and develop a resistant skin slowing the process of vapor diffusion. Enhanced SVE

approaches, such as pulsing or high suction techniques, may improve the recovery of

weathered VOCs [Pederson et al 1991].

Soil characteristics related to gas permeability, such as grain size, moisture

content, soil aggregation, and stratification have a significant effect on the airflow rate

and distribution [Hutzler 1987]. SVE is typically best suited to coarse-grained, high-
15

permeability soils, although it can be effective in fine-grained soils, particularly where

interbedded permeable layers or macropores are present [Hutzler 1987]. Gas

permeability of the soil is an important characteristic related to the performance of an

SVE system. Estimates of formation gas permeability from grain size distribution and

saturated hydraulic conductivity are subject to error in that soil moisture, gas slippage,

and swelling soils are omitted from the estimate [Pederson et al 1991]. Moisture content

can have a significant effect on airflow, because higher water content reduces the fraction

of air-filled pores, thereby decreasing the connected pores where airflow can occur

[Pederson et al 1991]. The gas permeability at a contaminated site will determine the

feasibility of remediation by SVE, so an accurate estimate of permeability is necessary

for the design of an effective system. Gas pumping tests provide a method of

determining the gas permeability by measuring the discharge from the well and

subsurface pressures in the vicinity of the well [Pederson et al 1991].

An SVE system generally consists of one or more extraction wells screened in the

vadose zone, a vacuum pump or air blower, and piping to connect the extraction wells to

the vacuum. Other above ground equipment includes gas/liquid separation devices (or a

knockout pot), flow meters, pressure gauges, and in some cases vapor treatment systems

for the removal and recovery of vapors (Fig. 6) [Gierke 2000]. Air inlet wells or vents

may also be installed to adjust airflow.


16

Vacuum
Knockout Blower
Pot Offgas
Treatment

Figure 6. Conventional SVE configuration for the removal of volatile contaminants from the vadose
zone with typical above ground based on [Gierke 2000].
17

THEORETICAL MODELS

The effects of induced fractures on the flow of air to an SVE well is evaluated, in

part, by using theoretical analysis. Two models are employed in this aspect of the study;

an analytical solution similar to that developed by [Klinkenberg 1941; Shan et al 1992]

for flow to a single extraction well, and a numerical solution for flow to a well

intersecting a fracture. Both analyses give the pressure distributions in the vicinity of the

well.

FLOW TO A WELL

The governing equation for steady-state gas flow through porous media is [Shan

et al 1992]

~
(k P 2 ) 0 (1)

~
where k = gas permeability tensor, and P = gas pressure. Darcy s Law is assumed to be

valid for gas flow, and the gas is assumed to have a constant composition and viscosity

under isothermal conditions [Shan et al 1992]). The problem will be formulated in 2-

dimensional, cylindrical coordinates with the screened interval of the well represented as

a line sink [Shan et al 1992]. The depth to the center of the screen, d, and its length , L,

are needed to describe the dimensions of the well screen (Fig. 7a).
18

Ground Surface Open


0 r to Atmosphere

d kz
z

b kr
L

Figure 7a. Well geometry for SVE models based on [Falta 1993; Shan et al 1992].

The boundary conditions of the model consist of atmospheric pressure at the

ground surface and at a radius of r = . No flow boundaries are located at depth of d = b,

and a radius of r = 0 above and below the well screen.

P Pa at 0 r ,z 0 (2)

P Pa at 0 z b, r (3)

P L
0 at 0 z d ,r 0 (4)
r 2

P L
0 at d z b, r 0 (5)
r 2

P
0 at z b (6)
z
19

Pressure at the screened interval of the well is described by

P Qs Pa L L
at d z d (7)
r 2 Lrw kr P 2 2

where Qs = volumetric flowrate at standard pressure and temperature, Pa = atmospheric

pressure, = viscosity, L = length of the screen, rw = radius of the well, and kr =

horizontal permeability.

ANALYTICAL SOLUTION

Solving the problem is facilitated by using the transform

u Pa2 P2 (8)

Substituting (8) in (1) gives

~
k u 0 (9)

and the boundary conditions are

u 0 at z 0 (10)

u 0 at 0 z b, r (11)

u L
0 at 0 z d ,r 0 (12)
r 2
20

u L
0 at d z b, r 0 (13)
r 2

u
0 at z b (14)
z

u Qs Pa L L
at d z d (15)
r Lrw kr 2 2

The solution is derived by using the equation for a potential due to a point source

in infinite media. A line source is obtained by integrating the point source over the

length of the well screen. Superposition is used to create the upper and lower boundaries

of the model.

The potential due to a point source in spherical coordinates in infinite media is

determined using the following boundary condition

P Q L L
at d z d (16)
r 4 rw2 k r 2 2

which transforms to

u Qs Pa L L
at d z d (17)
r 2 rw2 k r 2 2

This yields

Patm Qs 1
u r, z (18)
2 kr r
21

Converting (18) into cylindrical coordinates gives

Patm Qs 1
u r, z (19)
2 kr r 2
z z
2

which is the same as that developed by Shan et al [1992]. Anisotropy is included by

defining

kz
r r (20)
kr

where kz and kr are the permeability of the formation in the vertical and horizontal

direction, respectively [Shan et al 1992]. The model is axisymmetric, so the flow rate is

independent of the circumferential location. A uniform vertical line source is obtained by

integrating (19) from L/2 to L/2, which are the upper and lower limits of the well screen

L/2 1
u C1 dz (21)
2
L/2
r (z z )2

where C1 is

Patm Qs
C1 (22)
2 Lk r

The resulting solution for a continuous line sink in an infinite, homogeneous, anisotropic

medium is
22

u1 (r , z ) C1 g L, r , z (23)

where

L 2z 4r 2 (2 z L) 2
g L, r , z ln (24)
L 2z 4r 2 (2 z L) 2

which is equivalent to that developed by Shan et al [1992]. Superposition is used to form

the upper and lower boundaries. A constant pressure upper boundary and a no flow

lower boundary is created by introducing the function

f L, r , z , d , b f L, r , z d f L, r , z d

1n f L, r , z 2nb d f L, r , z 2nb d f L, r , z 2nb d f L, r , z 2nb d


n 1
(25)

where d = depth to the well, b = depth to the no flow boundary. Therefore, the potential

due to a continuous vertical line source including boundaries is

u CH , NF r , z C1 r , z, d , b, L (26)

where

r , z , d , b, L g L, r , z , d , b (27)

The pressure in the subsurface is


23

P r, z u Patm (28)

which is homologous with equation (5) in Shan et al [1992].

Pressure at the Well

The process of creating an open borehole for a well may result in a sheath, or skin,

of lower permeability material around the well screen. This skin has the potential to

affect flow to the well. The pressure drop in a well is controlled by the resistance of the

formation, the viscosity of the fluid, and the additional resistance concentrated around the

well bore resulting from drilling and completion techniques [Van Everdingen 1953]. The

pressure drop caused by this additional resistance is defined as a skin effect, which can

have considerable consequence on the productivity of a well, particularly in fine-grained

formations where smearing, or disruption of the pore structure, is may occur. The skin

effect will be included by asssuming the flow is horizontal in the vicinity of the screen,

and the pressure is governed by

2
P2 1 P2
0 (29)
r2 r r

Using

v P2 (30)

results in
24

2
v 1 v
2
0 (31)
r r r

It is required that the potential at the outside of the skin

and the flow rate to the well are known for the solution
rs
given above. The boundary conditions follow as
rw

v vs at r rs CL
Figure 7b. Schematic of
well radius and skin radius.

(32)

where rs is the radius of the skin (Fig. 7b). And following (7) and (15)

dP Qs Pa
, r = rs (33)
dr 2 Lrs k s P

which yields

dv Qs Patm
, r = rs (34)
dr Lrs k s

where ks is the permeability of the skin. The general solution of (31) is

vr C 2 C3 ln r (35)

and from (32) and (33) it is apparent that


25

Qs Patm
C3 (36)
Lk s

and

Qs Patm
C2 vs ln rs (37)
Lk r

Substituting and transforming gives the pressure at the well as

Qs Patm r
P(r ) P12 ln , r w < r < rs (38)
Lk s rs

where P1 is the pressure at the well. Solving for P1 in the absence of a skin (ks = k) gives

Qs Patm r
P12 Phw2 ln w (39)
Lk rs

where Phw is the pressure at r = rs. Substituting (39) into (38) yields

Qs Patm r Qs Patm r
P(rw ) Phw2 ln w ln w (40)
Lk rs L ks rs

rearranging

Qs Patm k r
P (rw ) Phw2 1 ln s (41)
Lk ks rw

and simplifying
26

P(rw ) Phw2 2C1 (42)

where the skin factor, , is

k rs
1 ln (43)
ks rw

and C1 is presented in (22)

Stream Function

The potential, u, and the stream function, , are conjugate harmonic functions.

Therefore the relationship between the potential and the stream function is described by

the Cauchy-Riemann equations, which can be integrated in cylindrical coordinates to

obtain the solution for the stream function [Shan et al 1992]. One of the Cauchy-

Riemann equations is

u
r (44)
z r

Integrating

u
1 r dz (45)
r

and using (24) results in


27

2 2
2 L 2 L
1 r, z C1 r z r z (46)
2 2

which is the stream function for a finite line source in an infinite media. It follows from

(25) that the stream function for a finite line source bounded above by constant pressure

and below by a no flow boundary is

NF ,CH 1 L, r , z, b, d (47)

where is defined in (25). The stream function value determined by (47) is normalized

by calculated at the constant head boundary at the ground surface. The normalized

stream function is defined as

d (48)
zd 0

where zd 0 is the stream function determined at the ground surface. The normalized

stream function varies from zero to one, and contour lines correspond to the fractional

mass flow rate through the subsurface [Shan et al 1992].

WELL INTERSECTING FRACTURE

The model for a well intersecting a fracture uses finite difference analysis coded

in Visual Basic to solve for the pressure distribution in the vicinity of a thin conductive
28

fracture. The governing equation for the problem, Laplace s equation, given in (1), is

expanded in cylindrical coordinates

v v Kr v
Kr Kz 0 (49)
r r z z r r

where

kr
Kr (50)

and

kz
Kz (51)

where Kr = radial hydraulic conductivity, Kz = vertical hydraulic conductivity, kr = radial

permeability, kz = vertical permeability, and = gas viscosity. The equation is solved

using a finite difference approximation that employs Gauss-Seidel iteration and

successive over relaxation [Wang and Anderson 1982]. In this method, the partial

differential equation is converted into a set of algebraic equations corresponding to

specific locations on a rectangular grid [Wang and Anderson 1982]. The finite difference

grid is block-centered, so the nodes are in the center of the grid blocks, and the block

boundaries one-half the distance between nodes.

The finite difference grid was designed to maximize accuracy while minimizing

execution time. This design was accomplished using closely spaced cells in the vicinity
29

of the well screen and the fracture, and more widely spaced cells elsewhere in the grid

(Fig. 8). The first ten columns closest to the well have a uniform spacing of 0.01m, and

then a geometric progression is used to increase the size of the columns, where the mesh

expands in size as we progress away from the well. The geometric progression was

accomplished by using

1 a
ri 10 r10 r1 (51)
1 a

to locate the boundary of each column, where r10 is the width of column 10, a is the

fractional increase in r over one increment, and i is a counter beginning at 1.

A row spacing of 0.01 m was used to represent the fracture, which is the thinnest

row used in the grid. The row height was increased using a geometric progression

upward to the ground surface, and another geometric progression downward to the base

of the model (Fig. 8). It was important for the grid to be readily adaptable to different

fracture depths to allow for the evaluation of the effects of fracture location. The

adaptability was accomplished by selecting the distance from the fracture to the ground,

Z, the total number of points in the grid between the fracture and the ground surface, n,

and the size of the first increment, zf, which was set to the height of the cell used to

represent the fracture. Brent s method [Press et al 1989] of finding the root of a

nonlinear equation was used to find a in

1 a
Z zf (52)
1 a

Locations of the tops of the rows overlying the fracture are obtained from
30

1 ai
zf i zf zf (53)
1 a

where zf is the location of the top of the cell representing the fracture. Locations of the

boundaries between rows under the fracture were obtained by substituting the distance

between the fracture and the lower boundary as Z in (52) and applying the same approach

outlined above.

Constant head

r
z

No flow

Figure 8. Conceptual model of numerical grid spacing showing horizontal and vertical
geometric progression of block size away from the fracture.

The model is axisymmetric, under steady-state, heterogeneous, anisotropic

conditions. The boundary conditions of the model have constant head along the ground

surface, and no flow boundaries on the other three sides. The model has two layers to

more accurately represent conditions at the site. The dimensions of the model have the

bottom of flow at d = 8 m, and the radial extent of the model at r = 20 m. The fracture is

represented as a thin conductive disc at a depth of d = 3.0 m with a radius of r = 4.2 m,

which intersects a block held at constant head to represent the screened interval of the

well (Fig. 9).


31

0 Constant head 20 m
kformationI
d=3 m
r=4.2

kformationII
kfracture fracture
aperture

8m
No flow
Figure 9. Numerical conceptual model showing boundary conditions and fracture
dimensions, with critical variables displayed.

Any point in the grid can be located by a specified integer ordered pair (i, j), with

i along the r-axis, and j along the z-axis. Heads or pressure values at point (i, j) are

calculated based on the four surrounding nodes in the grid, denoted by (i-1, j), (i+1, j), (i,

j-1), (i, j+1). In finite difference approximation, differences between nodal points

2
v
replace derivatives. A finite difference approximation of at (i, j) is obtained by
r2

taking the difference between the centered first derivatives at i 1 r and i 1 r,


2 2

which according to Huyakorn and Pinder [1983] is

v 1 vi 1, j vi , j vi , j vi 1, j
K Ki 1 ,j Ki 1 ,j (54)
r r ri 2 ri 1
2 ri 1
2 2

which simplifies to
32

v
Kr A vi 1, j vi , j B vi , j vi 1, j (55)
r r

where

2K i, j K i 1, j
A (56)
ri K i ri 1 Ki 1 ri

and

2K i, j K i 1, j
B (57)
ri K i ri 1 Ki 1 ri

An approximation in the z-direction results in

v
Kz C vi , j 1 vi , j D vi , j vi , j 1 (58)
z z

where

2K i, j K i, j 1
C (59)
zj Kj zj 1 Kj 1 zj

and

2K i, j K i, j 1
D (60)
zj Kj zj 1 Kj 1 zj

and
33

Kr h
E vi 1, j vi 1, j (61)
r r

where

2
E (62)
2 ri ri 1 ri 1
ri
K i, j Ki 1, j Ki 1, j

Substituting results in

A vi 1, j vi , j B vi , j vi 1, j C vi , j 1 vi , j D vi , j vi , j 1 E vi 1, j vi 1, j 0 (63)

Solving for the heads at (i, j) yields

A E vi 1, j B E vi 1, j C vi , j 1 D vi , j 1
vi , j (64)
A B C D

There will be one equation similar to (64) for each interior point in the grid.

Equations are solved by sweeping through columns one row at a time. Therefore

according to (64), two newly computed values are used in each iterative process. Thus

the formula for Gauss-Seidel iteration is

vim, j 1 A B C D A E vim1, j B E vim1,1j C vim, j 1 D vim, j 11 (65)

where, m represents the iteration level. The change in the value of each grid cell between

two Gauss-Seidel iterations is called the residual. If the model is converging, each
34

iteration brings the estimate closer to the exact solution, thus reducing the residual [Wang

and Anderson 1982]. This process continues until the maximum residual is minimized

below a set tolerance, which for this problem is 5.0 x 10-6.

The residual is multiplied by a relaxation factor using the method of successive

over relaxation (SOR). The SOR equation for updating vi,j at the (m+1) iteration is

vim, j 1 1 vim, j vim, j 1 (66)

where is the relaxation factor. An initial value of =1.7 was used for the analyses

presented here. The relaxation factor is decreased by 0.1 if the maximum residual begins

to increase and the model is failing to converge. The use of a relaxation factor between 1

and 2 allows for computational steps that minimize the execution time for the model to

converge, without overstepping the answer [Wang and Anderson 1982].

Stream Function

The streamlines were calculated numerically by integrating the vertical flux over

the entire ground surface to obtain the total flow rate. This value is divided by 10, with

the starting point for each streamline located where the cumulative flow rate is equal to

multiples of one-tenth of the total flow. Points are placed along the ground surface at

each 10% increment representing the initial points of the particle trace that will form the

streamlines. Points are traced by integrating the velocity field from the starting location

to the well [Pollock 1988]. Bilinear interpolation is used to determine the velocity at

arbitrary locations based on velocities at grid points. The time steps were adjusted based
35

on the velocity magnitude such that each displacement were approximately 0.05 m

[Pollock 1988].

MODEL VERIFICATION

The solution for a conventional well was verified using the gas solver code based

on the analysis by Shan et al [1992]. The numerical solution was verified by comparing

results to the analytical solution described earlier. A pressure of 0.9 atmospheres was

applied over a 0.12 m long interval centered on a point 3m deep in the numerical model.

A gas permeability of 10-11 m2 was used in the model. The numerical model was used to

predict the pressure in the formation and the discharge from the well. The lower

boundary is at 8m and the far-field boundary is 20 m away from the well. A grid of 80

by 80 cells was used. The column spacing begins with a 0.3 m radius, followed by ten

evenly spaced columns of 0.01 m. The width of the columns increased geometrically

between r = 0.13 m and r = 20 m. A height of 0.01 m was used for the row centered on

the well screen, with 40 rows above and 40 rows below the fracture. The height of the

rows increased geometrically from the center of the well screen upward to the ground

surface, and downward to the base of the model. A tolerance of 10-6 was used to identify

convergence.

The numerical model predicts a discharge of 5.88 scfm, which was used in the

analytical model to predict the distribution of pressure. The results show that the

pressures are predicted with reasonable accuracy, although slight differences between the

analytical and numerical results are apparent. The error was determined as the residual
36

between the numerical and analytical solutions normalized to the pressure difference

(Patm-Pwell = 0.1) used in the numerical model. Results were determined along three

representative vertical lines (Fig. 10). The maximum error between the models is 0.031

and it occurs at a depth of 2.9m and r = 0.065m. This is slightly above the top of the well

screen. The average error along that column is 0.008. The maximum error along the

other vertical profiles occurs at a similar depth. In general, the maximum and average

errors decrease with increasing distance from the well.

The stream function was calculated using a numerical procedure involving flow

allocation and particle tracking described above. The streamlines produced by the

process were compared to contours of the analytical solution for the stream function. The

convergence criteria was decreased to 10-8 for the streamline calculations.

The results show that the streamlines are generally similar for the two solutions

(Fig. 11). The greatest error occurs at the ground surface where the stream lines

determined analytically lie approximately 6 percent further from the well than the

numerical stream lines. The pairs of streamlines converge, however and are nearly

identical over roughly two thirds of their total length. The error near the ground surface

is related to the convergence criteria and the grid size at the ground surface. The error

can be reduced by decreasing the convergence criteria and increasing the number of

points in the grid.


37

-2.0

-2.5
r = 0.115 m

-3.0 r = 0.58 m

-3.5 r = 0.065 m

-4.0
0.94 0.95 0.96 0.97 0.98 0.99 1.00

Pressure (atm)
Figure 10. Pressure distributions predicted by the analytical and numerical models at three
radial distances (r = 0.065m, 0.115m, 0.58m)

0
Numerical
0.90.8
-2 0.70.6 0.2
0.5
0.4
0.3
-4 0.2 0.1
Analytical

-6 0.1

-8
0 2 4 6 8 10 12 14 16

Figure 11. Stream function in the vicinity of a vertical well determined by


contouring the analytical solution (thin lines) and using a numerical
procedure (thick lines).
38

FIELD EXPERIMENTS

Field experiments consist of gas pumping tests performed on conventional wells

and on wells intersecting sand-filled fractures in the vadose zone. Pumping tests

performed on conventional wells provide data for estimating formation permeabilities,

and developing a baseline for conventional SVE well performance at the field sites. Other

factors were investigated to characterize conditions at the site, including water content,

porosity, grain-size distribution, and local permebility values obtained by shelby-tube

column tests.

In order to conduct the field experiments, a total of 20 conventional wells were

installed at sites using a variety of drilling methods, and 4 sand-filled hydraulic fractures

were created. Details associated with materials required for well completions, gas

pumping tests, and the characterization of hydraulic fractures created for this study are

described in the following sections.

SITE CONDITIONS

Experiments in the field were conducted at two sites located in Pickens County,

western South Carolina (Fig. 12). One site is underlain by a granitoid gneiss, and the

other by a biotite schist. Bedrock at both sites is overlain by saprolite, which is overlain

by a massive sandy, silty, clay-rich interval. Two experimental sites were selected for

this study to evaluate the effect of differences in bedrock on the performance of SVE

wells.
39

Figure 12. South Carolina map with location of Pickens


County

Regional Geology

The field sites are located in the Piedmont physiographic province of the southern

Appalachian region. The Appalachian Highlands contain many features common to other

prominent mountain chains, such as the Alps, the Wopmay Orogen, and the North

American Cordillera [Hatcher et al 1986]. Collective features include a foreland fold and

thrust belt, a metamorphic core, cryptic sutures, a plutonic-volcanic belt, and several

large and small accreted terranes. The Appalachians result from Tertiary uplift of

dissected Mesozoic and Tertiary surfaces as erosional unloading occurred [Hatcher et al

1986].
40

The Southeastern United States have been divided into 5 physiographic provinces

to facilitate dileneation of geologic features in the orogen. The Piedmont Province is

bounded by the Coastal Plain province to the southeast and the Blue Ridge province to

the northwest (Fig. 13).

Physiographic Provinces
APPALACHIAN PLATEAUS
BLUE RIDGE
COASTAL PLAIN
PIEDMONT
VALLEY AND RIDGE

W E

200 0 200 Miles S

Figure 13. The physiographic provinces of Southeastern U.S. with the location of Pickens County.

Three major allochthonous terranes underlie western South Carolina [Nelson et al

1990]. From northwest to southeast, these terranes are the eastern portion of the Blue

Ridge, the Inner Piedmont, and the Charlotte Terrane. The field sites are located in the

Inner Piedmont, which is separated from the Charlotte Terrane to the southeast by the
41

Lowndesville Shear Zone, and the Blue Ridge to the northwest by the Brevard Fault (Fig.

14). Each terrane consists of stacks of northeast trending thrust sheets, with the Inner

Piedmont containing the Chauga-Walhalla Thrust Complex, the Six Mile Thrust Sheet,

the Paris Mountain Thrust Sheet, and the Laurens Thrust Sheet [Nelson et al 1990].

Blue Ridge
t
on
#

dm Geologysc
ie
rP n
Atlantic Coastal Plain
ne ai Blue Ridge
In rr
Te Charlotte Terrane
tte Latimer Complex
r lo Inner Piedmont
ha n
ai South Carolina
C
l Pl Geologysc
ta Atlantic Coastal Plain
o as Blue Ridge
C Charlotte Terrane
ic
nt Latimer Complex
Latimer Complex tla Inner Piedmont
A
N

W E
70 0 70 140 Miles
S

Figure 14. Allochthonous terranes associated with the Piedmont physiographic province
of western South Carolina

The Piedmont and eastern Blue Ridge Provinces are associated with the southern

portion of the internal metamorphic core of the Appalachian Highlands. Due to a lack of

fossils, extensive recrystallization, complex folding, and poor surface exposure, assigning

ages to the Inner Piedmont has been difficult [Griffin 1974]. Geologic investigations

suggest that the metamorphosed crystalline rocks underlying the region are
42

allochthonous, but the timing and mechanics associated with the placement of the

Piedmont Terrane is uncertain.

Bedrock underlying the Piedmont and Blue Ridge Provinces consists of a variety

of granites and gneisses, and foliated schists and phyllites, ranging from PreCambrian to

Mesozoic in age [Miller 1990]. Some of the metamorphic rocks were originally

sediments, resulting in clean to impure metasandstones, metagreywackes, and

metaconglomerates, whereas others were formed from intrusive or volcanic igneous

rocks. Metamorphosed volcanic tuff, ash, and flows are common locally. A majority of

the rocks have undergone several periods of metamorphism, with the heat and pressure

associated with the final period of metamorphism resulting in a mineralogy and texture

consistent with regional rocks [Miller 1990]. Igneous intrusions are common in the

Piedmont and Blue Ridge Provinces, ranging in size from small dikes and sills, to large

plutons. Large intrusions consist of granite, quartz monzonite, and gabbro, whereas dikes

and sills consist of syenite, andesite, diabase, and pegmatite [Miller 1990].

Regolith consisting of soil, alluvium, and weathered rock, overlies most of the

bedrock throughout the area. The regolith commonly consists of saprolite overlying the

bedrock. Saprolite is created by in-situ weathering, resulting in a material that retains

much of the structure of parent rock, but with minerals altered by physical and chemical

weathering [Miller 1990]. In other locations, the regolith includes material that has been

transported and deposited as colluvium or alluvium. The thickness of the regolith is

extremely variable throughout the region, ranging from 0 to more than 50 m [Miller

1990] .
43

Local Conditions

The field sites for this study are within the boundaries of the Simpson Agriculture

Experimental Station (SAES), which is located 5.0 km east of the city of Pendleton,

Pickens County, South Carolina (Fig. 15). The SAES is a series of adjacent farms and

pastureland owned by Clemson University extending approximately 8.0 km in a north-

south direction. The sites for this study are situated in the northern and southern ends of

the SAES.

The northern site is underlain by the Caesars Head Granite, whereas the southern

site underlain by an interlayered Biotite Gneiss and Biotite-Muscovite Schist [Nelson et

al 1990]. The Caesars Head Granite is a light-gray, medium-grained, discontinuously

banded to nonbanded biotite granitoid gneiss or gneissic granitoid. The composition is

mainly granodioritic, but can range from quartz monzonite and monzogranite to quartz

diorite. Discontinuous black, biotite-rich bands that have been folded are common.

Kaolin and biotite-rich bands were observed at the north field site to have a dip less than

10 degrees. Preliminary Pb207/ Pb206age of zircons is estimated at about 435 Ma [Nelson

et al 1990].

The interlayered Biotite Gneiss and Biotite-Muscovite Schist unit consists of

undivided biotite-plagioclase-quartz gneiss and biotite schist. The unit also includes

subordinate layers of megacrystic biotite gneiss, sillimanite-mica schist, and minor

amphibolites [Nelson et al 1990]. Schistosity at the south site appears to be relatively

flat-lying, with a maximum dip less than 10 degrees, based on measurements taken from

core samples.
44

North
Site

Bishop Branch Rd

Lebanon Rd South
Site

Figure 15. Map to North and South field sites

Both field sites are currently or have previously been used for agricultural

purposes, with the resulting soil horizon properties being indicative of the land use (Fig.

16) [Faure 1998]. There are two major layers observed at the sites, which are an

overlying relatively structureless interval, and an underlying saprolitic interval, with the

structureless interval subdivided into five separate horizons. At the north site, the

uppermost soil horizon (Ap) is approximately 0.15 m thick. This horizon is brown, and

rich in organic matter and humus with many roots present. All of the rock structure has

been obliterated in the surficial horizon due to cultivation processes. Below the
45

uppermost soil horizon, lies a 0.15 m thick interval (E). This interval has less organic

matter than the Ap horizon, but still contains roots and evidence of biological activity,

such as burrows. This interval is dominated by sand and silt-sized particles. The color of

this interval is slightly more red than is the Ap horizon, but it is lighter in color than is the

underlying region. Beneath the E horizon is a 0.15 m thick transitional region (BE),

which grades into an underlying horizon (Bt) approximately 1.2 m thick. The Bt horizon

is reddish, and darker in color than the overlying transition zone, and contain variable

amounts of sand, silt, and clay-sized particles. Roots are present in the Bt horizon, as

well as relic pegmatitic material, such as massive quartz, coarse-grained micas, and

kaolin blocks and bands. This horizon is underlain by another transition zone (BC),

which is approximately 0.6 m thick. This transition zone exhibits properties similar to

the Bt horizon, with an increase in iron oxides, such as hematite and goethite. This lower

boundary of the transition zone marks the contact between the relatively structureless,

clay-rich horizon, and the underlying saprolitic material. Below this horizon, lies a

saprolitic region (C), where remnant rock structures occur, and the color lightens to a

reddish-tan. Bands of quartz, kaolin, and mica are present. Weathered ferromagnesium

minerals and manganese oxides are also present as black bands and isolated granules.

At the south site, the uppermost soil horizon is 0.30 m thick, which, similar to the

north site, is brown in color and rich in organic matter with many roots present (Ap).

Below the Ap horizon lies an interval 0.30 m thick (E) that has less organic matter and is

more red than the Ap horizon, but lighter in color than the underlying horizon. The E

horizon is dominated by sand and silt particles. A 0.15 m thick transitional horizon (BE)

lies below E horizon. Beneath the BE horizon is an interval 1.0 m thick (Bt), which is
46

reddish, with variable amounts of sand, silt, and clay, and the presence of roots. This

horizon is underlain by another transition zone, which is 0.6 m thick (BC), exhibiting the

first down-section appearance of relic structures, such as coarse-grained micas and bands

of kaolin. Saprolitic material (C) occurs below this horizon, which is tan and contains

more pods of micas and kaolin relic pegmatites than the overlying material.

The soil profile at the two sites is similar. It consists of an upper relatively

structureless region extending from the ground surface to a depth of approximately 2.3 m.

This region is underlain by saprolitic material, which exhibits relic structures associated

with the parent material. The contact between these two layers varies between 2.2 and

2.6 m below ground surface at the north site, and between 2.3 and 2.8 m at the south site.

The uppermost soil horizon is classified as an Ap soil horizon, which is defined as

a mineral horizon at the ground surface with properties resulting from cultivation, or

similar kind of disturbance [Faure 1998]. Below the Ap horizon lies an E horizon. The

main feature of an E horizon is the loss of silicate clay, iron, aluminum, or some

combination of these minerals. The transition zone underlying the E horizon is denoted

as a BE. This is a layer dominated by the properties of a B horizon, but exhibiting

subordinate properties of an E horizon. A B horizon is dominated by the obliteration of

all or much of the parent rock structure, and the accumulation of silicate clay, iron,

aluminum, humus, and/or silica. The formation of granular, blocky, or prismatic clay

structures are also indications of a B horizon. Following the BE horizon is Bt horizon,

which is characterized by the properties of a B horizon and the accumulation of silicate

clay in particular. Clay coatings on the surface of peds, in pores, or as bridges between

mineral grains are evidence of a Bt horizon. Beneath the Bt horizon is the second
47

transition zone classified as a BC horizon. Again exhibiting properties of both horizons,

but dominated by characteristics of a B horizon. The underlying C horizon is defined as

a horizon, excluding bedrock, that is little affected by pedogenic processes, and lack

properties of an O, A, E, or B horizon. Included as C layers are sediment, saprolite, and

unconsolidated bedrock. At some depth below the C horizon, lies hard bedrock, which is

classified as an R horizon.

Ap Brown, humus-rich, many roots 0


E present. Rock structure obliterated.
m
BE
Reddish color with roots present.
Sand, silt, and clay content variable.
Bt Coarse-grained micas, quartz, and
1m
kaolin blocks and bands present

depth
Reddish color with few roots
present. Sand, silt and clay variable.
BC Coarse-grained micas and kaolin
present. Iron oxides and massive 2m
quartz present at north site.
Banding visible at both sites.
C Reddish-tan color at north site, tan
color at south site. Weathered
ferromagnesiums and manganese
a. North Site oxides at the north site.
3m
b. South Site

Figure 16. Soil horizons and description for a. North site b. South site

Soil samples were taken at 0.3 m intervals to a depth of 3.0 m with a split-spoon

sampler. Estimates of bulk density, porosity, and degree of saturation were obtained

using a gravimetric method. Sample volume and weight were recorded in the field. The

samples were oven-dried at a temperature of 100 to 110ºC for 24 hours to remove

gravitational, capillary, and hydroscopic water. The water content by weight was

calculated using the wet and dry weights of the samples. Particle density was obtained

from the literature, which was estimated to be 2.55 g/cm3 based on the particle densities
48

of quartz and clay [Faure 1998]. Bulk density, porosity, and degree of saturation were

calculated from the known parameters.

The soil at the north site has an average bulk density of 1.31 g/cm3, ranging from

1.60 g/cm3 at a depth of 0.3 m to 1.10 g/cm3 at a depth of 2.7 m. Porosity ranges from

0.57 at 2.7 m to 0.37 at 0.3 m, with an average porosity of 0.49. The average degree of

saturation has a value of 0.73, ranging from 0.84 at a depth of 1.5 m to 0.63 at 2.7m

(Table 1).

Bulk density at the south site ranges from 1.11 g/cm3 at a depth of 0.6 m to 1.59

g/cm3 at 1.5 m, with an average value of 1.38 g/cm3. Porosity averages 0.46, ranging

from 0.40 at depths of 1.0 m and 2.0 m to 0.56 at a depth of 0.6 m. Degree of saturation

at the south site ranges from 0.41 at 0.6 m to 0.99 at a depth of 2.0 m, with an average

value of 0.80 (Table 1).

Both soils have variable bulk densities, porosities, and degrees of saturation that

indicate a locally heterogeneous subsurface. However, there appears to be a pattern in

these soil characteristics that support the concept of a subsurface dominated by two

layers. The soil descriptions of the site indicate a lithologic change occurring at a depth

around 2.0 to 2.5 m, with the overlying material being relatively structureless, and clay-

rich compared to the underlying saprolitic material. The descriptions are confirmed by

the degrees of saturation, which drop from 0.82 at a depth of 2.4 m to 0.63 at a depth of

2.7 m at the north site, and from 0.98 to 0.70 within the same interval at the south site.

Bulk densities and porosity also reflect this transition with lower bulk densities and

higher porosities associated these depths.


49

Table 1. Water content, porosity and associated data for a. North Site b. South Site

a. North Site

Depth Vol. Mass P.D. Water B.D. Porosity Deg.


(m) (cm3) (g) (g/cm3) (g/cm3)
Content Sat.
0.3 26.0 39.0 2.55 0.29 1.50 0.41 0.70
0.6 16.0 17.8 2.55 0.23 1.11 0.56 0.41
0.9 40.5 61.9 2.55 0.31 1.53 0.40 0.77
1.2 40.5 57.4 2.55 0.31 1.42 0.44 0.70
1.5 40.5 59.3 2.55 0.39 1.46 0.43 0.92
1.8 40.5 57.2 2.55 0.44 1.41 0.45 0.99
2.1 40.5 60.3 2.55 0.40 1.49 0.42 0.96
2.4 40.5 53.8 2.55 0.47 1.33 0.48 0.98
2.7 40.5 53.3 2.55 0.34 1.32 0.48 0.70
3.0 40.5 49.2 2.55 0.43 1.21 0.52 0.82
average 1.38 0.46 0.80

b. South Site

Depth Vol. Mass P.D. Water B.D. Porosity Deg. Sat


(m) (cm3) (g) (g/cm3) Content (g/cm3)
0.3 10 16.0 2.55 0.29 1.60 0.37 0.78
0.6 21 31.1 2.55 0.34 1.48 0.42 0.81
0.9 19 25.8 2.55 0.32 1.36 0.47 0.68
1.2 24 28.6 2.55 0.36 1.19 0.53 0.68
1.5 22 28.1 2.55 0.42 1.28 0.50 0.84
1.8 21 26.8 2.55 0.39 1.28 0.50 0.78
2.1 19 25.9 2.55 0.27 1.36 0.47 0.58
2.4 22 28.1 2.55 0.41 1.28 0.50 0.82
2.7 22 24.2 2.55 0.36 1.10 0.57 0.63
3.0 17 19.5 2.55 0.36 1.15 0.55 0.65
average 1.31 0.49 0.73
50

Soil moisture tension was determined at the north site using Soil Moisture

tensiometers with a 1.0 bar porous ceramic cup. Tensiometers were installed in a circular

pattern with a 0.3 m horizontal spacing approximately 2.5 m northwest of the

conventional well cluster. The vertical spacing is 0.3 m, ranging from a depth of 0.3 m to

1.5 m below ground surface. Gauge pressures were recorded at approximately one week

intervals over the course of a month from late-September to late-October of 2001. In

general, the soil moisture tension decreases with depth, with the soil closest to the ground

surface having the greatest moisture tension (Table 2). The soil at a depth of 0.9 m

appears to have a moisture tension that is consistently lower than overlying or underlying

soils.

Table 2. Pressure heads (cm of water) determined from tensiometers at the North Site over
1 month.

Depth (m) 09/25/2001 09/27/2001 10/04/2001 10/08/2001 10/22/2001


0.3 180 430 670 560 580
0.6 160 340 500 510 520
0.9 80 210 310 310 390
1.2 220 280 360 370 470
1.5 80 220 280 270 370

Grain-size analyses were performed using soils from both sites. The samples

taken for the evaluation of water content were combined to form two groups for each site

based on the properties of the soil horizons, with the upper layer ranging from 0.90 m to

2.15 m, and the lower layer ranging from 2.15 m to 3.00 m. Samples were sieved with a

300 mesh sieve to separate sands and finer materials. Percentages of silt and clay were
51

determined using a water suspension method (Table 3). Extraction wells and hydraulic

fractures were installed at a depth 3.0 m, placing them in the lower region, which for the

north site was classified as a Loam, and for the south site a Sandy Clay Loam (Table 3).

Clay contents of the upper layer compared to the lower layer for both sites confirm the

interpretation of a clay-rich overlying region, and a saprolitic underlying region.

Table 3. Results from sieve analysis for field sites

Location Depth % Sand % Silt % Clay texture


North Site 0.90-2.15 m 46.5 22.9 30.6 Sandy Clay Loam
North Site 2.15-3.00 m 47.2 32.6 20.2 Loam
South Site 0.90-2.15 m 44.4 19.6 36.0 Clay Loam
South Site 2.15-3.00 m 64.9 14.7 20.4 Sandy Clay Loam

EXPERIMENTAL METHODS

The effects of sand-filled hydraulic fractures on the flow of air to SVE wells was

determined, in part, by the performance of wells during gas pumping tests conducted at

the field sites. Conventional, gravel-packed wells were installed to function as extraction

wells and/or piezometers for a standard SVE system. Sand-filled hydraulic fractures

were created at both the north and south sites, with the injection well used to create the

hydraulic fractures serving as the vapor extraction well. Specialized multi-level

piezometers were the primary source of pressure data during gas pump tests performed

on wells intersecting fractures.

Gas pumping tests consist of measuring the difference between the ambient

atmospheric pressure and the air pressure in the soil during vapor extraction. Gas is
52

pumped from a well that had been screened in the unsaturated zone, which reduces the

pressure in the subsurface [Fan 1987]. The pressure is measured by piezometers installed

at various vertical horizontal locations, and is used along with the pressure and the

discharge rate at the extraction well to determine the effects of the fracture on subsurface

airflow [Pederson et al 1991]. Pressures were measured at the well head through a port

on the vapor extraction manifold. The data were used to determine the specific gas

capacity of the wells, and obtain estimates for the gas permeability of the formation.

Measurements of pressure at the well and in the subsurface along with the discharge rate

also allowed the skin factors associated with the extraction wells to be estimated [Van

Everdingen 1953].

It was necessary to determine pressure losses due to the length of the well casing

to ensure that we have an accurate estimate of pressure at the well screen. Losses were

determined by taking simultaneous measurements during gas pump tests of the pressure

at the ground surface, and at the bottom of the well. Down hole pressures were obtained

through a capillary tube lowered inside the casing. The flow rate was reduced by 25%

increments, and the pressures observed at the ground surface and at a depth of 3.0 m were

recorded. Pressure losses due to the length of a 0.025 m diameter PVC well casing were

determined to have a maximum of 3.3 in of water associated with the conventional

extraction well with the greatest discharge (Well SW-3, Table 4).
53

Table 4. Pressures recorded during gas pump tests to determine pressure losses due to
casing length at a. North Site, NW-3 b. South Site, SW-3

a. NW-3

Discharge (scfm) Pressure (in. H20) Pressure (in. H20) Pressure difference
@ surface @ 3.0 m (in. H20)
14.50 30.0 26.7 3.3
11.50 21.9 20.2 1.7
7.50 13.6 12.1 1.5
4.00 8.1 6.7 1.4

b. SW-3

Discharge (scfm) Pressure (in. H20) Pressure (in. H20) Pressure difference
@ surface @ 3.0 m (in. H20)
6.90 44.3 42.2 2.1
5.20 30.7 29.2 1.5
3.50 19.0 17.9 1.1
1.75 8.4 7.5 0.9

Local gas permeability measurements and sand permeabilities were obtained in

the field using a shelby-tube column tests. Soil samples were taken using shelby-tubes in

0.6 m increments to a depth of 3.0 m. Shelby-tubes with samples were attached to an air

injection head fitted with a pressure transducer. One end of the sample tube was left

open to the atmosphere, while the other was sealed by the air injection head. Airflow was

induced by a 1hp blower, which was connected to a flow meter, ranging from 0-60 scfh,

to measure the discharge (Fig. 17). Air was injected into the shelby-tube at various flow

rates and the corresponding pressure was used to calculate gas permeability using
54

Q L
k (67)
A P

where Q = flow rate injected into the sample, = gas viscosity, L = length of the sample,

A = cross-sectional area of the sample, and P = observed pressure difference.

A similar process was used to determine the permeability of the sands used to fill

the hydraulic fractures. The shelby-tubes were manually packed with sand at varying

water contents to determine the effect of moisture on the gas permeability of the sands

filling the fracture.

Air Source
(Blower)
Pressure
Flow Meter Transducer

Sealing
Cap

Shelby Tube
(with sample)

Open to
Atmosphere

Figure 17. Schematic of shelby tube permeameter. Air is injected from


source through flow meter into the shelby tube. Corresponding pressure
difference compared to atmospheric is used to calculate a permeability.
55

Experimental Sites

Each experimental site consists of a cluster of conventional wells used to

determine achievable SVE performance in saprolite, and two extraction wells intersecting

sand-filled hydraulic fractures. The hydraulic fractures are characterized by injection

pressure and delivery rate, and the resulting fracture form.

North Site

The north site is located near the Bull Testing Center of the SAES, in a pasture

adjacent to an equipment storage shed (Fig.18). This land is used for cultivation of

livestock feed, which is either harvested or left for grazing, depending on the time of

year. The north site includes a cluster of ten conventional wells within a 2.0 m2 area

located roughly 8.0 m west of the center of the storage shed located on the SAES. The

well cluster consists of relatively closely spaced (approximately 0.3 m apart)

conventional wells installed using a variety of drilling methods. There is one well

intersecting a hydraulic fracture filled with medium-grained sand approximately 12.0 m

to the north of the conventional well cluster, and a well intersecting a hydraulic fracture

filled with coarse-grained sand approximately 15.0 m to the southwest of the

conventional well cluster. Multi-level piezometers were installed in the vicinity of the

conventional well cluster and wells intersecting fractures.

Displacement caused by fracturing was measured at the ground surface during the

fracture events using simple surveying techniques [Murdoch and Slack 2002]. An array

of vertical stakes with visible metric scales was centered at the injection well extending
56

4.5 m in six radial directions. Elevations of the uplift stakes were read along a horizontal

plain using a surveying transit positioned approximately 15.0 m from the array. Initial

elevation measurements were taken prior to the fracture event, during fracture

propagation, and immediately following completion of the process. The pattern of uplift

is used to estimate the form and the distribution of sand in a shallow hydraulic fracture

[Murdoch and Slack 2000]. Plots of maximum uplift for the fractures created in this

study are presented in the following sections.

The hydraulic fractures created at the north site for this study were initiated at a

depth of 3.0 m. The wells were installed by pushing 5.0 cm (2 in) diameter steel casing

and a retractable inner rod and point assembly to the desired depth (Fig. 1). The notch

was cut using high-pressure water by the process outlined in the introductory section of

this report. Sediment laden water was captured as it flowed out of the top of the casing in

order to estimate the amount of soil being removed by the water. Three 5-gallon buckets

were filled with fluid for each notch. One fracture is propped with a medium sand (16-40

slot), and the other fracture is propped with a coarse sand (6-20 slot).

The medium-sand fracture was created with 450 kg of sand, and 660 L of guar

gum gel. A 75 L pad of unlinked gel was injected to initiate fracture propagation prior to

introduction of the sand into the slurry, resulting in a ratio of bulk sand volume to total

slurry 0.3. The critical driving pressure, or the pressure at which fracture propagation

started, is 33 psi, and the fracture propagated at approximately 24 psi (Fig. 19).

The critical driving pressure of the coarse-sand fracture is 32 psi, with pressure

decreasing gradually to 22 psi before the fracturing process was discontinued (Fig. 19).

A pad of 80 L of gel was used in creation of the coarse-sand fracture, but injection was
57

halted after only 185 kg of sand were injected because a vent developed approximately

2.0 m to the northwest of the injection point (Fig. 20).


58

Medium Sand Fracture

Storage Shed
Tensiometer Cluster

Conventional Well
Cluster

Coarse Sand Fracture

10 m
N
Explanation
Conventional Well
Injection Well
Multi-Level Piezometer

Figure 18. North Site Map


59

a. medium-sand fracture
45 12
40
10
35

Sand Volume (ft3)


Pressure (psi)

30 8
25
6
20
15 4
10
2
5
0 0
0 5 10 15 20 25 pressure
Time (min) sand

b. coarse-sand fracture
35 4

30
3
25 Sand Volume (ft3)
Pressure (psi)

20
2
15

10
1
5

0 0 pressure
0 2 4 6 8 10 12
Time (min) sand

Figure 19. Injection pressure and sand volume as a function of time recorded during
fracturing event at the North Site for a. medium-grained sand fracture b. coarse-grained sand
fracture.
60

0.1
a.
0.2
0.0 0.0
0.5
0.5
-0.1 0.6

0.4
0.1 0.4
0.3

0.0 0.1

0.1

Injection point

4.0
3.0
5.8
5.0 b.
2.0
6.0

0.0 6.2

5.6
vent

1.5
2.4
5.5 N

5.1
0.5
1.0
ND
0.0 5.3

1m
0.1

Figure 20. Net surface displacement (cm) recorded from the fracture events
at the north site for a. medium sand fracture b. coarse sand fracture
61

South Site

The south site is located near the southern end of the SAES. Enter the SAES, go

past the first shed, and turn right between the buildings. Continue past the fire ant

research facility until paths merge between the pesticide research fruit grove and a

cultivation plot. The field site is located in a clearing past a telephone pole on the right.

The field site consists of a cluster of conventional wells, a well intersecting a medium-

grained sand fracture, and a well intersecting a coarse-grained sand fracture (Fig. 23).

Multi-level and conventional piezometers were installed in the vicinity of the hydraulic

fractures. The well cluster consists of seven conventional wells, and a multi-level

piezometer installed within a 3.0 m2 area. The coarse-grained sand fracture is

approximately 10.0 m southeast of the conventional well cluster, and the medium-sand

fracture is 15.0 m to the east-northeast of the well cluster (Fig. 21).

The medium-grained sand fracture was installed using 660 L of guar gum gel,

injecting 450 kg of sand. A pad of 75 L of gel was needed before the sand could be

introduced into the slurry, creating a ratio of bulk sand volume to total slurry volume of

0.3. The critical driving pressure was 38 psi, with pressure maintaining at approximately

24 psi during the propagation of the fracture (Fig. 22).

The coarse-grained sand fracture required a larger pad, of 115 L, to initiate

fracture propagation. Gel requirements for this fracture exceeded other fractures installed

for this study, totaling 830 L. 450 kg of sand were injected in the process, resulting in a

ratio of bulk sand volume to total slurry of 0.25. The critical driving pressure for the
62

coarse-grained sand fracture was 35 psi, with pressures fluctuating between 25 and 34 psi

during fracture propagation (Fig. 22).


63

Agricultural Plot

Coarse Sand Fracture

Overgrown Region
Well Cluster

Telephone Pole

Medium Sand Fracture

10 m

Explanation
N Conventional Well
Injection Well
Multi-Level Piezometer

Figure 21. Site Map for South Site.


64

a. medium-sand fracture
40 12
35
10
30

Sand Volume (ft3)


Pressure (psi)

8
25
20 6
15
4
10
2
5
0 0
0 5 10 15 20
pressure
Time (min) sand

b. coarse-sand fracture
40 12
35
10
Sand Volume (ft3)

30
Pressure (psi)

8
25
20 6
15
4
10
2
5
0 0
0 5 10 15 20 25 30 35 pressure
Time (min) sand

Figure 22. Injection pressure and sand volume as a function of time recorded during fracturing
process at the South Site for a. medium-sand fracture b. coarse-sand fracture.
65

0.2 a.
0.6
0.5

1.0
0.3 0.7
1.5
1.5
1.0
1.3
1.6

1.1 1.1
ND

0.2 0.5

0.3

Injection Well

0.0

0.2
0.5
0.3 0.0 b.
0.8
0.9
0.6
1.0
1.1 N
1.2
0.6
0.9

0.5 0.1

0.2 1m

Figure 23. Net surface displacement (cm) recorded from the fracture events
at the south site for a. medium sand fracture b. coarse sand fracture
66

Experimental Materials

The experimental sites include clusters of conventional wells, representing a

conventional SVE system, and two extraction wells intersecting sand-filled hydraulic

fractures. Piezometers were installed in the vicinity of the fractures using a conventional

well completion, which consists of a well screen or gravel pack, and a specialized

inflatable multi-level piezometer.

Soil Vapor Extraction Assembly

Airflow was induced in the subsurface using a 1hp 50Hz rotary vane Rotron

blower operating at 110 V. A 5.5 hp generator provided power for tests at the south site,

which is remote from a power source. The vapor extraction manifold consists of an

Omega flow meter, either ranging from 0-40 scfm or 0-60 scfh depending on the flow

rate, and a ball valve to control the discharge (Fig. 24). The manifold, which includes the

flow meter, is constructed of 0.025 cm (1.0 in.) diameter schedule-40 PVC, and is used to

connect the air inlet of the blower to the extraction well. PVC reducing bushings were

required to neck down the 0.038 m (1.5 in) diameter threads on the blower to the

diameter of the piping. Dwyer Mark III digital manometers were used for pressure

measurements in the field. Two manometers were used for this study, one with a range

of 0-199.9 inches of water (0-49.7 kPa) for measurements at the extraction well, and the

other with a range of 0-19.99 inches of water (0-4.97 kPa) for more refined low suction

measurements.
67

Blower
Flow Meter

Transducer
Port Air
flow
Attached to
Well Head

Figure 24. SVE setup for this study including 1hp blower, 1 in. PVC piping, transducer port,
and interchangable flow meter.

Wells and Piezometers

Two methods of completion were used for wells in this study: a conventional

technique for creating single-screen wells; and a technique for creating multi-level

piezometers. A generalized conventional well completion involves creating an open

borehole and installing casing with a well screen. In the case of conventional wells

installed for this study, the casing is open at the bottom and is enveloped by pea gravel to

function as a well screen. The screened section of these wells was omitted because of the

short interval used in the installations. A casing is lowered into the borehole pre-filled

with 0.15 m of gravel, which is then followed by another 0.15 m of gravel after the casing

is inserted. The gravel is overlain by 0.15 m of a mixture of one-half crushed bentonite

and one-half bentonite pellets, and then cement to the ground surface (Fig. 25). The

bentonite is intermittently hydrated with a few ounces of water by trickling water from a

bottle to create a seal separating the gravel pack from the cement. It is important to keep

water from entering the gravel pack because the relative-gas permeability of the
68

surrounding formation may be significantly reduced. A sand-mix concrete was used to

seal the annulus above the bentonite with the addition of Tetraguard AS-20 at a

concentration of 1% by volume. Tetraguard AS-20 is a water reducing agent to

minimize shrinking and cracking of the concrete.

Three different drilling methods were used to create the boreholes for

conventional wells: hand augers; machine augers; and shelby-tubes. Slight variations in

borehole geometries resulted from the drilling methods. A machine augered borehole has

a diameter of 10.2 cm (4 in), a shelby-tube borehole has a diameter of 7.6 cm (3 in), and a

hand augered borehole has a diameter of 5.7 cm (2.25 in). For all wells, the borehole was

brushed before completion with one pass of a single spiral nylon brush corresponding to

the size of the borehole to reduce the effects of smearing.

PVC
Casing

Grout
2.70 m
formation

Bentonite
0.15 m

0.30 m
Gravel Pack

Figure 25. Conventional well completions.


69

Conventional screened wells were used both as extraction wells and piezometers

for the conventional SVE systems. The vertical resolution of the conventional wells is

approximately 0.3 m, which is coarser than needed to characterize the pattern of flow in

the vicinity of hydraulic fractures.

Pressure distributions profiles were obtained in the field using a specialized multi-

level piezometer [Jones et al 1999]. The piezometer consists of plastic sheath with vinyl

tubing extending inside the sheath from the ground surface down to individual ports. The

ports allow for the passage of air through the sheath to a filter pad (Fig. 26).

The plastic sheath is 7.6 cm (3.0 in.) lay flat polyethylene, with 0.1 mm (0.004

in.) wall thickness. The vinyl tubing has a 3.2 mm (1/8 in.) outer diameter, and a 1.6 mm

(1/16 in.) inner diameter. The ports connecting the vinyl tubing to the filter pads are

Clippard Minimatic brass fittings, and the filters are 6.5 cm2 pieces of Scotchbrite

abrasive pad. The filter pads are attached to the brass fittings by 10/32 hexagonal nuts

glued to the inside of the pad with a fast drying epoxy. Plastic washers are used on the

inside and outside of the plastic sheath where the brass fittings connect to the filter pads

to ensure a seal is maintained as fittings pass through the plastic sheath (Fig. 26). The

borehole is created by split-spooning every 0.6 m interval down to 0.5 m past the depth of

the lowest port on the multi-level piezometer. The extra 0.5 m of borehole allows room

for the extra 0.5m of plastic sheath that contain a 0.3 m-long sand-filled weight. Split-

spooning results in a 4.8 cm (1.9 in) diameter borehole, which is brushed with a 5.7 cm

(2.25 in) single spiral nylon brush. The sheath assembly is lowered into the open

borehole using the weight of the anchor. The sheath is inflated with air using the outlet

of the vapor extraction blower, or with a shop vacuum, to remove folds and creases in the
70

plastic. The sheath is then filled with dry sand using a hand-held funnel and scoop. This

forces the filters against the borehole wall, and creates a seal around the pad. The volume

of sand required to fill the piezometer was estimated based on the radius and height of the

borehole. A piezometer that spans 4.6 m requires a bulk volume of approximately 2.5 gal

of sand. Bridging may occur during the process of filling the sheath with sand, resulting

in observed sand volumes required to fill the piezometer that are less than estimated

based on the piezometer dimensions. Occasional blasts of air from the blower or shop

vacuum appear to solve the problem.

The borehole is brushed in an attempt to minimize the effects of smearing caused

by the split-spooning process. Later analysis has revealed the possible presence of an

annulus between the sheath of the multi-level piezometer and the formation. This could

be a result of the brushing or split-spooning process, which could disturb the wall of the

borehole.

Once installed, the multi-level piezometers are protected at the ground surface by

a vault. An initial hole with a diameter of 0.3 m was created with a depth of 0.2 m before

piezometer installation. This provides space for a vault to be created around the

piezometer. The vault consists of a 0.25 m long piece of 10.16 cm (4 in) diameter pipe

and cap. The pipe is driven with a hammer roughly 0.1 m into ground surrounding the

exposed plastic sheath and vinyl tubes. The pipe is then capped and the open hole

surrounding the pipe is filled with cement. Access to the piezometer is gained through

the cap.
71

Tubing To
Ground Surface

Sand

Filter Pad
Formation

Brass
Fitting

Plastic
Washers
Plastic
Sheath

Figure 26. Close-up of sand filled multi-level piezometer. Shows filter pad
against formation, with connections to vinyl tubing leading to ground surface.
72

RESULTS

The solutions for airflow to a conventional well with a skin and to a well

intersecting a fracture determine the pressure distributions in the subsurface based on the

horizontal and vertical permeabilities and the discharge rate from the extraction well.

Approximations of the radial permeability, anisotropy ratio kv/kh, and skin factor can be

obtained from the observed pressure distribution and discharge by using the analytical

solution in a parameter estimation routine.

The numerical model was validated by comparing field observations to the

predicted data. Actual permeability ratios were calculated using the estimated

permeabilities of the sand in the fractures and the soil at the site. These estimates were

obtained from the analysis of gas pumping tests and the shelby-tube permeameter. The

observed specific capacities of wells intersecting sand-filled hydraulic fractures were

compared to specific capacities predicted by the numerical model, which was calibrated

using field permeability estimates to more accurately represent site conditions.

THEORETICAL ANALYSIS

The models provide the best available method for extrapolating information from

one experimental site to another. The effects of variables associated with the process

may also be studied without incurring the cost of setting up field-scale systems. In this

section of the report, well performance is assessed using the flow efficiency, which is the
73

ratio of the observed specific capacities to the optimum specific capacities of

conventional wells at the field sites, and the pressure and flow in the vicinity of wells.

The effect of well geometry and anisotropy ratio are determined by investigating the flow

efficiency for a variety of conditions.

Critical variables associated with the models are the permeabilities of the fracture

and the formation, and the geometry of the fracture. The ratio between the permeabilities

of the fracture and the formation is a key factor in determining the effects of the fracture

on well performance. Separate analysis based on fracture propagation mechanics show

that the volume of injected sand is perhaps the single most important factor related to

fracture geometry, in that the sand volume determines the aperture and radial extent of

the fracture [Murdoch 2002]. The effects of the fractures on subsurface pressure and

flow are evaluated with dimensionless plots of the pressure distribution and stream

function for varying permeability ratios and fracture geometries.

Specific Capacity

One aspect of the well performance evaluation is based on specific gas capacity,

Sc, which is a scalar measure of well performance, where

Qs
Sc 2 2
(68)
Patm Pwell

The specific gas capacity of an SVE well follows from the definitions of related

terms, specific capacity of a water well and the productivity index of an oil well. The

specific gas capacity differs slightly from those terms in that it involves a difference in
74

squared pressures instead of simply a difference in pressures or hydraulic heads. This

difference occurs because the compressibility of gas is included in the analysis of an SVE

well, whereas compressibility is ignored when analyzing the performance of wells

recovering water or oil [Earlougher 1977].

Specific gas capacity depends on the geometry of the well and gas phase

permeability of the formation. The specific capacity is expected to be constant over a

moderate range of flowrates and pressures at the well, which is consistent with field

observations, where the specific capacity is a linear function of the pressure observed at

the well (Fig. 27). Presumably Sc changes at high suctions when the Klinkenberg effect

becomes important [Klinkenberg 1941], but at values of suction typical of most field

implementations of vapor extraction, Sc is expected to be constant. Accordingly, it is

possible to predict the discharge as a function of the applied suction from the well whose

specific capacity is known (Fig. 27).

16 y = 116.91x
14 R 2 = 0.9953

12
Discharge (scfm)

10
y = 36.462x North Site
8
R2 = 0.9858 South Site
6

4
2
0
0 0.05 0.1 0.15 0.2 0.25

Patm2-P- obs 2
Pwell2

Figure 27. Observed pressures at varying discharges for an individual well at the north
and south sites. Slope represents specific capacity the well.
75

It is convenient to evaluate performance of a well by scaling its specific gas

capacity to that of a control well under ideal conditions, Sco. The ratio

Sc
F (69)
S co

is referred to as flow efficiency as it is similar to a term used in the petroleum industry

[Earlougher 1977; Streltsova 1988].

Conventional Wells

The flow efficiency of a well may be affected by changes in permeability

compared to an idealized control well. One important example of this permeability effect

is a low permeability skin that may occur in the vicinity of the well bore. A low

permeability skin will cause the flow efficiency to be less than 1.0, indicating that the

observed specific capacity is less than that of a control well.

The flow efficiency of a well affected by a skin is related to the skin factor

according to

rw , d , d , b, L
F (70)
rw , d , d , b, L 2
76

where (rw,d,d,b,L) is the function defined in (27) evaluated at the center of the well

screen. The specific gas capacity can be readily determined from the field measurements,

and the flow efficiency can be determined by scaling to an idealized value. Several

methods for determining Sco were evaluated and are described below, but for current

purposes it can be asssumed that values of F can be determined from field measurements

of specific capacity. Therefore, an estimate of the skin factor can be obtained from

specific capacity data

1 F
rw , d , d , b, L (71)
2F

The value (rw,d,d,b,L) is readily determined based on the geometry of the well. For,

example, the geometry of the conventional wells (b = 3 m, d = 8 m, rw = 0.05 m, L=0.3

m) used during the field work for this project yields = 3.59.

The skin factor is a non-linear function of the flow efficiency (71) and (Fig. 28).

Skin factor increases sharply as the flow efficiency decreases below 1.0, whereas it

decreases relatively gradually for F >1.0. As a result, skin factors for damaged wells

with F < 1.0 can be relatively large positive numbers, whereas they are relatively small

negative numbers for improved wells F>1.0. The skin factor asymptotically approaches a

negative value as F becomes large, so it is impossible to create wells with large negative

skin factors. For example, the smallest skin factor that can occur on a well characterized

by (rw,d,d,b,L) =3.59 and F is = -1.79 (Fig. 28).


77

10
8

Skin Factor
6
4
2
0
-2
0.1 1 10
Flow Efficiency
Flow Index

Figure 28. Skin factor as a function of flow efficiency for wells that are damaged F<1 and
wells that are improved, F>1.

The geometry of the well and its effect on well performance was investigated for

various anisotropy ratios using the flow efficiency as a function of the skin factor, ,

based on (71). Changing the depth of the well screen has a negligible effect on the flow

efficiency as a function of the skin factor (Fig. 29), but increasing the length of the screen

results in higher flow efficiencies for well skins adversely effects flow to the well ( >0),

and lower flow efficiencies for well skins enhancing flow to the well ( <0) (Fig. 29).

Reducing the radius of the well results in a similar response, with higher flow efficiencies

where >0, and lower flow efficiencies where <0 (Fig. 29).

Therefore, for situations where the skin is adversely effecting flow to the well,

increasing the screen length lessens the effect of a low permeability skin, because

increasing the screen length increases the area of the well and decreases the flux through

the skin at a given well discharge. Decreasing the flux reduces the head losses occurring

as gas travels through the skin resulting in a higher specific capacity thereby diminishing
78

the detrimental effect of the skin. The effect of well radius is due to the relationship

between the radius of the skin and the radius of the well, according to (43). Because, the

ratio of radii between the skin and the well must stay constant for a given skin factor,

decreasing the radius of the well also decreases the radius of the skin, thus producing a

greater flow efficiency for a given skin factor. The effect of anisotropy on well

performance was determined by varying the permeability ratio, kv/kh, from 0.1 to 10.

Conditions during the field tests resemble a case where the center of the well is at a depth

of 3.0 m, with a screen length of 0.3 m and a well radius of 0.05 m (Fig. 29), so those

well dimensions were used to determine the effects of anisotropy on flow to a

conventional well. For conditions where kr < 1.0, the flow efficiencies are increased for

well skins adversely affect flow to the well ( >0), and decreased for well skins

enhancing flow to the well ( <0). Conversely, for conditions where kr > 1.0, the flow

efficiencies are decreased for >0, and increased for <0.

This effect was observed because of the coordinate transform used to achieve

anisotropy in the model. For cases where the horizontal permeability is greater than the

vertical, the grid blocks were compressed horizontally, which reduces the radial thickness

of the well skin and increases the flow efficiency. Vertically dominated systems required

stretching the grid blocks in the horizontal direction, which increased the radius of the

skin adversely effecting flow to the well.


79

3
d = 3.0 m, L = 0.3m
d = 6.0 m, L = 0.3m
d = 3.0 m, L = 0.6m
2
d = 3.0 m, L = 1.2 m
well radius = 0.05m

0
-2 0 2 4 6 8 10
3
rw = 0.05 m
rw = 0.025 m
2
rw = 0.15 m
d = 3m, L = 0.3 m
1

0
-2 0 2 4 6 8 10
3
kratio=kv/kh = 1
kratio= 0.1
2
kratio = 10
d = 3m, L = 0.3 m, rw=0.05m
1

0
-2 0 2 4 6 8 10
skin factor,

Figure 29. Flow efficiency as a function skin factor for various well geometries
and anisotropies.
80

Fractured Wells

The numerical model was used to generate flow efficiencies for wells intersecting

fractures filled with material whose permeability contrasts with that of the formation, kr =

kfrx/kformation. The flow efficiency is determined by normalizing the predicted Sc for a well

with kr > 1.0 to Sco calculated using kr = 1.0.

Hydraulic fractures are represented as disks of uniform aperture in order to

evaluate the effects of basic fracture geometry on well performance. Flow efficiencies

for fractures of various radius, depth, and aperture were calculated for different values of

kr (Fig. 30).

The results show that flow efficiency is independent of fracture radius for kr <

200, but flow efficiency increases with radius for kr > 200. Interestingly, even the

relatively small radius of R = 1.0 m produces 10-fold increases in well performance when

kr = 1000. A depth of 3.0 m and an aperture of 0.005 m were used to determine the

effects of fracture radius (Fig. 30).

The depth of the fracture has only a minor effect on flow efficiency for a fracture

with a radius of 4.2 m and an aperture of 0.005 m. The effect is negligible where kr

<1000, but the flow efficiency is greater for shallow fractures compared to deep fractures

where kr = 1000 (Fig. 30).

Aperture is the most important of the three fracture geometry variables, affecting

the flow efficiency over the entire range of kr for a fracture at a depth of 3.0 m with a

radius of 4.2 m (Fig. 30).


81

R = 4.2
Rm=3m
R=2m
10 R=1m

= 0.005 m
b = 3m

1
1 10 100 1000 10000

d=2m
d=3m
d=4m
10
d=6m

= 0.005 m
R = 4.2 m

1
1 10 100 1000 10000

= 1.0
cm
= 0.5
cm
= 0.25
10 cm

d =3m
R = 4.2 m

1
1 10 100 1000 10000
kfrx/k

Figure 30. Flow efficiency as a function of the permeability ratio for various fracture
depths, radius, and aperture
82

Effects of Variable Aperture

The aperture of an ideal hydraulic fracture will be greatest near the injection point

and will decrease with radial distance [Murdoch and Slack 2002]. The thickness of sand

propping the fracture open, which is assumed to be the aperture, can vary markedly over

short distances, but the average thickness appears to taper smoothly toward the edge of

the fracture. The distribution of sand can be anticipated using a theoretical analysis of the

propagation of a shallow hydraulic fracture based on linear elastic fracture mechanics

[Murdoch 2002]. The aperture distribution in an idealized, flat-lying, horizontal

hydraulic fracture of radius a is

2 2
r 0 1 r a (72)

where 0 is the maximum aperture. Both the radius of the fracture and the maximum

aperture depend on the depth, properties of the formation, and volume of sand injected.

Rearranging eq. 9 in [Murdoch 2002] gives

a 1.117 d 3 8 E 1 4 K IC1 4V 1 4 (73)

34 12 12 12
0 0.7646d E K IC V (74)

where d is the depth of the fracture, E is the elastic modulus, KIC is the fracture

toughness, and V is the volume of sand.

Substituting (73) and (74) into (72) results in an expression for the distribution of

fracture aperture as functions of the volume of injected sand, depth, and formation
83

properties. The analysis used to develop the expressions above is for fractures that are

large relative to their depth, whereas small fractures are expected to behave differently.

Fractures that are extremely large relative to their depth will probably become

asymmetric and will differ from the geometry assumed above [Murdoch and Slack 2002].

As a result, the analysis given above is appropriate when d < a < 2d, where a is the radius

of the fracture and d is the depth of fracture initiation.

Fractures created for field tests during this project are within several meters of the

ground surface, so a depth of 3 m will be used as an example. Fracture toughness and

elastic modulus of materials at the field site are unknown, but approximate values of KIC

= 0.05 MPa m1/2 and E = 20 MPa are consistent with other shallow field settings

[Murdoch 2002] and yields values of aperture that are consistent with field observations.

Sand volumes used to fill hydraulic fractures at a depth of 3 m are in the range of several

tenths of a cubic meter, and 0.3 m3 was used during the field tests.

The radial length and the aperture increases as a function of the injected volume,

according to (72) through (74), and (Fig. 31). Increasing the volume of injected sand

from 0.1 to 0.5 m3 increases the maximum aperture from 5.5 mm to 12.5 mm, which is

consistent with (74). Similarly, increasing the injected volume by a factor of 5 will

increase the radial length by a factor of 1.5, from 4.0 to 6.0 m. The actual radial length of

the fracture in the field is expected to be slightly less that that shown in (Fig. 31)

becausesand is unable to fill the narrow apertures at the leading edge of the fracture.

Typically, apertures less than 0.002 m are expected to mark the extent of the sand-filled

portion of the fracture. For theoretical analysis this effect is ignored and it is assumed

that the profiles of sand are as shown in (Fig. 31).


84

0.012
Sand
0.010 vol0.1 m3

Aperture
3
0.008 0.3 m
3
0.5 m
0.006
0.004
0.002
0.000
0 1 2 3 4 5 6
r (m)
Figure 31. Aperture as a function of the radial distance from the injection well for
various sand volumes

The flow efficiency increases with the permeability ratio, kr, for various volumes

of sand in a hydraulic fracture (Fig. 32). The change in flow efficiency is minor for kr =

10, whereas for kr = 100, the flow efficiency is between 2.5 to 4, depending on the

amount of sand injected. Flow efficiency continues to increase as kr is raised to 1000

times that of the formation, resulting in a flow efficiency that ranges from 10 to 20.

If a 100% increase in the flow efficiency is a significant effect, then a fracture

permeability 30 to 50 times greater than the formation permeability is needed to achieve

this effect, depending on the volume of the fracture. Therefore, based on the numerical

solution, the fracture permeability must be at least 1.3 to 1.5 orders of magnitude greater

than the formation permeability to have a significant effect on specific capacity (Fig. 32).
85

10 Sand vol
0.1 m3
0.3 m3
0.5 m3
1
10 100 1000 10000
kfrx/k
Figure 32. Flow efficiency as a function of the ratio between the permeability of
the fracture and permeability of the formation for different fracture sand volumes.

Pressure Distributions

The pressure gradient in the subsurface is proportional to the flux, based on the

permeability of the formation. Pressure distributions were generated with the numerical

model to compare the performance of wells intersecting hydraulic fractures based on the

permeability ratio between the fracture and formation, kr, as well as the geometry of the

fracture. The pressures presented in this section have been normalized to the pressure at

the well, which was held constant at 0.9 atm, according to

Patm Pobs
Pd (75)
Patm Pwell
86

The pressure interval ranges from 0.01 to 0.30 the pressure at the well. For

performance comparisons it is convenient to establish the radius of pressure influence for

these plots by the radial location of the 0.1 pressure contour at the depth of the center of

the well screen.

The pressure is confined to the immediate vicinity of a conventional well, with the

radius of pressure influence located a few tenths of a meter away from the extraction well

(Fig. 33). A fracture that has a permeability 100 times greater than the formation results

in a pressure distribution that is broadened slightly, with decreased pressures in the

vicinity of the well. The radius of pressure influence is extended to 1.2 m from the

extraction well. More striking changes in the distribution of pressure are observed when

kr = 1000. The pressures field is intensified significantly around the fracture, and radius

of pressure influence is increased near 5.0 m. When the fracture permeability is

increased by another order of magnitude, to kr = 10,000, the effect continues to be

significant, with the radius of pressure influence now occurring beyond the 8.0 m radial

boundary of the model.


87

0 0
0.1 0.3
0.5
-2 -2
0.7

-4 -4 0.1 0.01
0.01
-6 0.9 -6

kr = 1 kr = 100
-8 -8
0 2 4 6 8 0 2 4 6 8
0 0
0.01 0.01
0.1
-2 -2

-4 -4

0.1
-6 -6
kr = 1000 kr = 10000
-8 -8
0 2 4 6 8 0 2 4 6 8
meters meters
Figure 33. Normalized pressure contours (solid lines) and streamlines (dashed
lines) for varying ratios of fracture sand permeability and formation permeability.
50% streamline is highlighted for flow comparisons.

The performance of wells is determined, in part, by the comparison of pressure

distributions and flow paths in the vicinity of wells. Comparisons are determined by

adjusting parameters associated with fractures. Separate analysis leads to the conclusion

that by specifying a fracture volume, the fracture aperture and radial extent can be

defined. Therefore, an analysis similar to the permeability ratio analysis can be

performed using normalized pressures associated with various injected sand volumes for

hydraulic fractures with a given sand permeability.


88

Conditions observed during field experiments suggest that 100 < kr < 1000, so kr =

1000 was used in order to investigate the effects of fracture volume on well performance.

As expected, pressure is confined to the immediate vicinity of a conventional vapor

extraction well, with the radius of pressure at a distance of only a few tenths of a meter

(Fig. 34). However, with the introduction of a fracture with only 0.1 m3 (350 lbs.) of

sand, we observe a significant effect on the pressure field (Fig. 34). The radius of

pressure influence is increased to 3.7 m, which is equivalent to the radius of the fracture,

and pressures are intensified in the vicinity of the well screen and the fracture. According

to fracture form analysis, as the volume of the fracture increases, the radius and aperture

are both increased [Murdoch 2002]. Because of this relationship between the fracture

volume and the radius, it appears that the radius of pressure continues to increase as a

function of the fracture radius. For a 0.1 m3 fracture, the extent of suction occurs at a

radial distance of approximately 3.7 m from the well, which is equal to the predicted

radius of the fracture at that volume. A similar relationship is observed for fracture

volumes of 0.3 m3 (1050 lbs. of sand) and 0.5 m3 (1750 lbs. of sand), with the radius of

pressure influence being equivalent to the predicted radius of the fracture (Fig. 34).

Therefore, with regard to fracture volume and its effect on subsurface airflow, the radial

extent of suction is comparable to the size of the fracture. It should be noted that larger

diameter fractures, such as a 0.5 m3 fracture at a depth of 3.0 m, may be difficult to create

in the field, because of their tendency to climb to the ground surface.


89

0 0
0.1 0.3 0.01
0.5
-2 -2
0.7

-4 -4
0.01
0.1
-6 0.9 -6

Sand Vol.=0.0 0.1m3


-8 -8
0 2 4 6 8 0 2 4 6 8
0 0
0.01 0.01

-2 -2

-4 -4

0.1 0.1
-6 -6

0.3m3 0.5m3
-8 -8
0 2 4 6 8 0 2 4 6 8
meters meters

Figure 34. Normalized pressure distributions and streamlines for varying fracture/sand
volumes.

Stream Function

The effects of hydraulic fractures on subsurface flow paths is determined by

varying the permeability ratio and volume of sand in the fracture and evaluating the

location of the 50% flow line at the ground surface. For a conventional well at a depth of

3.0 m, 50% of the flow enters the ground approximately 4.5 m away from the well and

curves downward towards the well screen (Fig. 33 and Fig. 34). In general, for short-
90

screened wells that are significantly above the lower boundary, the 0.5 streamline enters

the ground approximately 1.5 times the depth to the well screen. This radial distance is

increased to 5.5 m (1.8 times depth) for a well intersecting a fracture kr = 1000, and to

more than 6.0 m (2.0 times depth) for a fracture with kr = 10,000 (Fig. 33). Similar

effects are observed for varying the volume of fracture with kr = 1000 (Fig. 34). The

largest fracture investigated for this study (0.5 m3) increased the radial location of the 0.5

streamline to approximately 6.0 m, or 2.0 times the depth of the fracture. It appears that a

conductive disk-shaped fracture changes the pattern of airflow by moving the radial

extent outward away from the well along the ground surface, with the location of the 0.5

streamline increased by a factor of 1.33. This increase in radial distance means at the

ground surface, the 0.5 streamline for a well intersecting a productive fracture forms a

ring whose area is approximately 1.8 times that of conventional well.

FIELD OBSERVATIONS

Field observations were used to characterize the sites and acquire data to calibrate

the theoretical models. Measurements of discharge, specific gas capacity, and subsurface

pressure in the vicinity of wells during gas pumping tests constitute much of the

calibration data. Core samples were taken to determine the geology at the field sites, and

to obtain an estimate of fracture form. The presence of a low permeability skin around

well bores was also detected during field experiments as a possible source of observed

variations in discharge.
91

Fracture Form

Split-spoon core samples were taken during the installation of the multi-level

piezometers. Measurements of fracture depth and thickness were recorded from samples

taken in the vicinity of the coarse-grained fracture at the south site to estimate the fracture

geometry. The fracture was found at a depth of 3.00 m at radial distances of 0.3, 0.6, and

1.5 m away from the injection well, and at a depth of 2.92 m at a radial distance of 3.0 m

from the well. The fracture was absent in samples taken at a radius of 4.5 m.

Fracture apertures range from 0.013 m to 0.005 m, with the greatest aperture

observed 0.3 m from the well, and the smallest aperture observed at a radial distance of

3.0 m from the well (Fig. 35).

0.014
0.012
a pe rture (m )

0.01
Observed apertures
0.008
0.006
0.004
0.002
0
0 1 2 3 4 5
ra dius (m )

Figure 35. Observed fracture apertures as a function of the radial distance from the
well used to create the coarse-grained fracture at the south site.
92

Gas Pumping Tests

Gas pumping tests were conducted at the field sites in order to obtain information

about the effects of hydraulic fractures on subsurface airflow in the vicinity of SVE

wells, and determine which method of conventional well completion would best suit SVE

at the field sites. Pumping tests were performed in the conventional well clusters using

each well for extraction, with the remaining wells used as pneumatic piezometers. The

pipe used to create the hydraulic fractures served as the extraction well during vapor

extraction tests. Multi-level piezometers were the primary tool for obtaining pressure

distributions in the vicinity of wells intersecting fractures. A few conventional wells

were also installed near the fracture to ensure data from the multi-level piezometers were

reliable.

For all pumping tests, a period of 10 minutes was allowed for equilibration in the

subsurface. This was determined in the field to be sufficient based on the time lapse

between changes in observed pressure. The pressures were assumed to be equilibrated if

intervals longer than 10 seconds passed between changes in the observed pressure.

Conventional Wells

The first step in understanding the field conditions is to characterize the

performance of conventional wells at the field sites. There are a total of 17 conventional

wells installed in clusters at the two sites, with 10 at the north site, and 7 at the south site

(Fig. 37). However, two of the conventional wells at the south site were used only as

piezometers (SW-4, SW-5), because they were installed at a depth of 1.5 m using a 1.2
93

cm (0.5 in) diameter casing, which does not match the dimensions of the vapor extraction

manifold. Hence, there are effectively 5 conventional extraction wells at the south site.

The discharge of conventional wells at the north site ranges from 0.5 to 7.1

standard cubic feet per minute (scfm) (Table 5). The arithmetic mean of the discharges is

2.5 scfm and the geometric mean is 1.3 scfm. Pressures observed at the extraction wells

of the north site conventional well cluster range from 44.4 to 48.9 in of water. The

arithmetic mean pressure is 46.7 in of water and geometric mean is 46.7 in of water.

The discharge of conventional wells at the south site ranges from 4.0 to 14.5

standard cubic feet per minute (scfm) (Table 5). The arithmetic mean of the discharges is

8.5 scfm and a geometric mean is 7.4 scfm. The pressures observed range from 23.2 to

43.5 in of water. The arithmetic mean pressure is 34.2 in of water, and the geometric

mean is 33.2 in of water.

Table 5. Observed pressures and discharges for a. North Site b. South Site. Specific
capacities calculated from these data.

a. North Site
Well Pobserved (in. H20) Pobserved (atm.) Q (scfm) Sc (scfm/atm2)
NW-1 46.9 0.8848 0.6 2.3
NW-2 48.9 0.8799 0.5 2.2
NW-3 44.4 0.8892 7.1 33.8
NW-4 45.0 0.8878 6.1 28.6
NW-5 48.5 0.8809 0.4 2.2
NW-6 47.4 0.8836 0.5 2.3
NW-7 46.5 0.8858 1.0 4.7
NW-8 46.1 0.8868 1.9 8.7
NW-9 45.6 0.8880 5.9 27.7
NW-10 48.0 0.8821 0.6 2.7
94

b. South Site
Well Pobserved (in. H20) Pobserved (atm.) Q (scfm) Sc (scfm/atm2)
SW-1 43.5 0.8932 4.0 19.7
SW-2 23.2 0.9430 12.5 112.6
SW-3 25.7 0.9369 14.5 118.9
SW-6 39.6 0.9027 4.0 21.6
SW-7 39.1 0.9040 7.5 40.9

Specific capacities of wells within the conventional cluster at the north site range

from 2.2 to 33.8 scfm/atm2 (Table 5). The arithmetic mean is 11.5 scfm/atm2 and the

geometric mean is 6.1 scfm/atm2. The specific capacities of wells at the south site range

from 19.7 to 118.9 scfm/atm2, with an arithmetic mean of 62.7 scfm/atm2 and a geometric

mean of 47.1 scfm/atm2.

The specific capacity of a well is dependent on the discharge and the pressure

applied at the well screen according to (68). A low permeability skin around the well

screen can decrease the flux into the well and increase the pressure at the well, both

reducing the specific capacity. Initial observations revealed an apparent correlation

between the specific capacity of conventional wells and the drilling method used for well

installations (Fig. 38). The well cluster at the north site has wells installed using machine

augers, hand augers, and a combination of machine augers and the shelby-tube removal

of the screened interval. Conventional wells at the south site were installed using

machine augers and hand augers. Boreholes created using a hand auger generally

resulted in wells with the highest discharge and specific capacity, whereas those created
95

using machine augers resulted in the lowest discharge and specific capacity. This effect

will be investigated further by the evaluating the field data with theoretical analysis.

A gas pumping test performed at the north site on NW-3 serves as an example.

This test took place on Oct. 8, 2001, from 12:20 pm to 12:45 pm. The initial discharge

was 7.2 scfm at the onset of pumping, but decreased to 7.1 after 5 minutes, and remained

constant throughout the next 20 minutes. Pressure at the well was initially 45.9 in water,

and decreased to 45.1 in water in accordance with the discharge.

Observed subsurface pressures decrease as a function of radial distance from the

extraction well (Fig. 36). NML-2 and NW-5 are the only piezometers completed at

depths other than 3.0 m in the conventional well cluster (Table 6). NML-2 is a multi-

level piezometer with ports ranging from depths of 1.5 m to 3.6 m, and NW-5 is a

conventional well completed at a depth of 1.5 m. Pressures appear to range over 2.5 in

water, equaling a few hundred pascals, at a depth of 3.0 m and a radial distance of 0.3 m

from the extraction well. Failure to reach subsurface pressure equilibration is likely to

impact the results, as are local heterogeneities. Further analysis suggests that the

equilibration time of 10 minutes was only sufficient for moderate to high flow rates, with

more time being required for low flows. This effect would be compounded when the

extraction well and the observation well produce low flows. Failure to reach

equilibration would result in observed pressures that are systematically less than

pressures predicted by the model.


96

Table 6. Piezometer observations during gas pump test on MW-3 at north site

Depth(m) NML-2 NW-1 NW-2 NW-3 NW-4 NW-5 NW-6 NW-7 NW-8

1.5 0.48 1.02

2.4 1.38

3.0 1.69 1.11 3.03 5.30 1.70 3.58 3.74

3.6 2.28

5
Pressure (in. water)

0
0 0.2 0.4 0.6 0.8
radius
Radial (m)
distance (m)

Figure 36. Pressure as function of distance away from extraction well NW-3.
97

North Site Well Cluster

NW-6 NW-7

NW-2

NML-2 NW-4 NW-3 NW-9

NML-1
NW-10 NW-5
NW-8

NW-1

South Site Well Cluster

SW-6

SW-7

1.0
Explanation
Machine auger
N
SML-1 SW-1 Machine auger/Shelby tube

Hand auger

SW-4 SW-2 Multi-level Piezometer

SW-3
1m
SW-5

Figure 37. Well locations within conventional clusters at north and south site
98

North Site Well Cluster

2.2 4.6

2.2
28.6 33.8 27.7

2.7 2.23
8.7

2.3

South Site Well Cluster

21.6

40.9

1.0 Explanation

N Machine auger

Machine auger/Shelby tube

118.9 Hand auger

Multi-level Piezometer
112.6
Specific Capacity (scfm/atm)

19.7

1m

Figure 38. Observed specific capacities (scfm/atm2) of conventional wells in clusters at north
and south sites.
99

Fractured Wells

Four hydraulic fractures were created at the field sites: two at the north site and

two at the south site. Each site has a fracture propped with medium-grained sand and a

fracture propped with coarse-grained sand. The permeability of the sands in the fractures

was determined using a shelby-tube permeameter (Fig. 17). The permeability of the

medium-grained sand is 3.0 x 10-9 m2, and the permeability of the coarse-grained sand is

5.5 x 10-9 m2.

At the north site, the medium-grained sand fracture has a discharge of 26.5 scfm,

and the coarse-grained fracture has a discharge of 23.5 scfm, yielding a mean discharge

of 25.0 scfm. A pressure of 17.5 in of water was observed at the extraction well

intersecting the medium-sand fracture at the north site, and 24.8 in of water was observed

at the well intersecting the coarse-sand fracture.

At the south site, the medium-grained fracture has a discharge of 25.0 scfm, and

the coarse-grained fracture has a discharge of 28.0 scfm, yielding a mean discharge of

26.5 scfm. The well intersecting the medium-sand fracture at the south site has an

observed pressure of 20.1 in of water, while the well intersecting the coarse-sand fracture

has a suction of 12.3 in of water.

The medium-grained fracture at the north site has a specific capacity of 315.5

scfm/atm2, while the coarse-sand fracture has a specific capacity of 199.2 scfm/atm2. The

average specific capacity is 257.4 scfm/atm2 for wells intersecting fractures at the north

site. The medium-grained fracture at the south site has a specific capacity of 260.4

scfm/atm2, and the coarse-sand fracture has a specific capacity of 383.6 scfm/atm2,

averaging 322.1 scfm/atm2.


100

The pumping test conducted on the coarse-sand fracture at the south site serves as

an example of a test on a well intersecting a hydraulic fracture. The test was performed

on Oct. 30, 2001 from 1:25 pm to 2:00 pm. The initial discharge was 30.5 scfm, reducing

to 28.0 scfm after 10 minutes, and remaining constant for the next 25 minutes.

Pressures observed during this test (Table 7) increase with increasing distance

from the extraction well, and increasing distance above and below the fracture (Fig. 39).

The fracture was observed to terminate between 3.0 and 4.5 m away from the well (Fig.

35) at a depth of 2.99 m based on split-spoon data. Pressures become more uniformly

distributed for distances beyond the radius of the fracture, and are observed to have lower

values above the fracture than below the fracture at the same vertical distances. It also

appears that at 0.4 m above the fracture, the observed pressures within a radius of 1.5 m

converge.
101

Table 7. Recorded pressure observations during gas pump test for the coarse fracture at the
south site

Depth (m) ML-1 E ML-2 E MW-3 ML-5 E ML-10 E ML-15

2.1 2.89

2.4 3.51 3.34 3.73 3.20 2.51

2.7 4.44

3.0 4.95 4.76 4.37 4.53 3.30 2.44

3.3 4.73

3.6 3.93 4.32 3.12 2.89 2.33

3.9 3.48

4.3 3.42

1.5

2.5
depth (m)

0.3m
3 0.6m
1.5m
3.0m
3.5
4.5m

4.5
0.985 0.99 0.995 1
pressure (atm)

Figure 39. Observed pressures as a function of depth along selected radial distances.
Pressures were observed during a gas pump tests on the coarse sand fracture at the south site.
Fracture is at a depth of 3.0m.
102

EVALUATION OF FIELD DATA

Field observations are used to determine the validity of both the solution for flow

to a well with a skin and flow to a well intersecting a fracture. Pressures observed during

gas pumping tests are compared with pressure distributions predicted with the numerical

and analytical models. Specific capacities predicted by the numerical model as a function

of the permeability ratio are compared to observed specific capacities from the field.

Skin factors associated with well installation methods and resulting specific capacities are

also presented in the following section.

According to a fracture form analysis [Murdoch 2002], the aperture and radius of

the fracture are determined by its volume. The expected aperture as a function of the

radial distance away from the injection well for varying fracture volumes is presented in

(Fig. 31). Fracture apertures were obtained in the field by taking split-spoon core

samples (Fig. 35) and were compared with the predicted apertures according to the

analysis (Fig. 40). Because the fractures created in the field had 1000 lbs. of sand

injected, it is expected that the apertures would be similar to those predicted along the 0.3

m3 (1050 lbs. of sand) curve. It was observed that the predicted apertures are similar to

the observed apertures, with a slightly thicker fracture close to the injection well, and the

termination of the fracture occurring within 5.0 m.


103

0.012
Sand
0.010 vol0.1 m3
3
0.008 0.3 m
3
Observed
0.5 m Sand Thickness
0.006
0.004
0.002
0.000
0 1 2 3 4 5 6
r (m)
Figure 40. Comparison of observed apertures from coarse-sand fracture at south site to
predicted apertures for a 0.3 m3 volume fracture, with E = 20MPa KIC = 0.05 MPam1/2.

The analytical solution solves for subsurface pressures based on horizontal and

vertical permeabilities and discharge from the well. This solution is inverted to give

estimates of the permeabilities based on observed pressures and well discharge during gas

pumping tests by minimizing the difference between the observed pressures and the

predicted pressures. A skin factor associated with each extraction well is also determined

based on the difference between the observed and predicted pressure at the well at a

specified flow rate. Individual pumping tests were analyzed to determine a horizontal

and vertical permeability, and a skin factor. Pumping tests were also analyzed

simultaneously, or as an ensemble, to give values of the horizontal and vertical

permeabilities that best fit the data. However, skin factors were only determined for the

individual analyses. Estimates were obtained by embedding (26) and (42) in a solving

routine that minimizes the residual between the observed pressures and the pressures
104

predicted by the model to gain best fit values for the horizontal permeability, anisotropy

ratio, and in case of individual tests, the skin factor.

The quality of the fit obtained from the analytical solution is expressed by the

Pearson product moment correlation coefficient, r. The correlation coefficient is

determined by

SSR (n 1)
r 1 (72)
SSM (n m 1

where n is the number observation pairs, and m is the number of parameters involved in

the analysis. Using (72) to calculate the correlation coefficient can only be positive, but

because r values can range from 1 r 1, the slope of the best fit line through the data

must be used to determine the appropriate positive or negative value. Individual r values

were calculated for pumping tests, as well for the ensemble analyses. Correlation

coefficients for individual analyses ranged from 0.998 to 0.999 at the north site, with an

average value of r= 0.999, and from 0.997 to 0.999 at the south site, with an average

value of r=0.999, whereas the ensemble analyses yield correlation coefficients of

r=0.931 for the north site, and r=0.889 for the south site. Confidence limits are set on r

using the tables for 99% confidence limits presented by Steel and Torrie [1960]. The

correlation coefficient for the north site is 0.931 according to (72), and has a 99%

confidence interval ranging from 0.830 to 0.960. The south site has an r value of 0.889,

and has a 99% confidence interval ranging from 0.500 to 0.970.

The sensitivities of the vertical and horizontal permeabilities were determined

based on 1% significance levels for the appropriate degrees of freedom, which yields a
105

1% significance level of 0.470 for the north site, and 0.708 for the south site. The r

values were contoured, wherein contour lines connect points of equal correlation

coefficient under varying anisotropic conditions (Fig. 41). The 1% significance levels

determine the boundaries of the plots, such that actual field conditions are 99% likely to

be contained within the contours of the plots.

A variety of methods were used to determine representative permeabilities for the

field sites. Pumping tests analyzed simultaneously yield the most accurate values for

representative horizontal and vertical permeabilities based on the correlation coefficients,

r. This method results in kh = 5.25 x 10-12 m2, kv = 3.81 x 10-11 m2, and r = 0.930 for the

north site, and kh = 1.32 x 10-11 m2, kv = 1.03 x 10-11 m2 ,and r = 0.889 for the south site

(Tables 8 and 9). These values were obtained by allowing for anisotropic conditions in

the model. Well performance comparisons based on the permeability ratio between the

fracture and the formation required a single representative value of the formation

permeability. Constraining the models to isotropic conditions yielded permeability

values with correlation coefficients similar to anisotropic conditions. At the north site k =

1.51 x 10-11 m2 and r = 0.920, and the south site k = 1.15 x 10-11 m2 and r = 0.887 (Tables

8 and 9). The sensitivity analysis was performed on the anisotropic ensemble models for

both sites, because that method resulted in the highest correlation coefficient associated a

representative horizontal and vertical permeability. Individual pump tests were also used

to gain representative permeabilities by averaging the results arithmetically and

geometrically (Figs. 43, 44, and 45). Permeabilities based on the arithmetic averages

resulted in the lowest correlation coefficients, with the north site permeabilities having

r = 0.023 and south site having r = 0.000 (Table 9). The values obtained by geometric
106

averaging of the individual results were plotted on the sensitivity analysis. At the north

site kh = 1.47 x 10-11 m2, kv = 3.03 x 10-11 m2, and r = 0.830, and at the south site kh = 2.03

x 10-11 m2, kv = 1.66 x 10-11 m2, and r = 0.644(Table 8 and 9).

The sensitivity analysis for the south site revealed that permeability fits with r

values greater than 0.850 ranged from 1.00 x 10-11 m2 to 1.75 x 10-11 m2 in the horizontal

direction, and from 4.00 x 10-12 m2 to 1.85 x 10-11 m2 (Fig. 41). The sensitivity analysis

for the north site indicates that the permeability fits with r values greater than 0.900 range

from 2.30 x 10-12 m2 to 1.59 x 10-11 m2 in the horizontal direction, and from 9.00 x 10-12

m2 to beyond 1.33 x 10-11 m2 in the vertical direction (Fig. 41).

Table 8. Permeability estimates from individual pump tests analysis and coupled analysis.

North Site South Site


khorizontal (m2) kvertical (m2) khorizontal (m2) kvertical (m2)
7.36 x 10-13 1.11 x 10-10 5.73 x 10-11 7.45 x 10-12
7.33 x 10-12 2.97 x 10-11 6.54 x 10-12 4.91 x 10-11
1.16 x 10-10 5.80 x 10-12 2.24 x 10-11 1.25 x 10-11
1.43 x 10-11 2.86 x 10-11

Arithmetic 4.31 x 10-11 4.43 x 10-11 2.87 x 10-11 2.30 x 10-11


mean
Geometric 9.73 x 10-12 2.72 x 10-11 2.03 x 10-11 1.66 x 10-11
mean
Standard ±5.46 x 10-11 ±4.62 x 10-11 ±2.60 x 10-11 ±2.27 x 10-11
deviation
Isotropic 1.51 x 10-11 1.51 x 10-11 1.15 x 10-11 1.15 x 10-11
(ensemble
analysis)
107

Anisotropic 5.25 x 10-12 3.81 x 10-11 1.32 x 10-11 1.03 x 10-11


(ensemble
analysis)

Table 9. Pearson product moment correlation coeffecient values for permeability estimation
methods for the North and South sites.

Permeability Estimation Method North Site South Site


Pearson Coefficients Pearson Coefficients
Arithmetic mean 0.023 0.000
(individual analysis)
Geometric mean 0.830 0.644
(individual analysis)
Isotropic conditions 0.920 0.887
(coupled analysis)
Anisotropic conditions 0.930 0.889
(coupled analysis)
108

1.2E-10
a.
1.0E-10
kvertical (m2)

8.0E-11

6.0E-11

4.0E-11

2.0E-11

2.0E-12 6.0E-12 1.0E-11 1.4E-11


2
khorizontal (m )

3.0E-11

b.
2.5E-11

2.0E-11
kvertical (m2)

Anisotropic Best Fit


1.5E-11 (Ensemble Analysis)
Isotropic Best Fit
(Ensemble Analysis)
1.0E-11
Geometric Mean
(Individual Analysis)
5.0E-12

5.0E-12 1.5E-11 2.5E-11 3.5E-11

khorizontal (m2)

Figure 41. Sensitivity of permeability estimates from analytical solution for a. North Site and
b. South Site. Sensitivities based on 1% significance levels.
109

Skin Factor

Skin factors obtained from the analytical solution for individual pumping tests

show a correlation between the well skin and the observed specific capacity. There also

appears to be a correlation between the drilling methods and the resulting specific

capacities and skin factors (Fig. 42 and 43). Of the three drilling methods used at the

sites, hand-augered boreholes have the smallest positive skin factor, whereas that of

machine augered boreholes is the greatest. Wells installed in boreholes created by the

shelby-tube removal of the screened interval have a specific capacity and a skin factor of

intermediate value.

The skin factors at the north site range from 0.84 to 26.0 (Fig. 42), and skin

factors at the south site range from 0.54 to 6.02 (Fig. 43). NW-5 was installed in a

hand-augered borehole at the north site at a depth of 1.5 m, which put the screened

interval of this well 1.5 m above the other wells in the center of the clay-rich horizon,

resulting in a well with a skin factor of 26.06.


110

400

350

Specific Capacity (scfm/atm)


300

250

200

150

100

50

0
Med.Frx.
NW-1 NW-2 NW-6 NW-10 NW-7 NW-8 NW-3 NW-4 NW-5 NW-9 Crs.Frx.

30

25

20
Skin Factor

15

10

-5

Machine auger
Machine auger/Shelby Tube
Hand auger
Fractured well

Figure 42. Specific capacities and skin factors associated with all wells at the North site.
111

400

350

Specific Capacity (scfm/atm)


300

250

200

150

100

50

0
M ed. Frx
SW -1 SW -2 SW -3 SW -7 SW -6 Crs. Frx

30

25

20
Skin Factor

15

10

-5

Machine auger
Hand auger
Fractured well

Figure 43. Specific capacities and skin factors associated with all wells at the South Site
112

The analytical solution for determining a skin factor was also used to analyze the

wells intersecting fractures using only the flow rate and pressure observed at the well

during gas pump tests. As discussed in the Theoretical Analysis Results section of this

report, the relationship between the flow efficiency and skin factor is non-linear (71).

This non-linearity makes it possible to have relatively large values for skin factor when

flow is adversely affected by the well skin ( > 0), but small values when the wells show

improved flow ( < 0) (Fig. 28).

Wells intersecting sand-filled fractures are characterized by negative skin factors.

At the north site, the medium-sand fracture has a skin factor of 0.44, and the coarse-sand

fracture has skin factor of 0.37 (Fig. 42). At the south site, the medium-sand fracture

has a skin factor of 1.17, and the coarse-sand fracture has a skin factor of -1.25 (Fig.

43).

Individual analysis of the conventional wells at the south site show two other

wells with negative skin factors corresponding to unusually high specific capacities, with

SW-2 and SW-3 having skin factors of 0.54 and 0.63, respectively. The anomalous

skin factors associated with these wells could be attributed to local heterogeneities, such

as a high permeability layer or even cracks leading down from the ground surface.

Another source of these values could be associated with the well completion, as shrinking

of the cement used to fill the annulus of the conventional wells could result in a conduit

connecting the ground surface to the gravel pack of the well.


113

Flow Efficiency

The numerical model predicts the discharge and the pressure distribution in the

vicinity of the fracture as a function of the formation permeabilities, the fracture sand

permeability, and the pressure at the well. This model was used to generate specific

capacity curves as a function of the permeability ratio for varying sand volumes as

discussed in the Theoretical Analysis Results portion of this report. The predictions of

specific capacity can be compared to the observed specific capacities based on

permeability ratios similar to field conditions. Determining a permeabililty ratio for a

particular fracture requires a single representative permeability for the sand in the

fracture, and for the formation. The permeabilities of the fractures are assumed to be the

characterized by the permeabilities of the sands filling the fractures, which were

determined with the shelby-tube permeameter. Site specific formation permeabilities

were estimated using the ensemble analytical solution constrained to isotropic conditions

(Table 8). This yields kr= 364.0 for the coarse-sand fracture, and 198.7 for the medium-

sand fracture at the north site. At the south site, kr= 478.3 for the coarse-grained sand

fracture, and kr= 260.9 for the medium-grained sand fracture.

Validating the numerical predictions of flow efficiency as a function of the

permeability ratio required not only permeability ratios from the field, but also actual

flow efficiencies. Flow efficiency is the ratio of observed specific capacity to the specific

capacity of a conventional well operating under ideal conditions, Sco (69). Specific

capacities ranged over an order of magnitude within the conventional well clusters,

depending on the well completion method. Therefore, determining a specific capacity


114

that represents the performance of conventional wells is a somewhat challenging task.

Possible representative specific capacity values were taken as the arithmetic and

geometric average specific capacities within a well cluster. Individual averages

associated with well completion techniques were also used, as clusters consist of uneven

numbers of wells installed with different methods. For instance, the north site has 4 hand

augered wells, 4 machine augered wells, and 2 shelby-tube wells, whereas the south site

has 4 hand augered wells, and 1 machine augered well. Averaging the specific capacities

of a well cluster as a whole would result in a value that is skewed to the specific

capacities associated with the predominant drilling method. The specific capacity of a

well lacking skin was determined theoretically, and was also considered as a candidate

for Sco.

Flow efficiencies for the conventional wells at the north site range from 0.035 to

0.537, and at the south site from 0.250 to 1.507 (Table 10). These values are based on

the performance of a conventional well operating under ideal conditions determined

analytically to represent Sco .

Flow efficiencies for well intersecting fractures at the north site range from 5.0 to

133.0 for the medium-grained fracture, and from 3.2 to 84.0 for the coarse-grained

fracture, depending on the method used to estimate Sco (Table 11). Flow efficiencies at

the south site range from 3.3 to 12.1 for the medium-grained fracture, and from 4.9 to

17.8 for the coarse-grained fracture. Flow efficiencies based on the average of specific

capacities associated with machine-augered wells to determine Sco result in the highest

values, which are significantly greater than those predicted by the model (Fig. 44). Flow

efficiencies, with Sco estimated from the analytical solution for conventional wells
115

assuming no skin, represent the lowest observed values, and are below predicted flow

efficiencies from the numerical solution. Other flow efficiencies determined by the

averaging of specific capacities range between these two extremes, with hand augered

wells having the least skin factors and greatest specific capacities, resulting in the second

lowest values for flow efficiency. Not surprisingly, flow efficiencies based on the

arithmetic and geometric averages of the observed specific capacities to determine Sco lie

within the range of individual completion method averages.

The variations in flow efficiency are partly due to the effect of a low permeability

skin created by well completion techniques. Wells installed using a machine-auger

resulted in the lowest specific capacities observed at the field sites. Whereas wells

installed using hand-augers to create the borehole resulted in the highest specific

capacities (Fig. 42 and 43). Therefore, the flow efficiencies are significantly greater

when the performance of machine-augered wells is used for Sco , rather than the

performance of hand-augered wells (Table 11). The specific capacity of a well lacking a

skin was determined theoretically, and represents the performance of a well under ideal

conditions. Based on the performance of a well under ideal conditions, all wells at the

field sites perform below optimum levels.


116

Table 10. Flow efficiencies for conventional well based on analytical determination of Sco.

Well North Site South Site


1 0.037 0.250
2 0.035 1.427
3 0.537 1.507
4 0.455
5 0.035
6 0.036 0.274
7 0.074 0.518
8 0.138
9 0.440
10 0.043

Table 11. Flow efficiencies for fractured wells with varying methods of determining Sco.

Overall Estimates Individual Estimates


Analytical Arithmetic Geometric Hand Machine Shelby-
Auger Auger Tube
North Site
Medium- 5.0 27.4 51.4 10.5 133.0 47.3
sand fracture
Coarse-sand 3.2 17.3 32.5 6.6 84.0 29.9
fracture
South Site
Medium- 3.3 4.2 5.5 3.6 12.1
sand fracture
Coarse-sand 4.9 6.1 8.1 5.3 17.8
fracture
117

100

10

1
1 10 100 1000 10000
k /k
frx

Flow Efficiency
Well Cluster Estimates
North Site Analytical optimum
South Site Arithmetic average
Medium-sand fracture Geometric average
Coarse-sand fracture Well Completion Averages
Hand auger
Machine auger
Shelby-tube

Figure 44. Comparison of predicted Flow Efficiency to observed values as a function of the
permeability ratio for a 0.3 m3 fracture. Various methods of determining representative
specific capacities for conventional wells resulted in a range of Flow Efficiencies for each
fracture.
118

The differences between the predicted and the observed flow efficiencies must be

attributed to either errors in the estimates of the performance of conventional wells at the

field sites or to overestimates of the permeability ratio. The pearson coeffiecients

associated with the individual pump tests indicate a good fit to the data (0.999), so it can

be assumed that the error does not lie in the analytical estimates for the specific capacity

of a conventional well with no skin. Hence, the deviation from the predicted flow

efficiencies is likely associated with an overestimation of the permeability ratio. Sources

of this overestimation could be due to differences in the idealized form of the fracture

opposed to the actual form in the field [Gidley 1989]. This portion of the theoretical

analysis assumes the fracture to be a dome-shaped feature with a smoothly tapering

aperture. Excavations have revealed that fracture apertures vary by 50% or more

compared to the average aperture [Murdoch and Slack 2002]. This variation can result in

fractures that pinch down to a small thickness, reducing the effective permeability of the

fracture. Increased water content associated with the fracturing process could also reduce

the relative gas permeability of the sand in the fracture or the enveloping formation.

Pressure Distribution

The analytical model was used to estimate the horizontal and vertical

pemeabilities at the field sites based on the pressure distribution in the subsurface and the

discharge from the extraction wells during gas pump tests. Fracture sand permeabilities,

as well as local site permeabilities were obtained with a shelby- tube permeameter

described by (Fig. 17). These values were used as initial inputs for the numerical model

to generate a pressure field for a heterogeneous case with permeabilities similar to


119

conditions in the field. The model was calibrated by trial and error adjustment of the five

permeability values within the model. Pressures observed during gas pumping tests were

then used to compare to theoretical predictions of subsurface pressures. For this

example, observations from a gas pump test performed on the coarse-sand fracture at the

south site were compared to the predicted pressures along vertical axes at specified radial

distances (0.3 m, 0.6 m, 1.5 m, 3.0 m, 4.5 m) away from the injection well (Fig. 45).

The theoretical analysis of the coarse-sand fracture at the south site considers two

distinct layers, with the boundary between the layers at a depth of 2.5 m, corresponding

to row 20 in the finite difference grid. The permeabilities of the upper layer are

kh=1.20 x 10-12 m2 and kv=8.50 x 10-14 m2, and the permeabilities of the lower layer are

kh=1.32 x 10-11 m2 and kv=4.38 x 10-12 m2. The fracture has a permeability of

8.00 x 10-9 m2, with a volume of 0.3 m3 at a depth of 3.0 m. Pressure at the well is set at

0.96 in of water, which is consistent with observations at the well. The model typically

converges in 1500 to 2000 iterations, with a tolerance of 5.0 x 10-6.

The pressures predicted by the numerical model fit the observed pressures from

field experiments moderately well, resulting in a pearson correlation coefficient of

r = 0.870, according to (72). The model predicts a discharge of 28.9 scfm, which is

similar to observations in the field (28.0 scfm). Pressures predicted by the model are

overestimated in the immediate vicinity of the fracture, with the maximum difference of

0.006 atm (2.44 in of water) occurring closest to the extraction well at the depth of the

fracture (3.0 m)(Fig. 45).


120

3
Depth (m)

8
0.980 0.982 0.984 0.986 0.988 0.990 0.992 0.994 0.996 0.998 1.000

Pressure (atm.)

Radial distance
0.3 m Field Observation

0.6 m Predicted from


Numerical solution
1.5 m
3.0 m
4.5 m

Figure 45. Pressures predicted by the numerical model compared to field observations for a
coarse-sand hydraulic fracture at a depth of 3.0 m.
121

a. b.
-1 -1

-2 -2
0.20 0.15
-3 0.30 0.25 0.20 0.15 0.25
-3
0.30
meters

-4 -4

-5 -5

-6 -6

-7
-7

1 2 3 4 5 6 7 1 2 3 4 5 6 7

meters meters

Figure 46. Contours of pressures a. predicted by the numerical model and b. observed during
a gas pumping test performed on the coarse-sand fracture at the south site.

The pressures observed in the field during the gas pumping test performed on the

coarse-sand fracture at the south site were contoured by hand and compared to the

pressures predicted by the numerical model (Fig. 46). Pressures have been normalized to

the pressure at the well in order to make comparisons between the plots more convenient.

Overall, the shape of the contours are similar, with 15% of the pressure applied at the

well having a radial extent of 4.5 m for the numerical model, and 5.0 m for the observed

pressures, both at the depth of the fracture. Pressures in the immediate vicinity of the

fracture closest to the well are overestimated by the numerical model, with the radial

extent of the 0.3 pressure contour occurring 1.3 m from the extraction well, while

observations during the gas pumping test have the 0.3 pressure contour extending 0.5 m

away from the extraction well.


122

Sources of error between the observed and predicted pressures could be due to

local heterogeneities omitted by the simulation. Issues associated with the performance

of the multi-level piezometers used to acquire field data during gas pumping tests, such

as the higher permeability annulus, are also a possible source.

For the purpose of the model, it was assumed that the geology at the field sites

consists of a two layers, with a flat-lying boundary between the layers. Core samples

taken at the field sites reveal an uneven surface representing the boundary between the

upper and lower layers, which can range 0.3 meters above or below the assumed depth of

the boundary (2.5m). Misrepresentation of the boundary between the layers could result

in an inability of the numerical model to account for irregularities in the observed

pressures. Core samples and subsequent excavations at the field sites have also revealed

local heterogeneities, with small-scale variations that are omitted in the theoretical

analysis.

The multi-level piezometers were developed for this study because they provide a

vertical resolution difficult to achieve with conventional wells. However, this method of

installing observation wells is not without its limitations. Correct functioning of the

piezometers require a seal to be created that forces the pads against the borehole wall and

isolates neighboring filter pads. A region of higher permeability could develop in the

annulus between the plastic sheath of the piezometer and the formation. Such a region

would reduce the difference in pressure between measurements at neighboring points

along the piezometer. The pressures predicted by the numerical model show sharp

increases in the suction in the immediate vicinity of the fracture. A region of higher

permeability around the borehole of the piezometer would decrease pressure losses
123

between data points, thus diminishing the observed high pressures associated with air

flow in the fracture.

As discussed in the Flow Efficiency portion of the Evaluation of Field Data

section of this report, the fracture can be represented as a dome-shaped feature with a

tapering aperture. This idealized form has been proven to adequately represent the

overall form of the fracture (Fig. 40), but neglects to account for relatively small-scale

variations in the aperture observed in the field [Murdoch and Slack 2002]. Necking

down of the aperture near the injection point may be responsible for the reducing the

pressures observed in the immediate vicinity of the fracture closest to the well, while

allowing for more accurate predictions of far-field pressures.


124

CONCLUSIONS

Sand-filled hydraulic fractures improve the recovery of contaminants from

environmental wells in fine-grained formations. The performance of SVE wells can be

evaluated by comparing the specific gas capacity of wells, defined as

Sc Q , and the subsurface pressure and flow in the vicinity of wells. The
2 2
Patm Pwell

flow efficiency is a simple way to characterize the performance of a well by normalizing

the observed specific capacity to the specific capacity of a well operating under ideal

conditions. Well performance is evaluated using flow efficiency, and subsurface pressure

and flow in the vicinity of wells, based on the permeability ratio between the fracture and

the formation and geometry of the fracture.

CONVENTIONAL WELLS

Drilling and well completion techniques can create a low permeability skin

enveloping the well screen, which may markedly reduce the specific capacity of wells.

The specific capacities of conventional wells at the field sites range from 2 to 34

scfm/atm2 at the north site, and from 20 to 119 scfm/atm2 at the south site. A skin factor

associated with each well is determined from the analytical solution based on data from

gas pumping tests. Skin factors for conventional wells range from 0.84 to 26 at the north

site, and from 0.54 to 6 at the south site. The ranges of skin factors correspond to

observed specific capacities, in that wells with lowest skin factors result in highest
125

specific capacities, and wells with highest skin factors yield lowest specific capacities.

Flow efficiencies for conventional wells at the north site range from 0.035 to 0.537, and

at the south site from 0.250 to 1.507. Flow efficiencies for conventional wells are based

on the analytical estimate of the specific capacity for a conventional well with no skin.

FRACTURED WELLS

The permeability of the fracture must be approximately 50 times greater than that

of the formation for a 100% increase in the specific gas capacity (F = 2) according to the

numerical model. Permeability ratios estimated from field data at the north site are 360

for the coarse-grained sand fracture, and 200 for the medium-grained sand fracture. The

permeability ratio estimated for the coarse-grained sand fracture at the south site is 480,

and the permeability ratio for the medium-grained sand fracture is 260.

Observed specific gas capacities for wells intersecting fractures show, on average,

an order of magnitude increase compared to the performance of conventional wells at the

field sites. Flow efficiencies at the north site range from 5 to 133 for the medium-grained

fracture, and from 3 to 84 for the coarse-grained fracture. Flow efficiencies at the south

site range from 3 to 12 for the medium-grained fracture, and from 4 to 17 for the coarse-

grained fracture. Variations in flow efficiency depend on the method used to determine

the baseline specific capacity of conventional wells, Sco. Flow efficiency as a function of

the permeability ratio was determined numerically, with Sco based on the specific

capacity of a conventional well operating under ideal conditions.


126

Observed flow efficiencies of wells intersecting fractures at the field sites based

on the performance of a conventional well performing under ideal conditions fall short of

the predicted values. This is likely due to the idealized fracture form used in the

theoretical analysis, which assumes the fracture have an aperture that tapers smoothly,

with uniform sand distribution. Excavations and borings of fractures show that the

apertures can vary as much as 50% over short intervals [Murdoch and Slack 2002], which

can significantly reduce the effective permeability of the fracture. Theoretical analyses

indicate that neither conventional wells, nor wells intersecting fractures behave ideally in

the field. It is important to recognize the factors controlling well performance in order to

improve the performance of both conventional wells and wells intersecting hydraulic

fractures.

Pressure distributions and flow in the vicinity of wells were also used to evaluate

the effects of a fracture intersecting an SVE well by comparing well performance under

varying permeability ratios and fracture geometries. The radius of pressure influence at

the depth of the fracture is increased from approximately 0.3 m to 5.0 m away from the

extraction well by the presence of a hydraulic fracture with a permeability 1000 times

that of the formation with a radius of 4.2 m at a depth of 3.0 m.

The effects of factors related to fracture geometry on well performance were

evaluated using the numerical model with the fracture represented as a disc-shaped

feature of uniform thickness. Fracture aperture has the greatest influence on well

performance, affecting flow efficiencies over the entire range of permeability ratios

greater than 1.0, with thicker fractures producing greater flow efficiencies for a given

permeability ratio. Radius has a moderate effect on well performance, only affecting the
127

flow efficiency for permeability ratios > 200, at which point the well performance for

smaller diameter fractures show less of an improvement. Depth has a negligible effect on

well performance, with the flow efficiency being essentially independent of depth for a

fracture with a radius of 4.2 m and an aperture of 0.005 m.

Analysis based on the mechanics of fracture propagation has shown that the

radius and aperture of a fracture can be estimated as a function of the volume of sand in

the fracture. A relatively small fracture (0.1 m3) at a depth of 3.0 m, and with a

permeability ratio = 1000, will increase the radius of pressure influence by roughly a

factor of 10. Increasing the volume increases the radius of the fracture, which coincides

with the radial extent of pressure influence.

Flow paths are evaluated using the radial location of the 0.5 streamline at the

ground surface. A fracture with a radius of 4.2 m and a permeability ratio = 1000 will

increase the radial distance to the 0.5 streamline from 4.4 m to 5.5 m compared to a

conventional well. This increase in radial distance results in an area that is approximately

65% greater containing 50% of the total flow through the ground surface.

MODEL PERFORMANCE

The analytical model for a conventional well predicts subsurface pressures fields

resulting in correlation coefficients of 0.999 for analysis of individual gas pumping tests,

0.930 for the ensemble analysis of the north site, and 0.889 for the ensemble analysis of

the south site.


128

The numerical model can predict the pressure field in the vicinity of an extraction

well with reasonably accuracy, resulting in a correlation coefficient of 0.870. However,

the model underestimates pressures in the immediate vicinity of the fracture. This

underestimation of pressures may again be due to the idealized fracture form used in the

numerical model, which would give the fracture a greater effective permeability than

observed in the field.

Sensitivity analysis for the north site revealed that permeabilities may vary

between 2.3 x 10-12 m2 and 1.6 x 10-11 m2 in the horizontal direction, and from 9.0 x 10-12

m2 to many orders of magnitude beyond the 1.3 x 10-11 m2 boundary of the plot in the

vertical direction. Sensitivity analysis for the south site revealed that permeabilities

range from 1.0 x 10-11 m2 to 1.8 x 10-11 m2 in the horizontal direction, and from

4.0 x 10-12 m2 to 1.9 x 10-11 m2 in the vertical direction. More accurate estimates of upper

layer and lower layer permeabilities would refine the model calibration, and possibly

result in a better fit.


129

APPENDICES
130

APPENDIX A

NORTH SITE GAS PUMPING TEST OBSERVATIONS

Observations gathered during gas pumping tests on wells not included in the body

of text are listed below. Pressures given in inches of water.

NW-1
Q =0.6 scfm
P @well =46.9 in H 2 0
depth (m) NW-1 NW-2 NW-3 NW-4 NW-5 NW-6 NW-7 NW-8
1.5 0
3 0.04 0.03 0.02 0.02 0.02 0.06

NW-2
Q =0.5 scfm
P @well =48.9 in H 2 0
depth (m) NW-1 NW-2 NW-3 NW-4 NW-5 NW-6 NW-7 NW-8
1.5 0
3 0 0 0 0 0 0

NW-3
Q =7.1 scfm
P @well =44.4 in H 2 0
depth (m) NML-2 NW-1 NW-2 NW-3 NW-4 NW-5 NW-6 NW-7 NW-8
1.5 0.48 1.02
2.4 1.38
3 1.69 1.11 3.03 5.3 1.7 3.58 3.74
3.6 2.28

NW-4
Q =6.1 scfm
P @well =45.0 in H20
depth (m) NML-2 NW-1 NW-2 NW-3 NW-4 NW-5 NW-6 NW-7 NW-8
1.5 0.49 1.13
2.4 1.9
3 1.74 1.28 1.92 4.59 2.37 2.76 2.31
3.6 2.16
131

NW-5
Q =0.4 scfm
P @well =48.5 in H20
depth (m) NML-2 NW-1 NW-2 NW-3 NW-4 NW-5 NW-6 NW-7
1.5 0
2.4 1.25
3 0 0.09 0.1 0.11 0.08 0.07 0.07
3.6 0

NW-6
Q =0.5 scfm
P @well =47.4 in H20
depth (m) NML-2 NW-1 NW-2 NW-3 NW-4 NW-5 NW-6 NW-7 NW-8
1.5 0 0
2.4 0
3 0 0 0 0 0 0 0
3.6 0

NW-7
Q =0.5 scfm
P @well =47.4 in H20
depth (m) NML-2 NW-1 NW-2 NW-3 NW-4 NW-5 NW-6 NW-7 NW-8
1.5 0.04 0.03
2.4 0.98
3 0.13 0.07 0.17 0.26 0.23 0.24 0.18
3.6 0.19

NW-8
Q =1.9 scfm
P @well =46.1 in H20
depth (m) NML-2 NW-1 NW-2 NW-3 NW-4 NW-5 NW-6 NW-7 NW-8
1.5 0.06 0.18
2.4 1.16
3 0.27 0.69 0.47 0.71 0.49 0.32 0.39
3.6 0.41

NW-9
Q =5.9 scfm
P @well =45.6 in H20
depth (m) NML-2 NW-1 NW-2 NW-3 NW-4 NW-5 NW-6 NW-7 NW-8
1.5 N/D N/D
2.4 N/D
3 N/D N/D N/D N/D N/D N/D N/D N/D
3.6 N/D
132

NW-10
Q =0.6 scfm
P @well =45.8 in H20
depth (m) NML-2 NW-1 NW-2 NW-3 NW-4 NW-5 NW-6 NW-7 NW-8
1.5 N/D N/D
2.4 N/D
3 N/D N/D N/D N/D N/D N/D N/D N/D
3.6 N/D

Medium Fracture, North Site

Q =26.5 scfm P @well =17.5 in H 2 0


depth (m) NML-1'N NW-2.5'N NML-4'N NW-5'N
2.4 6.29 4.09
2.7 6.32 4.22
3 6.34 3.95 4 3.87
3.3 5.14 1.95
3.6 5.07 3.92
3.9 4.89 3.75
4.2 4.87 3.6

Coarse Fracture, North Site

Q =23.5 scfm P @well =24.8 in H 2 0


depth (m) NW-3.5'N NW-5'N
1
1.3 3.31
1.6 0
1.9
133

APPENDIX B

SOUTH SITE GAS PUMPING TEST OBSERVATIONS

Observations gathered during gas pumping tests on wells not included in the body

of text are listed below. Pressures given in inches of water.

SW-1
Q =3.9 scfm
P @well =43.5 in H20
depth (m) SW-1 SW-2 SW-3 SW-4 SW-5
1.5 0.32 0.42
3 1.35 0.78

SW-2
Q =12.5 scfm
P @well =23.2 in H20
depth (m) SW-1 SW-2 SW-3 SW-4 SW-5 SML-1
1.5 0.96 1.28
2.4 3.26
3 4.67 5.9 3.71
3.6 4.23

SW-3
Q =14.5 scfm
P @well =25.7 in H20
depth (m) SW-1 SW-2 SW-3 SW-4 SW-5 SML-1
1.5 N/D N/D
2.4 N/D
3 N/D N/D N/D
3.6 N/D

SW-6
Q =4 scfm
P @well =39.6 in H20
depth (m) SW-1 SW-2 SW-3 SW-4 SW-5 SML-1
1.5 N/D N/D
2.4 N/D
3 N/D N/D N/D
3.6 N/D
134

SW-7
Q =7.5 scfm
P @well =39.1 in H20
depth (m) SW-1 SW-2 SW-3 SW-4 SW-5 SML-1
1.5 0.61 0.73
2.4 1.63
3 1.31 1.05 0.99 1.71
3.6 1.16

Medium Fracture, South Site

Q =25.0 scfm P @well =20.1 in H 2 0


depth (m) SML-3'W
2.1 0.96
2.4 4.66
2.7 5.75
3 8.29
3.3 5.74
3.6 4.13
3.9 2.86
4.2 2.48
135

APPENDIX C

ANALYSIS OF GAS PUMPING TESTS FOR THE NORTH SITE

Observed and predicted absolute pressures from individual analysis not included

in the body of text. Pearson correlation coefficients also given below.

NW-1

101326
101324
101322 r = 0.999
Pressure (Pa)

101320
101318
101316
101314 observed
101312
predicted
101310
101308
0 0.2 0.4 0.6 0.8 1 1.2
radius (m)

NW-3

101200
101000
100800
Pressure (Pa)

r = 0.999
100600
100400
100200
observed
100000
predicted
99800
0 0.1 0.2 0.3 0.4 0.5 0.6
radius (m)
136

NW-4

101200

101000
r = 0.999

Pressure (Pa)
100800
100600

100400
observed
100200
predicted
100000
0 0.2 0.4 0.6 0.8
radius (m)

NW-5

101310
101308
Pressure (Pa)

101306
r = 0.999
101304
101302
101300
101298 observed
101296 predicted
0 0.2 0.4 0.6 0.8
radius (m)

NW-6

101326

101325
r = 0.997
Pressure (Pa)

101324 observed
predicted
101323

101322
0 0.2 0.4 0.6 0.8 1 1.2
radius (m)
137

NW-7

101330
r = 0.999
101320
101310

Pressure (Pa)
101300
observed
101290
predicted
101280
101270
101260
101250
0 0.2 0.4 0.6 0.8 1
radius (m)

NW-8

101300
101280
101260
r = 0.999
Pressure (Pa)

101240
101220 observed
101200 predicted
101180
101160
101140
101120
0 0.2 0.4 0.6 0.8
radius (m)
138

APPENDIX D

ANALYSIS OF GAS PUMPING TESTS FOR THE SOUTH SITE

Observed and predicted absolute pressures from individual analysis not included

in the body of text. Pearson correlation coefficients are also given below.

SW-1

101350
101300
101250 r = 0.999
Pressure (Pa)

101200
101150
101100
101050
101000 observed
100950 predicted
0 0.2 0.4 0.6 0.8 1
radius (m)

SW-2

101200
101000
100800 r = 0.997
Pressure (Pa)

100600
100400
100200
100000
observed
99800
predicted
99600
0 0.1 0.2 0.3 0.4 0.5 0.6
radius (m)
139

SW-7

101400

101200
r = 0.999

Pressure (Pa)
101000

100800
100600
100400 observed
predicted
100200
0 0.5 1 1.5 2
radius (m)
140

REFERENCES

Anderson, M. P. and W. W. Woessner. 1992. Applied Groundwater Modeling:


Simulation of Flow and Advective Transport. Academic Press.

Arps, J. J. 1955. How Well Completion Damage can be Determined Graphically. World
Oil: Production Section, April, 1955: 225-232.

Earlougher, R. C., Jr. 1977. Advances in Well Testing. American Institute of Mining.
p. 8-10, 42-44.

Edwards, K. B. and Jones, L. C. 1994. Air Permeability from Pneumatic Tests in


Oxidized Till. Journal of Environmental Engineering, v. 120, no. 2.

Falta, R. W., Pruess, K., and Chesnut, D. A. 1993. Modeling Advective Contaminant
Transport During Soil Vapor Extraction. Ground Water, v. 31, no. 6: 1011-1020.

Fan, C. Underground Storage Tank Correction Action: Application and Field Evaluation
of Vacuum Extraction Technology. United States Environmental Protection
Agency. p. 1-4.

Faure, G. 1998. Principles and Applications of Geochemistry: Second Edition. Prentice


Hall. p. 354-356.

Frank, U. and Barkley, N. 1994. Remediation of low permeability subsurface formations


by fracturing enhancement of soil vapor extraction. Journal of Hazardous
Materials, v. 40: 191-201.

Gidley, J. L., Holditch, S. A., Nierode, D. E., and Veatch, R. W. 1989. American
Institute of Mining. p. 1-33, 57-80.

Gierke, J. S. 2000. Remediation Technologies: Conventional Vapor Extraction; in


Vadose Zone Science and Technologies. edited by Looney, B. B. and Falta, R.
W. Batelle Press. v. 2: 951-969.
141

Hatcher, R. D., Jr., Thomas, W. A., and Viele, G. W. 1989. The Appalachian-Ouachita
Orogen in the United States: Geology of North America. Geological Society of
America. v. F-2: 1-7, 385-417.

Hartsock, J. H. and Warren, J. E. 1961. The Effect of Horizontal Hydraulic Fracturing


on Well Performance. Journal of Petroleum Technology. p. 1050-1056.

Hubbert, M. K. and Willis, D. G. 1957. Mechanics of Hydraulic Fracturing. Petroleum


Transactions, v. 210: 161-163.

Hutzler, N. J. 1987. Assessment of Full-Scale Vapor Extraction Systems. Proposal to


United States Environmental Protection Agency. p. 1-41.

Huyakorn, P. S. and Pinder, G. F. 1983. Computational Methods in Subsurface Flow.


Academic Press. p. 341-353.

Johnson, P. C., Kemblowski, M. W., and Colthart, J. D. 1990. Quantitative Analysis for
the Cleanup of Hydrocarbon-Contaminated Soils by In-Situ Soil Venting.
Groundwater, v. 28, no. 3: 413-429.

Jones, I., Lerner, D. N., and Baines, O. P. 1999. Multiport Sock Samplers; a low cost
technology for effective multilevel ground water sampling. Groundwater
Monitoring and Remediation, Winter 1999, p. 134-142.

Klinkenberg, L. J. 1941. The Permeability of Porous Media to Liquids and Gases. API
Drilling and Production Practice. p. 200-212.

Liskowitz, J., Schuring, J., Mack, J. 1994. Application of Pneumatic Fracturing


Extraction for the Effective Removal of Volatile Organic Compounds in Low
Permeable Formations. Focus Conference on Eastern Groundwater Issues. p. 1-7.

Massman, J. W. 1989. Applying Groundwater Flow Models in Vapor Extraction System


Design. Journal of Environmental Engineering, v. 115, no. 1: 129-149.

Murdoch, L. C. 2002. Analysis of an Idealized Shallow Hydraulic Fracture. Journal of


Geotechnical and Geoenvironmental Engineering. p. 1-36.

Murdoch, L. C. 1995. Hydraulic and Implulse Fracturing for Low Permeability Soils ;
in Petreleum Contaminated Low Permeability Soil: Hydrocarbon Distribution
142

Processes, Exposure Pathways, and In-Situ Remediation Technologies. edited by


Walden, T. American Petroleum Institute. p. C-1 - C-32, E-1 - E-34.

Murdoch, L. C. 1992. Hydraulic fracturing of soil during laboratory experiments, Parts


1-3. Geotechnique, v. 43, no. 2: 255-287.

Murdoch, L. C. Induced Fractures. United States Environmental Protection Agency. p.


39-55.

Murdoch, L. C. and Slack, W. 2002. Forms of Hydraulic Fractures in Shallow Fine-


Grained Formations. Journal of Geotechnical and Geoenvironmental
Engineering. p. 1-44.

Nelson, A. E., Horton, J. W., and Clarke, J. W. 1990. Geologic Map of the Greenville
1ºx2º Quadrangle, South Carolina, Georgia, and North Carolina. United States
Geological Survey. p. 1-3.

Pederson, T. A. and Fan, C. 1992. Soil Vapor Extraction Technology Development


Status and Trends, Proceedings of the Symposium on Soil Venting. United
States Environmental Protection Agency. p. 1-7.

Perkins, T. K. and Kern, L. R. 1963. Widths of Hydraulic Fractures. Journal of


Petroleum Technology. p. 939-949.

Pollock, D. W. 1988. Semianalytical Computation of Paths Lines for Finite-Difference


Models. Ground Water, v. 26, no. 6: 743-750.

Prothero, D.R. 1977. Interpreting the Stratigraphic Record. W.H. Freeman and
Company. p. 43-102.

Schuring, J. R., Chan, P. C. 1992. Removal of Contaminants from the Vadose Zone by
Pneumatic Fracturing. New Jersey Institute of Technology. p. 31-34, 53-54.

Shan, C., Falta, R. W., and Javandel, I. 1992. Analytical Solutions for Steady-State Gas
Flow to a Soil Vapor Extraction Well. Water Resources Research, v. 28, no. 4:
1105-1120.
143

Steel, R. G. D. and Torrie, J. H. 1960. Principles and Procedures of Statistics with


Special Reference to the Biological Sciences. McGraw-Hill Book Co. p. 188-
190, 451, 557.

Streltsova, T. D. 1988. Well Testing in Heterogeneous Formations. John Wiley and


Sons. P. 21, 158-160.

Van Everdingen, A. F. 1953. The Skin Effect and its Influence on the Productive
Capacity of a Well. Petroleum Transactions, v. 198: 171-176.

Vesselinov, V. V. and Neuman, S. P. 2001. Numerical Inverse Interpretation of Single-


Hole Pneumatic Tests in Unsaturated Fractured Tuff. Groundwater, v. 39, no. 5:
685-695.

Wang, H. F. and Anderson, W. P. 1982. Introduction to Groundwater Modeling: Finite


Difference and Finite Element Methods. Academic Press. p. 26-28.

Wolf, A. and Murdoch, L. C. 1993. A Field Test of the Effect of Sand-Filled Hydraulic
Fractures on Air Flow in Silty Clay Till. As Appears in the Proceedings from the
Seventh National Outdoor Action Conference.
144

Potrebbero piacerti anche