PB91-181818
                               EPA/600/2-91/012
                               April 1991
FEASIBILITY OF HYDRAULIC FRACTURING OF SOIL

        TO IMPROVE REMEDIAL ACTIONS
                       by
          L C. Murdoch, G. Losonsky, P. Quxton,
            B. Patterson, I. JOich, B. BrasweU
              Center Hill Research Facility
                University of Cincinnati
                Cincinnati, Ohio 45224
                 Contract 68-03-3379
                Work Assignment No. 8
                   Project Officer
                 Michael H. Roulier
                              . Laboratory
                CintinnatCOhio 45268
    RISK REDUCTION ENGINNERING LABORATORY
      OFFICE OF RESEARCH AND DEVELOPMENT
     U.S. ENVIRONMENTAL PROTECTION AGENCY
              CINCINNATI, OHIO 45268

-------
                                    TECHNICAL REPORT DATA     .  .
                             (Please read Instructions on the reverse before complttr
 1. REPORT NO.
   EPA/600/2-91/012
                              2.
 4. TITLE AND SUBTITLE
      Feasibility of Hydraulic Fracturing of Soil .  to
  Improve Remedial Actions
                                                           3.
                                                                 PB91-181818
                                                          5. REPORT DATE
                                                           April  1991
                                                          6. PERFORMING ORGANIZATION CODE
7. AUTHOH(S)

  L.C. Murdoch,  G.  Losohsky, P. Cluxton,  B.  Patterson,.
  T  nirh  anH  R  RracwolT
                                                           8. PERFORMING ORG/
                                                                               ION REPOI
                                                           10. PROGRAM ELEMENT NO.

                                                                .1Y1J
I. PERFORMING ORGANIZATION NAME AND ADDRESS

 University  of Cincinnati
 Center Hill  Research Facility
 5995 Center  Hill  Road
 rinrinnati-   OH
 TF.IYIfl
.CONTRAC
                                                           11. CONTRACT/GRANT NO.
                                                              68-03-3379 •
12. SPONSORING AGENCY NAME AND ADDRESS

  Risk Reduction  Engineering Laboratory
  Office of Research  and Development
  U.S. Environmental  Protection Agency
  P-inrinnati. OH—4S?fiR
                                                          13. TYPE OF REPORT AND PERIOD COVERED

                                                            fnmnl Pt.P	
                                                          14. SPONSORING AGENCY CODE
 5. SUPPLEMENTARY NOTES
     Project Officer:
                                                               FPA/finn/11
                      Michael  H.  Roulier  (513) 569-7796  or  FTS 684-7796
 6. ABSTRACT
       Hydraulic fracturing, a method of increasing fluid flow within the subsurface,
  should improve the  effectiveness of several remedial techniques,  including pump
  and treat, vapor extraction, bio-remediation, and soil-flushing.   The technique is
  widely used to increase the yields of oil wells, but is untested  under conditions
  typical of contaminated sites.
       The project consisted of laboratory experiments, where hydraulic fractures
  were created in a triaxial pressure cell, and two field tests, where fractures were
  created at shallow  depths in soil.  The lab tests showed that hydraulic fractures
  are readily created in  clayey silt, even when it is saturated and loosely-consolidated
  Many of the lab observations can be explained using parameters and analyses based
  on  linear elastic fracture mechanics.
       Following the  field tests, the vicinity of the boreholes was excavated to reveal
  details of the hydraulic fractures.  Maximum lengths of the fractures, as measured
  from the borehold to  the leading edge, averaged 4.0 m, and the average area was 19 m?.
  Maximum thickness of  sand ranged from 2 to 20 mm, averaging 11 mm.   As many as four
  fractures were created  from a single borehold, stacked one over the other at vertical
  spacing of 15 to 30 cm.
                                KEY WORDS AND DOCUMENT ANALYSIS
                  DESCRIPTORS
                                              b.lDENTIFIERS/OPEN ENDED TERMS  C.  COSATI Fidd/GlOUp
 8. DISTRIBUTION STATEMENT


 RELEASE TO  PUBLIC
                                             19. SECURITY CLASS (Tint Rtport)
                                               UNCLASSIFIED
            21. NO. OF PAGES
                307
                                             20. SECURITY CLASS (This poftj

                                               UNCLASSIFIED
                                                                        22. PRICE
EPA Form 2220-1 (R«v. 4-77)   PREVIOUS EDITION is OBSOLETE
                                             1

-------
                                NOTICE

      The information in this document has been funded by the United States
Environmental Protection Agency under Contract 68-03-3379 to the University of
Cincinnati. It has been subjected to the Agency's peer and administrative review,
and it has been approved for publication as an EPA document. Mention of trade
names or commercial products does not constitute endorsement or recommendation
for use.
                                 11

-------
                                FOREWORD

      Today's rapidly developing and changing technologies and industrial
products and practices frequently cany with them the increased generation of
materials that, if improperly dealt with, can threaten both public health and the
environment. The U.S. Environmental Protection Agency is charged by Congress
with protecting the Nation's land, air and water resources. Under a mandate of
national environmental laws, the Agency strives to formulate and implement actions
leading to a compatible balance between human activities and the ability of natural
systems to support and nurture life. These laws direct the EPA to perform research
to define our environmental problems, measure the impacts, and search for
solutions.

      The Risk Reduction Engineering  Laboratory is responsible for planning,
implementing, and managing of research, development, and demonstration
programs to provide an authoritative, defensible engineering basis in support of the
policies, programs, and regulations of the EPA with respect to drinking water,
wastewater, pesticides, toxic substances, solid and hazardous wastes, and Superfund-
related activities.  This publication  is one of the products of that research and
provides a vital communication link between the researcher and  the user
community.

      This study was undertaken to determine if hydraulic fracturing, a method of
increasing fluid flow in subsurface soils, could improve the effectiveness of several
remedial action techniques at sites  contaminated by hazardous wastes.

                         E. Timothy Oppelt, Director

                         Risk Reduction Engineering Laboratory
                                  in

-------
                                 ABSTRACT

      Hydraulic fracturing, a method of increasing fluid flow within the subsurface,
should improve the effectiveness of several remedial techniques, including pump
and treat, vapor extraction, bio-remediation, and soil-flushing. The technique is
widely used to increase the yields of oil wells, but it is untested under conditions
typical of contaminated sites.

      The project consisted of laboratory experiments, where hydraulic fractures
were created in a triaxial pressure cell, and two field tests, where fractures were
created at shallow depths in soil. The laboratory tests showed that hydraulic
fractures are readily created in clayey silt, even when it was saturated and loosely-
consolidated. Many of the laboratory observations can be explained using
parameters and analyses based on linear elastic fracture mechanics.

      Field tests were conducted during the summers of 1988 and 1989 at sites
underlain by Pleistocene glacial till. During the 1988 test, hydraulic fractures were
successfully created from cemented casing at depths of 2 to 4 m using equipment
designed for fracturing oil wells. The tests were limited to creating one fracture per
well, and they were hindered by the large oil-field equipment, which was difficult to
control because it was designed  to operate under flow rates and pressures much
greater than those-used during the test. Many .of the fractures created in 1988
closed completely because they were barren of proppant, a condition that
apparently resulted from insufficient control.

      For the 1989 test, a new method of setting casing was designed, and pumps
and blenders designed for injection grouting were used. Casing was set by inserting
a drill rod tipped with a conical point and driving the assembly into the ground. The
rod and point were retracted and a water jet was inserted to cut a disk-shaped notch
beneath the casing. A hydraulic fracture was nucleated at the notch by injecting gel
and sand into the casing.  Following fracturing, the rod and point were inserted and
the assembly was driven to a greater depth, where the process was repeated.

      Following the tests, the vicinity of the boreholes was excavated to reveal
details of the hydraulic fractures. In general, they were slightly elongate (aspect of
2:3) in plan view, and they were highly asymmetric with respect to their parent
borehole; in each case there was a preferred direction of propagation.  The fractures
created  in 1988 climbed gently and vented to the ground surface, whereas the ones
created  in 1989 were nearly flat-lying, and few of them vented.  Maximum lengths of
the fractures in 1989, as measured from the borehole to the leading edge, averaged
4.0 m, and the average area was 19 m2. Maximum thickness of sand in individual
fractures ranged from 2 to 20 mm, averaging 11 mm. As many as four fractures
were created from a single borehole, stacked one over the other at vertical spacings
of 15 to 30 cm.

      Four monitoring methods were evaluated: injection pressure, surface uplift,
surface tilt, and electrical resistivity. Results indicate that it should be feasible to
monitor the growth of hydraulic fractures at shallow depths.

      It is feasible to create hydraulic fractures at shallow depths in unlithified
sediment, and we recommend that this technology be tested during the remediation
of contaminated sites.

-------
                          CONTENTS
Foreword
Abstract
Figures
Tables
Acknowledgment
Introduction	1
           Statement of the problem	2
           Approach	3
           Major Conclusions	4
           Laboratory Studies: Summary of Results	4
           Field Studies: Summary of Results	6
           Recommendations	8

1: Potential Remedial Applications and Previous Investigations of
               Hydraulic Fracturing	6
           Assessment of Potential Applications	6
           Previous Studies of Hydraulic Fracturing	18

2: Observations During Laboratory Experiments	33
           Experimental Design	33
           Appearance of a Hydraulic Fracture in Center Hill Clay	43
           Records of Driving Pressure	54
           Summary and Discussion	63

3: Analysis of Hydraulic Fracturing of Soil	67
           Predicting the Onset of Hydraulic Fracturing	68
           Propagation of a Hydraulic Fracture in Soil	75
           Data Quality	116
           Discussion	120

4: Setting and Design of the Field Test -1988	122
           Site Characteristics	122
           Boreholes	132
           Method of Fracturing	137

5: Hydraulic Fractures Created During the Field Test	141
           Forms of the Hydraulic Fractures	142
           Dimensions of Hydraulic Fractures	160
           Direction of Propagation	161
           Summary: An Idealized Hydraulic Fracture	163
           Discussion: Development of the Idealized Fracture	165

6: Setting and Design of the Field Test - 1989	170
           Quality Assurance and Control	170
           Site Characteristics	170
           Method of Fracturing	172

-------
7: Hydraulic Fractures Created During the Field Test - 1989	181
            Quality Assurance and Control	181
            Forms of the Hydraulic Fractures	181
            Dimensions of the Hydraulic Fractures	204
            Thickness Profiles of the Hydraulic Fractures	205
            Direction of Propagation	205
            Discussion	205

8: Monitoring Hydraulic Fractures	214
            Injection Pressure	214
            Surface Tilt	229
            Surface Uplift	239
            Electrical Resistivity	.243
            Discussion	257


9: Summary and Conclusions	258
            Potential Applications	258
            Laboratory Studies: Summary of Results	259
            Field Studies: Summary of Results	261

References	265
Appendix: Records of Pressure, Volume, Surface Displacement, and
               Surface Tilt From 1988 Field Tests	282
                                   VI

-------
                               FIGURES
Number                                                                     Page

  1.1.    Production rate of an oil well in North Texas as a function of time	9
  1.2.    Rates of inflow into boreholes in till	11
  1.3.    Recovery capacity of a flat-lying, circular fracture in a confined
                 aquifer	17
  2.1.    Apparatus used for hydraulic fracturing experiments	35
  2.2.    Cut-away sketch of hydraulic fracturing cell	36
  2.3.    Grain-size distribution of soils used in the laboratory experiments	39
  2.4.    Bulk density as a function of moisture content from Proctor tests and
                 fracture tests	40
  2.5.    Photograph and sketch of the surface of a hydraulic fracture	44
  2.6.    Sketches of surfaces of three hydraulic fractures of various lengths	46
  2.7.    Idealized diagram of the leading edge and propagation paths	47
  2.8.    Idealized configurations of commonly-occurring fracture lobes	49
  2.9.    Leading edge of a hydraulic fracture	52
  2.10.   Length of undyed zone as functions of dyed length and moisture
                 content	53
  2.11.   Records of driving pressure as a function of time	55
  2.12.   Idealized record of driving pressure as a function of time	58
  2.13.   Records of injection pressure as a function of time for samples of
                 various moisture contents	60
  2.14.   Records of injection pressure as a function of time using various
                 lengths of starter slots	66
  2.15.   Idealized cross-section of a hydraulic fracture in a laboratory
                 sample	70
  3.1.    Critical  stress intensity as a function of half-length of starter slot	73
  3.2.    Critical  stress intensity as functions of moisture content and duration
                 of  consolidation	.'.	77
  3.3.    Conceptual model of growth of an idealized hydraulic fracture in
                 laboratory samples	80
  3.4.    Geometry used in analyses	84
  3.5.    Velocities of fluid within a fracture and of the fracture tip	91
  3.6.    Normalized length of unwetted tip with respect to length of etted
                 fracture	93
  3.7.    Normalized driving pressure, aperture, and half-length as functions
                 of time predicted by analyses	95
  3.7a.   Average propagation velocities from laboratory experiments	96
  3.8.    Driving  pressure and apertures as functions of length, during
                 inflation and propagation. The x symbol marks the end of
                 the fracture	98
  3.9.    Loading conditions used to develop analytical model	102
  3.10.   The effect of m on dimensionless driving pressure, half-length and
                 aperture	104
                                    vu

-------
Number                                                                    Page

  3. IS.   Records of driving pressure as a function of time from experiments
                and from analytical solution for samples of various moisture
                contents	110
  3.16.   Records of driving pressure as a function of time from experiments
                and from analytical solution for samples that differ only in
                the length of their starter slot	Ill
  3.17.   The ratio of wetted to total length of a fracture as a function of
                confining stress. Data from Medlin and Masse (1984)	117
  3.18.   Variation in exponents in eq. (3.40) as functions of confining
                stress	118
  4.1.    Topography and locations of boreholes, vents, and hydraulic
                fracturing equipment at the ELDA test site	123
  4.2.    Geology of the ELDA test site		124
  4.3.    Stratigraphic section from the vicinity of the ELDA test site	125
  4.4.    Method of creating small hydraulic fracture to measure the least
                horizontalconfining stress	128
  4.5.    Injection pressure as a function of time during laboratory
                experiments (a), and using field apparatus (b)	130
  4.6.    Grain size distributions in samples of ELDA Till	133
  4.7.    A borehole used to create hydraulic fractures	136
  5.1.    Hydraulic fractures HF5, HF6, and HF7. Structural contours are on
                the fracture surfaces	144
  5.2.    Trace of hydraulic fractures HF5 and HF6	146
  5.3     Trace of hydraulic fracture HF6	147
  5.4     Trace of hydraulic fracture HF7	148
  5.5.    Hydraulic fracture HF9	150
  5.6.    Trace of hydraulic fracture HF9	151
  5.7.    Hydraulic fractures HF10 and HF11	152
  5.8.    Hydraulic fracture HF12	154
  5.9.    Geology and trace of hydraulic fracture HF12	155
  5.10.   Hydraulic fracture HF13	157
  5.11.   Outline of HF13 showing area where  fracture cuts a bed of
                upwardly-grading gravel, sand and  silt	158
  5.12.   Trace of hydraulic fracture HF13	159
  5.13.   Outlines of hydraulic fractures and locations of a backhoe at the
                time of fracturing	.,	162
  5.14.   Idealized hydraulic fracture created at the ELDA test site	166
  6.1.    Scheme for hydraulic fracturing operation performed during the
                1989 field test	173
  6.2.    Fracturing lance, used to prepare boreholes for hydraulic fracturing
                during the  1989 field test	177
  6.3.    Five steps of hydraulic fracturing	178
  7.1.    Map of ELDA landfill site	182
  7.2.    Fracture map of trenches B and C	184
  7.3.    Fracture map of trenches D, E and F	185
  7.4.    Fracture map of trenches G, H and 1	186
  7.5.    Cross section B-B'	188
  7.6.    Cross section C-C	188
  7.7.    Cross section D-D'	189
                                    vni

-------
Number                                                                   £gg£

  7.8.    Cross section E-E'	189
  7.9.    Cross section F-F	189
  7.10.   Cross section G-G'	„	-	190
  7.11.   Cross section H-H'	190
  7.12.   Map of east wall of trench B	191
  7.13.   Map of east wall of trench C			192
  7.14.   Map of east wall of trench D	193
  7.15.   Map of east wall of trench E	194
  7.16.   Map of east wall of trench F	195
  7.17.   Map of east wall of trench G	-	196
  7.18.   Map of east wall of trench H	197
  7.19.   Map of north wall of trench 1	198
  7.20.   Fracture thickness for EL6F1, EL6F2 and EL6F3	.206
  7.21.   Fracture thickness for EL7F1 and EL7F2	20?
  7.22.   Fracture thickness for EL3F1 and EL3F2	208
  7.23.   Fracture thickness for EL1F1 and EL1F2	209
  7.24.   Fracture thickness for EL2F1 and EL2F2	210
  7.25.   Fracture thickness for EL4F1, EIAF2, E1>*F3 and EL4F4	211
  7.26.   Fracture thickness for EL5F1, EL5F2 and EL5F3	212
  8.1.    Pressure records from fracturing tests during 1989	215
  8.2.    Tiltmeter array from 1989 tests	231
  8J.    Tilt as a function of time			—232
  8.4.    Hydraulic fracture and surface tilt interpretation of HF5	235
  8.5.    Hydraulic fracture and surface tilt interpretation of HF6	237
  8.6.    Hydraulic fracture and surface tilt interpretation of HF7	238
  8.7.    Thickness and uplift contours for EL6F2 and EL6F3	240
  8.8.    Thickness and uplift contours for EL7F1	.241
  8.9.    Thickness and uplift contours for EL7F2	244
  8.10.   Thickness and uplift contours for EL5F1, EL5F2, EL5F3
                andL5F4	245
  8.11.   Apparent resistivity contour map on borehole 4, following fracture
                H4	248
  8.12.   Apparent resistivity contour map on borehole 6, following fracture
                H4	249
  8.13.   Apparent resistivity contour map on borehole 5, following fracture
                H4andH5	250
  8.14.   Apparent resistivity contour-map on borehole 4, following fracture
                H5	251
  8.15.   Apparent resistivity contour map on borehole 4, following fractures
                H4,H5andH7	„	252
  8.16.   Apparent resistivity contour map on borehole 7, following fractures
                H4,H5,andH7	253
  8.17.   Apparent resistivity contour map on borehole 4, following fractures
                H6	254
  8.18.   Apparent resistivity contour map on borehole 6, following fracture
                H6	255
  8.19.   Summary of apparent resistivity changes relative to borehole 4
                resulting from H4, H5, H6 and H7	256
                                   IX

-------
                                TABLES
Number                                                                    £ag£


  1.1.    Production rates before and after hydraulic fracture	7
  1.2.    Inflow rates into glacial till	13
  1.3.    Some possible constants in Eq. (1.9)	27
  2.1.    Characteristics of soils used in the study.	-	38
  3.1.    Parameters used in numerical analyses	-	89
  3.2.    Parameters used in analyses shown in figures	112
  3.3.    Exponents from various analytical solutions	114
  3.4.    Comparison of constants	115
  3.5.    Accuracy of laboratory measurements	119
  4.1.    Conditions of boreholes prior to testing	127
  4.2.    Saturated hydraulic conductivities of till	~	131
  43.    Physical characteristics of silty-clay till	132
  4.4.    Summary of data from field tests	139
  5.1.    Dimensions and dips of hydraulic fractures	160
  5.2.    Azimuths of features of hydraulic fractures	163
  7.1.    Fracture size	183
  7.2.    Fracture thickness	200
  7.3.    Proppant concentration	202
  7.4.    Flow parameters	203
  8.1.    Actual and predicted orientations of fractures	234

-------
                        ACKNOWLEDGMENTS

      The efforts of Doug Ammon, formerly of the USEPA, who initially
recognized the potential of hydraulic fracturing, and the continued support of
Herb Pahren and Mike Rouher of the USEPA have been essential to the
success of this project. Reviews by Herb Pahren, Robert Hartley and Ken
Dotson improved the quality of this report. John Stark, manager of the ELD A
Landfill, gave us permission to conduct the 1988 and 1989 field tests on
property owned by ELDA. Tom Busek, president of the Goettle Construction
Company, gave us permission to conduct tests during 1989 on property owned
by Goettle Construction. Don Steirman, professor of geophysics at the
Univerisity of Toledo, and his assistants conducted the resistivity surveys during
the 1988 field tests. Gary Holzhausen and Howard Egan of Applied
Geomechanics, Santa Cruz, CA, supplied us with tiltmeters and they analyzed
the tilt signals obtained during the 1988 field tests. Mark Roberts and his field
crew from Halliburton Services created the hydraulic fractures during the 1988
tests.

      Space limitations prevent us from acknowledging individually a number
of colleagues and associates who contributed to this project. Their efforts have
been greatly appreciated.
                                  XI

-------
                              INTRODUCTION

      The recovery of hazardous chemicals from contaminated ground is often
difficult and sometimes impossible using established techniques, so earth scientists
have begun to turn to related fields for innovative ideas. In petroleum engineering,
the problem of recovering hydrocarbons from reservoirs is closely analogous to the
problem of recovering contaminants from aquifers. A wide range of techniques has
been developed to enhance the recovery of oil from reservoirs, and one of the most
effective is hydraulic fracturing.  This research was motivated by the possibility of
using hydraulic fractures to improve the remediation of contaminated ground.

      The basic process of hydraulic fracturing, as it is used in the petroleum
industry, begins with the injection of fluid into a well until the pressure of the fluid
exceeds a critical value and a fracture is nucleated. A granular material, which is
usually sand and is termed  a proppant, is pumped into the fracture as it grows away
from the well. Transport of the proppant is facilitated by using a viscous fluid,
usually a gel formed from guar gum and water,  to carry the proppant grains into the
fracture. After pumping, proppant holds the fracture open while the viscous gel
breaks down into a thin fluid.  The thinned gel is then pumped out of the fracture,
creating a permeable channelway suitable for either the delivery or recovery of
liquid or vapor. As a result of hydraulic fracturing the flow rates of oil wells
commonly increase by factors of 1.5 to 8 (e.g. Howard and Fast, 1970). In some
cases where oil or gas is present in an adjacent formation but initially undetectable
in a well, hydraulic fracturing can dramatically increase the yield and make the well
an economic producer, according to Howard and Fast (1970).


STATEMENT OF THE PROBLEM

      Experience from oil wells suggests that hydraulic fracturing could increase
flow rates from wells used to recover ground water contaminants. However, to
realize this increase hydraulic fractures would have to be created and filled with
sand under conditions of contaminated regions. Oil reservoirs are typically deeper
and are composed of different materials than contaminated regions, so the
applicability of fracturing methods used by petroleum engineers is unknown.
Contaminants commonly occur in soils1 that are weaker and more compliant than
limestone or sandstone typical of reservoirs. Effects of soil properties on hydraulic
fractures are difficult to anticipate based on the results of previous studies of
hydraulic fractures in rock. Moreover, most contaminants occur at shallow depths
(several meters  to several tens of meters), so intersecting the ground surface and
venting could severely limit the length, and thus the performance of the fracture.
Hydraulic fractures virtually never vent when they are created in oil reservoirs,
which are several hundred  to several thousand meters deep, so the practical
problem of creating fractures at shallow depths has yet to be addressed.

      Previous investigations of hydraulic fracturing of soil have focused on
methods of preventing fracture initiation, largely ignoring the problems cited above.
As a result, little is known about both the basic problem of the mechanical behavior
1 The termso// is used in this work to mean an unlithified, or uncemented sediment

-------
of hydraulic fracture propagation in soil, and the applied problem of creating useful,
proppant-filled hydraulic fractures at shallow depths.
APPROACH

      The approach of this research was to adapt methods proven for hydraulic
fracturing of rock to applications of hydraulic fracturing of soil.  Laboratory
experiments were conducted by creating hydraulic fractures in rectangular samples
of remolded clayey-silt confined in a triaxial pressure cell. Results of those
experiments were analyzed using methods of linear elastic fracture mechanics, a
branch of elasticity theory that is widely used to analyze hydraulic fractures in rock.

      Two sets of field experiments were performed by creating hydraulic fractures
in Pleistocene glacial drift at depths of between two and four meters. The first set
was conducted during June 1988 in collaboration with a subcontractor, who used
equipment designed to create hydraulic fractures from oil wells; whereas the second
set was conducted during June and July 1989 by investigators from the  Center Hill
Facility, who used equipment that was either rented or designed for the project.
Hydraulic fractures were successfully created during the field tests, and then they
were exposed on the walls of trenches dug with a backhoe. Detailed descriptions of
exposures document the geometries of the fractures, highlighting the potential that
this process should have in remediation of contaminated soil.
MAJOR CONCLUSIONS

      The project consisted of studies in the laboratory and in the field, as well as
an investigation of possible applications and a review of previous work. Most
remedial systems requiring fluid flow either into or out of the subsurface could
benefit from hydraulic fracturing. Pump and treat systems are obvious candidates
because they employ procedures resembling those used in petroleum recovery,
where the benefits of hydraulic fractures are without question.  The yields of vapor-
producing wells, such as those recovering natural gas or steam, are improved by
hydraulic fracturing, so by analogy we expect that vapor extraction systems could
benefit from this technology.  Similarly, hydraulic fracturing could be used in
conjunction with steam stnpping-a process developed to improve yields of oil wells
and currently being tested under remedial conditions. Horizontal, sheet-like
hydraulic fractures placed below a contaminated region could be used as gravity
drains to intercept the leachate from soil flushing systems (Murdoch and others,
1987), improving the effectivness of that remedial action. Bio-remediation systems
stand to benefit in particular because either nutrients for microorganisms, or the
microorganisms themselves, could potentially be delivered as fine-grained solids in
hydraulic fractures.

Laboratory Studies: Summary of Results

   1. Hydraulic fracturing in the Center Hill clay (a type CL clayey silt) can be
      predicted using methods of linear elastic fracture mechanics, according to
      both empirical evaluation and estimates of the size of the crack tip process
      zone. Many published methods of analyzing hydraulic fracturing in rock,

-------
       which are based on linear elastic fracture mechanics, can be used to analyze
       applications in soil, with appropriate modifications for boundary conditions
       (e.g. ground surface), material properties (e.g. elastic constants, leakoff
       parameters), and state of stress encountered at shallow depths.

   2. The appearance of hydraulic fractures in silt, and clayey silt resemble
       published descriptions of hydraulic fractures in rock created during
       laboratory experiments.  Details of the fracturing process are sensitive to
       moisture content of the soil; moisture affects the appearance of the fracture
       tip, the magnitude of driving pressure required to initiate fracturing, and the
       form of the pressure record during propagation of the fracture.  Changes in
       all those factors with changes in moisture content are particularly large when
       moisture conditions are near the plastic limit of the soil.

   3. The critical stress intensity factor can adequately predict the pressure required
       to initiate hydraulic fracturing of soil. It is independent of the length of
       starter slot (or pre-existing fractures), for the slot lengths used in this study,
       and it predicts the driving pressure at the onset of fracturing to within 10
       percent, on average. It was more accurate for samples whose moisture
       contents (moisture content = wt. water/wt. solid) exceeded 21% and less
       accurate for drier samples.

   4. The value of the critical stress intensity is highly sensitive to moisture content,
       decreasing sharply from roughly 200 kPa cm1/* to 30 kPa cm1/2 as moisture
       content increased by a few percent in the range of the plastic limit.  Critical
       stress intensity goes to zero for samples of Center Hill  clay greater than 32%
       moisture, although hydraulic fractures are readily created under those
       moisture conditions.

   5. Theoretical analyses based on linear elastic fracture mechanics and linear
       viscous fluid mechanics explain many of the details of the laboratory
       experiments, including the development of an unwetted tip, the average
       propagation velocity, changes in the form of the pressure record with changes
       in slot length or moisture content.

Field Studies: Summary of Conclusions

   1. Hydraulic fractures of useful size were created and propped with sand at
       shallow depths in glacial drift during field tests in 1988 and 1989.

   2. A technique was developed that consistently produced hydraulic fractures
       propped with sand in glacial drift. The technique uses a unit, which is readily
       available for injection grouting, to inject a slurry of sand and cross-linked
       guar gum. To create a fracture, an apparatus composed of an outer casing
       and an inner rod was driven out through the bottom of a hollow-stem auger.
       The inner rod was removed and a water jet was used to cut a disk-shaped slot
       in soil enveloping the casing. Slurry was injected into the casing with a
       positive displacement pump, nucleating a hydraulic fracture at the slot.

   3. Hydraulic fractures created during the 1989 test were nearly flat-lying; their
       steepest dip was 5°. In plan view, they were slightly elongate with aspect

-------
       ratios of 2:3, and they were highly asymetric with respect to their parent
       boreholes. They were createdat depths ranging from 1 to 4 m, and the
       shapes and sizes of fractures created at depths between 1 and 2 m are known
       in detail from excavations. The dimensions are as follows:  maximum
       distance from the borehole ranged from 0.8 to 7.2 m, and averaged 4.0 m.
       Areas covered by the fractures ranged from 0.8 to 36.7 m2, and averaged 19.2
       m2. All the fractures contained sand, ranging in maximum thickness from 2
       to 20 mm, and the average maximum thickness was 11 mm.

   4. The fracturing apparatus facilitates the creation of multiple fractures from a
       single borehole. Stacks of flat-lying fractures with vertical spacings from IS
       to 30 cm were created with the device.

   5. The fractures were highly asymetric with respect to their parent borehole; that
       is there was always a preferred direction of propagation.  This direction was
       related to loading at the ground surface, with the fracture either growing
       away from a backhoeparked next to the borehole, or growing in the direction
       of maximum slope.  This indicates that topography, or surface loading (e.g.
       vehicles, buildings), will affect the direction of propagation of shallow
       hydraulic  fractures.

   6. Pressure records, surface uplift, and surface tilt all are available methods of
       monitoring hydraulic fractures at shallow depths. Electrical resistivity shows
       promise, although further research is required before that method is a
       practical monitoring tool.
   7. The results of hydraulic fracturing at a site i               _,	„
      consolidated soil or fill are expected  to differ markedly from'die results
      described here. Further, during this  study fractures were created in
      unweathered material, so the effects  of creating hydraulic fractures in
      weathered material is unknown.
RECOMMENDATIONS

      The principal recommendation is to expedite further investigation of field
applications. The results of this study indicate that hydraulic fracturing potentially
offers important improvements in the effectiveness of in situ remedial technologies,
particularly at sites underlain by overconsolidated soil or bedrock, and our
recommendation is to develop techniques necessary to assess this potential.

      Specifically, we recommend the following:

   1. Hydraulic fracturing should be tested during the remediation of several
      contaminated sites.  Those field tests should  assess the impact that hydraulic
      fracturing has on several different remedial technologies (e.g. pump and
      treat, vapor extraction, soil flushing, bio-remediation).  They should also
      assess the impact of site conditions (e.g. geology, state of stress, depth of
      contamination, etc.).

-------
2. Design and construct equipment intended to create hydraulic fractures at
    shallow depths. Equipment should be a compromise between the powerful
    and sophisticated ng used during the 1988 test (which lacked the required
    control), and the modest and simple unit used during the 1989 test (which
    lacked the required power and resiliency).

3. This study focused on creating hydraulic fractures in overconsolidated drift,
    and the results of hydraulic fracturing in normally consolidated material may
    differ markedly from the results reported here.  We recommend that future
    studies develop methods of creating hydraulic fractures at shallow depths in
    normally consolidated soil, where in situ stresses will favor vertical fractures
    that are highly susceptible to venting.

4. Develop techniques of delivering nutrients and microorganisms used for bio-
    remediation.

5. Design methods of completing and developing ground water wells intersecting
    hydraulic fractures.

6. Adapt or develop analyses to a.) determine treatment parameters from
    pressure records (e.g. Nolte and Smith, 1981,1987; Nolte, 1988); b.) predict
    the size, shape and orientation of hydraulic fractures at shallow depths,
    particularly considering effects of topography; c.) optimize the position and
    dimensions of hydraulic fractures based on predictions of the transport of
    contaminants in the vicinity  of the fractures.

-------
                               SECTION ONE

 POTENTIAL REMEDIAL APPLICATIONS AND PREVIOUS INVESTIGATIONS

       The final decision on the use of hydraulic fractures during remediation of
 contaminated soilwill depend on whether the benefits they produce outweigh the
 effort of creating them. Most of the following report deals with the processes and
 problems of creating hydraulic fractures in soil, but it seems prudent at this point to
 review some of the potential applications and estimate the order of benefits that can
 be expected. Here we will assume that fractures of useful size can be created and
 filled with a permeable proppant.


 ASSESSMENT OF POTENTIAL APPLICATIONS

       Most remedial systems that require fluid flow either into or out of the
 subsurface could benefit from the effects expected to accompany hydraulic
 fracturing. Pump and treat systems are obvious candidates, as are systems involving
 vapor extraction, soil flushing, or steam stripping. Hydraulic fractures offer the
 novel possibility of delivering solid material, formed into granules and mixed with
 proppant, to the subsurface. Nutrients for microorganisms, or the microorganisms
 designed for bio-remediation themselves, could be delivered as solid grains to
 contaminated regions.

      Filling a hydraulic fracture with material of low permeability, such as a
 bentonite-rich soil or grout, was proposed by Buck and others (1980) as a method of
 isolating contaminated material. This proposal has been investigated
 experimentally in the laboratory (Brunsing and Henderson, 1984), and tested in the
 field at a site near Whitehouse, Florida (Brunsing, 1987). The ability to place a flow
 barrier beneath a contaminated region without removing overburden is the principal
 advantage of this application. The principal drawback lies in the inability to verify
 continuity of a low permeability layer. Breaking into segments is a property of
 hydraulic fractures that seems to be inevitable.  That property, combined with the
 presence of objects such as tree roots, will cause discontinuities in the fracture-filling
 that would be impossible to predict and difficult to detect using current methods. It
 is feasible that hydraulic fractures filled with grout could act as short-term barriers,
 or impediments to flow, but the likelihood of undetectable discontinuities inhibits
 their use as long-term barriers.

      Increasing the rate of flow from a well is expected to be the most widespread
 application of hydraulic fracturing in remediation. The magnitude of increase will
 be site-specific, but a general estimate of what to expect can be obtained from the
 results of related applications and from theoretical analyses.

      Records of oil production are irrefutable testimony of the benefits of
 hydraulic fracturing, and ratios of production rates are a suitable yardstick with
which to measure those benefits. According to data from several dozen oil wells
 (Howard and Fast,  1970; Table 1.1), production ratios (production rate after
fracture/initial rate) are at least  1.5 and range up to very large values. In general,
the ratios range from 1.5 to roughly 10 for wells that were producing before
fracturing, and the ratios are very large for wells showing negligible production prior
to fracturing.

-------
                                         BATE OF OIL PRODUCTION
STATE
Alaska
Calif.
Canada
Colorado
Illinois
Kansas
Louisiana
Michigan
New Hex.
H.Dakota
Ohio
Oklahoma
Penn.
Texas
Heat Vs.
FORMATION
Eeaai sand
Stephana sand
Vaqueros sand
Beaverhill la.
Cardium sand
Villenuve quartz
Weber sand
Cypress sand
McCloakey Is.
Arbuckle la.
Hernnfton la.
Kansas City Is.
Annona Chalk
Hackberry sand
Marauilana sand
Blchfield la.
Graybur* la .
San Andres Is.
Mesa Verde
Madison la.
Berea sand
Bartlesville sand
Baaal sand
Granite wash
Mississippi Is.
Pennsylvanian
Bradford aand
Bron dolomite
Caaerina aand
Canyon limestone
Reklaw sand
Strawn sand
Vilcox sand
Benson aaad
in Barrels of Oil per Day
Before After Ratio
frx Irx (after/before)
1128
20
10
0
50
0
20
4
4
10
6TO>
6
5
30
15
22
15
20
4901
111
0
25
80
120
trace
130
6
6
26
SO
72
10
3.S*
1584
120
70
75
204
4*
110
35
60
40
20501
30
20
115
104
ISO
110
60
5000*
296
60
275
350
650
168
4T5
55
65
230
130
113
35
1100*
66.1*
1.4
6.0
7.0
4.1
5.5
8.7
12.5
4.0
3.1
5.0
4.0
3.0
6.9
6.8
7.3
3.0
10.2
2.6
-
11.0
4.4
* I6
3.6
9.1
10.8
8.8
2.6
1.6
3.5
73.0
18.9
                      Dakota aand        25       125      5.0
                      Frontier aand       16       70       4.4
                      Madison linestone   473      715      1.6
                      Phosphoria Is.      1050*    3050s    2.9
                      Tensleep aaand      20       65       3.2

                                      *  units of (Hcubio ft./day)
Table 1.1 Data describing the production rates from oil wells before and after
              hydraulic fracturing. From Howard and Fast (1970).

-------
       Production rates, from either fractured or unfractured wells, typically
decrease as a function of time due to depletion of the reservoir, clogging of pores, or
other processes. Behaviors of fractured wells through time were characterized by
Howard and Fast (1970) in their classic monograph,

             "The effect of [hydraulic] fracturing on both short and long term well
       productivity has been studied by many investigators, most of whom conclude
       that, regardless of the kind of treatment, four basic patterns of production
       behavior have been observed.

             Type A:   Sustained increase in well production accompanied by a
       flattening of the production decline curve following treatment.

             Type B: Sustained increase in production with the well's highest rate
       of production  after the treatment declining essentially at the same rate
       established before treatment.

             Type C: Transitory increase in production lasting from a few months
       to several months, after which the well continues to  follow the production
       decline trend observed prior to treatment.

             Type D: No increase in production, with the well continuing to follow
       its established, normal production history.

             In no case did a treatment have a discernibly detrimental effect on the
       production performance of the well."

       In one example of a Type A or B result (Fig. 1.1), the production rate at a
well in North Texas increased by a factor of four, from 20 to 80 barrels of oil per day
  £opd), and then gradually decreased to 30 bopd over the subsequent seven months.
  terestingly, a second hydraulic fracturing job increased the production rate to 80
bopd (Fig. 1.1).

       Yields of water wells are also increased by hydraulic fracturing (e.g. review
by Smith, 1989). Thirty years ago Koenig (1960) examined data from wells used for
waterflooding or waste disposal and reported that 78 percent of those wells showed
an increase in yield following hydraulic fracturing.  The ratio of yields ranged up to
100, with a median of 5.0.  More recently, Stewart (1974,1978) observed increases
in yields from water wells drilled in granite or schist formations in New Hampshire.
Yields of one well reported by Stewart increased from 0.30 to 1.821/s, (ratio of 6.0),
and at another well yields increased from 0.26 to 1.141/s (ratio of 4.4). A
particularly impressive increase in yield of 22 to 25 times is described by Mony
(1989), who cites a water well drilled in gneiss that yielded 0.00771/s prior to
fracturing and 0.16 to 0.191/s after fracturing. Waltz and Decker (1981) used
hydraulic fracturing to stimulate wells in crystalline rock in Colorado and reported
ratios of yields of 1.5  to 2.0. Williamson and Woolley (1980) created hydraulic
fractures from water wells penetrating igneous or metamorphic rocks in Australia,
and cited improvements in specific yield by factors of 5.0 to 6.0. Several other
authors (Hurlburt, 1989; Waltz, 1988; Baski, 1987; Macaulay, 1987) have recently
claimed that hydraulic fracturing of water wells consistently results in increased
yields that are economically significant.

-------
  Q
  D-
  O
  PQ
  %_^

  to
  C
  o
  O
  S-,
  OH
      100
80
       60 I-
                            Hydraulic fracture created
            Hydraulic fracture created
         0     2     4    6     8    10   12    14   16   18

                              Time (months)
                                                          20
Figure 1.1. Production rate of an oil well in North Texas as a function of time.

-------
      Rates of inflow into wells used for waste disposal (Stow and others, 1985;
Wolff and others, 1975; De Laguna, 1966 a and b), or as infiltration galleries are
increased by hydraulic fracturing. Rates of inflow were measured as part of this
project from five open boreholes intersecting sand-filled hydraulic fractures created
in unsaturated glacial drift.  The fractures were flat-lying, typically 1 cm in maximum
thickness and as much as 8 m in maximum dimension.

      For comparison, inflow rates were measured from three similar boreholes
penetrating unfractured drift.  Boreholes intersecting hydraulic fractures are EL2,
EL5, EL6, EL7, whereas those in unfractured ground are TB1, TB2, TB3.

      All boreholes were 5 cm in diameter and 2 m deep.  Tests were conducted
using Guelph Borehole Permeameter, manufactured by SoilMoisture Inc., Santa
Barbara, Ca. The permeameter enabled us to measure flow rates while maintaining
a constant water depth of one meter above the bottom of the borehole.

      Rate of inflow into unfractured boreholes diminished slightly with volume,
approaching steady-state rates after one to three liters.  The steady-state rates were
between 0.04 and 0.071/min, depending on the well, and the average rate was 0.055
I/min.

      The rate of inflow into the wells intersecting hydraulic fractures also
diminished with time, but larger volumes, as much as several tens of liters, were
required to achieve steady-state. Initially, inflow rates were as rapid as 2 to 31/min,
the upper limit that could be supplied by the permeameter. The rates diminished
with volume, but they were greater than those of the unfractured wells even after
large volumes. For example, the rate of inflow was roughly 0.41/min after 200 liters
of water flowed into EL7, and the rate was 0.175 after 350 liters flowed into EL4.
Steady inflow into the fracture wells occurred at 0301/min.

      In the works cited above, effects of a hydraulic fracture were illustrated using
the ratio of flow after fracturing to flow prior to fracturing. The method used to
create the hydraulic fractures precluded measuring inflow at a well before it was
fractured. The average inflow into an unfractured well in the vicinity of a fractured
well will be used to determine the ratios (Table 1.2).  Values range from 3.1  to 9.0
for steady-state conditions, and they are typically several times greater than that for
unsteady state (Fig.  1.2).

      Inflow into the boreholes that intersect fractures probably would have been
greater if screen and gravel packing was used to prevent collapse. The boreholes
filled with mud to depths of several dm while they contained water during inflow
testing.  Nevertheless, the data do suggest that hydraulic fractures in unsaturated
ground could provide marked improvements to remedial systems based on soil
flushing.

      Wells used for purposes ranging from delivery of water in unsaturated soil to
recovery of oil in reservoirs have shown strikingly similar responses after hydraulic
fracturing - their yield is consistently increased, and the magnitude of increase
ranges up to several orders of magnitude or more. Typically, the ratio of yields is
roughly 1.5 to 8 or 10 for wells that are producing prior to fracturing, but it can be
greater than 10 for wells that are initially poor producers.
                                     10

-------
 M
o-
0 15
0.10

0.05
0 15

0.10
0.05
0 OO
0.
0 15
0.10

O.OS
0.00
TB 1


) 1 2
TB 2

° o ° o o - o °
„ 0 0 0 ° c 00 O
o
0 02 0.4 OS 0.6 1.0 1.2 1.4
TB 3
° °V". . . •'
Oo°oo0°»c °° e 0ta000o00°C00coo
o

2

o
3 <

2
1
I
2

1

o o
EL 2
0
0 0
00
0
o
e
J 10 20 30 40 5C
EL 5

0
° °o°Q°°aaoe_0 a eoo ° ° °° o uo-oo"
2 4 6 8 10 12 14 16
EL 6
o
e °
° ° e
B8W>°° . # .'^ssato-.nv
                                          a  o    z    4   s    e    10   iz
                             Delivered Volume (liters)
    Figure 1.2 Rates of inflow into boreholes in till. TB1.TB2.TB3, are test borings,
                whereas EL2 through EL7 are borings intersecting hydraulic fractures.
                                    11

-------




x^*
• f-l
^s
w
0)
«4«J
• 1— 1
I— H
0*





U.3 1 ' • • • 	 	 1 ' ' ' ' 1 	 i i i i
o.4 | EL 4 -
0.3 i
*
0.2
0.1

0.0
(
1 1 ° •
» 8 :
o -
o ;
. • i







) 100 200 300 400


2n
• V
1.5

1.0

0.5

OD
: EL 7
j
i o
3 «<
° o o o o o dCo o
0 SK
1 1 1 1







                  50
100
150
200
             Delivered Volume  (liters)
Figure 12 (continued).
                          12

-------
TABLE 1.2  INFLOW RATES INTO SILTY CLAY
Borehole      Steady inflow          Average        Ratio
TB1              0.04
TB2              0.05
TBS              0.075                0.055

EL2              0.20                                3.6
EL4              0.175                               3.1
ELS              0.50                                9.0
EL6              0.20                                3.6
EL7              0.40                 0.30           73
Predicting Flow to Wells That Intersect Fractures

      Early schemes of predicting flow to wells that intersect hydraulic fractures
were based on empirical studies, electrical analogs, and simple theory. Results were
obtained for fractures that were either vertical and rectangular in shape (Prats,
1961, McGuire and Sikora, 1960; Dyes and others, 1958; van Poolen and others,
1958), or horizontal and circular in shape (Landrum and Crawford, 1961; Hartsock
and Warren, 1961; Morrisson and Henderson, 1960). Today, those schemes are
important for simple calculations, but more sophisticated numerical models
(Gringarten, 1982; Cinco-Ley and others, 1981,1978) are available for detailed
analyses.

      Many of the available analyses can be used to obtain the specific yield of a
well intersecting a hydraulic fracture, which when divided by the specific yield of the
well without a fracture results in a dimensionless measure of the improvement of
the well. This ratio will be termed the recovery capacity / in the following pages.
Recovery capacity in general depends on conditions of the fracture, such as its size,
shape, aperture, location, and permeability of proppant, and on conditions of the
reservoir or aquifer, such as its permeability, dimension, and the location of the well
within it

      Two analyses will provide useful insights: one that shows the importance of
fracture length, relative transmissivity, and radius of influence at steady-state, and
another that shows the importance of fracture length as a function of time. Prats
(1961) analyzed the effect of a vertical fracture of finite permeability ki and
thickness d in a confined aquifer bounded laterally by surfaces of constant pressure.
Relative transmissivity Tt is defined by Prats (1961) as


                                         *.«                         (1.1)
                                   13

-------
where ks is the permeability of enveloping soil or rock, a is fracture half-length.
Prats describes Tr as the ratio of the ability of fluid to travel along the fracture to the
ability of fluid to travel through the enveloping material and reach the fracture. The
flow at the fracture increases with relative transmissivity for values of Tt between 0
and 1.0. For values of Tt greater than 1.0, however, the effect of T, is negligible
(Prats, 1961) and the fracture behaves at steady-state as if its permeability were
infinite. This result is important because  it indicates that in many field cases the
permeability of a propped hydraulic fracture can be assumed to be essentially
infinite, thereby avoiding analytical difficulties associated with fractures of finite
permeability.

      Assuming that the relative transmissivity is essentially infinite, the steady-
state recovery capacity is


                          / = Mre/rw)/7n(re/(L/4))                      (12)

in which rw is the actual radius of the well, and re is the effective radius of the  area
drained by the well, and L is the total fracture length. In a field of regularly-spaced
wells, for example, the effective radius is roughly half the spacing between wells.

      According to eq. (12) the steady-state behavior of the fracture is the same as
that of a well whose radius is equivalent to L/4. A vertical fracture 10 m long would
behave as a well 25 m in radius, so that the equivalent radius of a fracture can be
much larger than most wells.

      Recovery capacity certainly increases with fracture length, but its magnitude
depends on the ratio of fracture length to effective radius.  According to this
analysis, the recovery capacity will be greatest when the fracture extends across the
area drained by the well. Assuming L = 2re, the maximum / is


                             Jam* -  1-44 ln(L/2rv)                        (1.3)
       For example, 2.3 is the maximum recovery capacity of a fracture 10 times as
long as the radius of the unfractured well, whereas the value increases to 5.6 for a
fracture 100 times the well radius, according to this analysis.

       As the effective radius increases, the steady-state recovery capacity decreases
and approaches unity.  For a single well in a large aquifer, therefore, the steady-
state yield from a fractured well will approach that of one from an unfractured well.
This conclusion is misleading, however, because it ignores potential increases in
yields during the transient period of recovery before steady-state is achieved.

       Recovery capacity is typically greatest shortly after fracturing and then it
decreases with time, as shown in Figure 1.1 for example. A solution to the transient
recovery capacity can be obtained in analytical form from dimensionless drawdown
functions for a well and for a fracture.  Dimensionless drawdown S^ of a vertical
well in a confined aquifer of thickness h is obtained by assuming the well behaves as
a line sink (Ramey, 1967), so
                                     14

-------
                                                                      (1.4)

where


                               I*, = 4tKh/Sr*                          (1.5)

and K and 5 are the hydraulic conductivity and the storage coefficient of the aquifer,
respectively; t is the time of pumping, p, Q, and r are the drawdown, pumping rate
and radius of the well, respectively.  The drawdown function sa«(l/t  1.0 are a result only of the difference in geometry between a sheet-like fracture
and a line-like well.

      Many hydraulic fractures created at shallow depths (less than roughly 300 m)
are flat-lying and roughly equant, a geometry that can be approximated by a planar
sink that is flat-lying and shaped like a circular disk. The drawdown function for a
fracture of that geometry in a confined aquifer (e.g. Gringarten and Ramey, 1974;
                                      15.

-------
 eqs. 51 and 53) was used in eq. (1.7) to produce Figure 1.3. The ratio of aquifer
 thickness 7* to well radius r was held constant and equal to 250, and the fracture was
 assumed to be at mid-height in the aquifer (z/T=0.5 where z is the height of the
 fracture above the bottom of the aquifer). Recovery capacity was determined as a
 function of time for ratios of a/r between 50 and 1000. The radius of the fracture
 equals the  aquifer thickness when a/r equals 250, and it is four times the aquifer
 thickness when a/r equals 1000. Although the dimensionless form is valuable for
 many applications, a reference value of fdr may be helpful in some cases: if K = 104
 cm/sec, 5 = 0.001, a = 10 m, T = 11.6 m; then tu x 10 = / in days.

       Results of the analysis indicate that the recovery capacity at any given time
 increases with relative fracture length. Recovery capacity is roughly unity when a/r
 = 50, indicating that the recovery from a fracture 50 times the radius of a well will
 be similar to that of an unfractured well. As the ratio a/r increases to 1000, the
 value of/ increases to 8.0, and even greater values of/ are obtained for earlier times
 or longer fractures.

       Magnitudes of/ diminish with time because geometric advantages of a planar
 sink decrease compared with those of a line sink as water is removed from regions
 at increasingly greater distances from the sinks. Even at relatively large values of to,
 however, the recovery capacity is significantly greater than 1.0 (Fig. 1.3), indicating
 that the fracture is performing better than the well. The performance of the
 fracture would be even better if we assumed that the well draining the fracture was
 screened for the full thickness of the aquifer, which is a reasonable assumption for
 most applications.  That case can be approximated by adding 1.0 to the right hand
 side ofeq. (1.7), or to values of/ shown in Figure 1.3.

 Summary

       In the previous section we have seen that hydraulic fracturing increases the
 yields (or inflow rates) of oil and gas wells, water wells, and infiltration wells. The
 magnitude  of increase is consistently between 1.5 and 5 times, routinely as much as
 10 times, and in some cases much more than 10 times'. In general, hydraulic
 fracturing offers the greatest relative increase to wells which show poor yields prior
 to fracturing. Of course, if the initial yield of a recovery well is extremely low, even
 the improvement offered by hydraulic fracturing may be insufficient to make
 recovery a viable remedial solution.  That judgement will depend on analyses of
costs, which exceed the scope of this investigation.

       Observations indicate that yields of fractured wells diminish with time, just as
 they do at unfractured wells. The relative improvement of the well diminishes
following fracturing, but the effect of the fracture can be important even after many
months or longer. In some cases, wells can be re-stimulated and their yields
increased creating another hydraulic fracture.

       A simple theoretical analysis, based on the difference in shape between plane
and line sinks, predicts improvements in specific yield that are similar, both in
magnitude  and behavior with time, to improvements observed in the field. It
follows that improvements in yield result principally from the geometric effects of
the fractures.
                                      16

-------
                                           T/r: 250
                                           z/T: 0.5
                                                                  1000
Figure 13. Recovery capacity of a flat-lying, circular fracture in a confined aquifer.
                                   17

-------
       We conclude that the magnitude of increase in yields of contaminant
 recovery wells will be similar to that of wells used to recover oil, natural gas, or
 water from rock.  Hydraulic fractures are expected to increase the yields of recovery
 wells in soil or rock by several times, or perhaps up to an order of magnitude.
 Detailed predictions of the effects of hydraulic fractures at a particular site will
 require more sophisticated analyses, but the general magnitude of the improvement
 in yield predicted by the data and analyses described above justifies further
 investigation into this process.


 PREVIOUS STUDIES OF HYDRAULIC FRACTURING

       To some investigators hydraulic fracturing is a blessing, to others it is a curse.
 This difference is driven by the effect of fracturing on particular processes, and it
 has resulted in two different research goals.  When hydraulic fracturing produces
 useful results, such as increasing the yield of a well, providing a measure of in situ
 stress, or generating insight into geologic process, investigators generally focus on
 predicting the characteristics of a fracture  (e.g. size, shape, orientation). When
 hydraulic fracturing causes problems, such as collapsing a dam, compromising a
 permeability test, or bleeding off grout or waste from a borehole, however,
 investigators generally focus on predicting, and thus preventing, initiation. To the
 latter group, the characteristics of hydraulic fractures are of little interest.

       A thorough review of the vast body of published work related to hydraulic
 fracturing would fill several volumes, so the following section is intended only as a
 brief overview.  Other reviews relevant to the present work have been published by
 Clearv (1988), Mendelsohn (1984), Veatch (1983 a and b), Geertsma and Haafkens
 (1979Y  The classic SPE monograph on hydraulic fracturing by Howard and Fast
 (1970) is a valuable source, and a revision  of that monograph will be published in
 the near future.

 The Hydraulic Fracture as a Tool

      As early as the mid-1930s it was widely recognized by workers in the
 petroleum industry that pressurizing a well could fracture the enveloping formation
 (Yuster and Calhoun, 1945). The first use of proppants to hold fractures open,
 however, was the key to developing a technique of widespread importance. The first
 description of the creation of propped fractures is generally attributed to Clark
 (1949), who outlined a process that closely resembles the ones used today. Clark
 claimed that hydraulic fractures had been created at 32 wells when his paper was
 submitted for publication in 1948. Several years later in 1953, a patent of the
 hydraulic fracturing method was reissued to R.F. Ferris, a colleague of Clark
 (Howard and Fast, 1970).
       During the 20 years following Clark's paper the ap
fracturing to oil and gas wells became routine, and by 1968 the technique had been
used at roughly 500,000 wells (Howard and Fast, 1970). Success rates, indicated by
an increased yield after fracturing, were between 75 and 80 percent during the early
years, but they increased to nearly 90 percent by the late 1960s, according to
Howard and Fast (1970).
                                    18

-------
       Typical fracturing operations pumped volumes on the order of several
 thousand to several tens or thousands of gallons prior to 1970, but in the following
 years the sizes of fracturing jobs increased to a million gallons or more. These so-
 called massive hydraulic fractures opened oil reserves mat could not be
 economically produced using previous methods. Costs involved with creating
 massive hydraulic fractures spurred the modern era of research into methods,
 materials, analyses and designs of hydraulic fracturing.

       Petroleum companies are the primary users of hydraulic fractures, but they
 are by no means the only ones. The generation of thermal energy from hot dry
 rocks is a technology that makes use of hydraulic fractures to create flow paths in
 basement rock. In principle, two wells drilled into hot rock are linked together by
 one or more  hydraulic fractures (e.g. Murphy, 1982; Ernst, 1980). Cold water is
 injected into one well, it flows through the hydraulic fractures and is heated by the
 wall rocks, and then is recovered at the other well where it is used to drive turbines
 for electrical power. This application was successfully implemented in a small-scale
 test at Fenton Hill, New Mexico, where two wells 2.75 km in depth were connected
 by a fracture 0.3 km in height (Kerr, 1987; Murphy,  1982).  The hydraulic fracture
 acted as a heat exchange producing 3 megawatts of power. A second phase of that
 project, intended  to create larger fractures and produce more power, encountered
 difficulties. Two boreholes were drilled to depths 9f 4.6 km, but hydraulic fractures
 created at one well failed to intersect the neighboring well, presumably injection
 resulted in slippage along many natural joints rather than dilation of a single
 fracture (Kerr, 1987). In a related project in Cornwall, England, investigators used
 gel to open natural fractures, successfully creating a circulation loop between two
 wells. This project was producing 5 megawatts of thermal power after two years of
 operation (Kerr, 1987).  Similar projects involving the circulation of water through
 hydraulic fractures Unking two  boreholes have been evaluated in France (Cornet
 and others, 1982), and Germany (Rummel and Kappelmeyer, 1982).

       Proppant was omitted from fractures created during the  projects cited above,
 because thermal breakdown of the guar gum-based gel used to transport proppant
 could cause sand to plug the well casing during fracturing. A modified gel, intended
 for use at high temperatures, was employed by Nakatsuka and others (1982), who
 created hydraulic fractures used to recover wet steam at the Nigorikawa geothermal
 field in Japan. They injected as much as 81 metric tons of sand  per fracture, and
 increased the yields of steam or hot water by factors as great as  2.3.

      Water wells can be stimulated by hydraulic fracturing, as indicated by data
 cited in a previous section. Most of the applications are to water wells in igneous or
 metamorphic rock, where the natural permeability is low and the wells are initially
 nearly dry (Hurlburt, 1989; Waltz, 1988; Baski, 1987; Macaulay, 1987; Williamson
 and Wooley, 1980; Stewart,  1974,1978). The use of proppant in hydraulic fractures
 at water wells is unnecessary in many cases because sufficient yields can often be
 achieved using water alone (Smith, 1989; Baski, 1987; Williamson and Woolley,
 1980).  Presumably this occurs because fractures in hard rock can be held open by
asperities on fracture surfaces (Paillet,  1985; Detournay, 1979).  Omitting proppant
from a fracture eliminates the need for a blender and a sand supply, considerably
reducing the expense of the  fracturing operation.
                                  19

-------
       Several other industries that use wells to gain access to the subsurface have
applied hydraulic fracturing. The mining industry, for example, extracts soluble
minerals such as halite, sylvite, sulfur, or uraninite by circulating fluids through
natural ore deposits. Solution mining operations commonly employ two wells, one
to produce and another to inject fluid. Hydraulic  fractures are used to increase the
flow rate between the two wells, in a design similar to that of geothermal energy
operations (Haimson and Stahl, 1970).

       Various industries have disposed of wastes by injecting them into hydraulic
fractures. In 19S8, injecting radioactive waste into hydraulic fractures was first
considered by the Atomic Energy Commision as a means of disposal (De Laguna,
1966). This application was evaluated for 20 years at the Oak Ridge National
Laboratory (Stow and others, 1985), where radionuclides were mixed with cement-
based grout and injected into a shale formation. The grout formed nearly flat-lying
hydraulic fractures roughly 300m below the ground surface (Stowe and others, 1985;
Holzhausen and others, 1985 a and b; De Laguna, 1966).

       Deep well injection is a popular technique  of waste disposal that can either
benefit or suffer from hydraulic fracturing. On one hand, hydraulic fracturing
during deep well injection will increase the rate of injection (Bouwer, 1978) and
thereby reduce the. cost of disposal. On the other  hand, hydraulic fractures can
propagate upward from the bottom of the disposal well, cutting through low
permeability formations intended to isolate the waste and possibly contaminating
overlying aquifers (Wolff and others, 1975).

       The state of stress in the earth's crust is a central topic of interest to
investigators of tectonics, structural geology, earthquake prediction, mining, well
stimulation, soil or rock mechanics, and related disciplines. Hydraulic fracturing has
been used by those scientists since the early 1960s, when Scheidegger (1960,1965),
Kehle (1964), and Fairhurst (1964) proposed that  the pressure required to initiate a
hydraulic fracture from a well bore was related to the magnitude of in situ stress.
Their method idealizes a wellbore as a pressurized cavity and assumes that
hydraulic fracturing occurs when the magnitude of tensile stress in the wall of the
cavity exceeds the tensile strength 5 of the formation. Assuming the wellbore is a
long cylindrical cavity in a porous formation, the maximum horizontal compressive
stress is given by
                          JHinax = 5 + 3St\mm '/>w ~Po                     (1.8)


in which pw is the pressure in the wellbore at the onset of fracturing, sHmin is the
minimum horizontal compressive stress, and/?0 is the pressure of pore-fluid in the
formation.

      To solve eq. (1.8), tensile strength 5 is determined by testing core samples
and the pressures po,pv are obtained from transducer measurements before and
during fracturing. The minimum confining stress  Hmirfs generally equated to the
wellbore pressure at the instant the fracture closes shut after pumping has ceased
(termed the instantaneous shut-in pressure, or ISIP). Kehle (1964) presents an
                                     20

-------
 eloquent argument supporting the use of the ISIP in determining  Hm» Most
 investigators agree that the ISIP is marked by a subtle change in slope of the
 pressure record after pumping has ceased, but there are a variety otmethods of
 estimating this pressure (e.g. Aamodt and Kuriyagawa,  1982; McLennan and
 Roegiers, 1982; Gronseth, 1979; Nolle, 1979; Bjerrum and Anderson, 1972). The
 methods of Nplte and Smith  (Nolte, 1979,1982; Nolte and Smith, 1981; Smith, 1981;
 Nolte and Smith; 1987) are particularly robust, yielding not only Hmmbut leakoff
 coefficients, proppant schedules, and other parameters used to design hydraulic
 fracture treatments.

       Flaws in the walls of boreholes, either natural fractures or cracks created
 during drilling or perforating a well, will nucleate hydraulic fractures and cause
 errors in in situ stress measurements if they are overlooked (Warpinski, 1983).
 Abou-Sayed and others (1978) present theoretical analyses based on linear elastic
 fracture mechanics that account for pre-existing cracks in a borehole during in situ
 stress measurement. Their analysis requires knowing the size and shape ot pre-
 existing fractures intersecting the borehole, information that is generally difficult to
 obtain for naturally-occurring fractures.  This difficulty is avoided by deliberately
 cutting a notch of known dimension in the wall of the borehole to dominate the
 effects of natural fractures.

       Geologists have recognized natural features that formed by processes closely
 resembling hydraulic fracturing. The features are tabular in form and they are filled
 by materials that differ in composition or texture from enveloping material. Igneous
 sheet intrusions (Pollard, 1978), clastic dikes (Shoji and Takenouchi, 1982),
 hydrothermal breccia dikes (Bryant, 1968; Farmin, 1934), and some mineralized
 veins (Kesler and others, 1981; Anderson, 1974; Phillips, 1972 and 1973) are
 interpreted as resulting from natural processes related to hydraulic fracturing. High
 pore pressures induced by seismic shocks accompanying earthquakes (Holzer and
 others, 1989) can fracture overlying sediments and erupt at the ground surface as
 sand blows, or sand boils (Shoji and Takenouchi, 1982).

       Studies of those natural features have certainly benefited from research
 conducted by the petroleum industry, but it has by no means been a one-way street.
 Natural analogs to hydraulic fractures  can be studied directly where they are
 exposed at the ground surface, a luxury seldom afforded investigators in the
 petroleum industry. Much of what is known about the details of large-scale
 fractures has been derived from studies of exposures  of igneous dikes and sills. In
 their argument for the state of stress as a fundamental control of the orientation of
 hydraulic fractures, for example, Hubbert and Willis (1957) first noted the
 mechanical similarity between dikes and hydraulic fractures. Then they cited the
 similarity between the pattern of dikes at Spanish Peaks, Colorado, and the pattern
 of principal stresses calculated by Ode (1957).  Later, Pollard  (1978,1973) and
 Polard and others (1975) showed tiiat dikes and sills are tabular in gross form, but
 are discontinuous in detail. The discontinuities are particularly well-developed near
 leading edges, where they appear as segments, or elongate lobes. Pollard (1978)
 argues, by analogy, that hydraulic fractures should also be discontinuous, and his
 argument is supported by later theoretical analyses (Pollard and others,  1982)
showing how a fracture could  break into segments in response to small fluctuations
in the in situ state of stress. This insight is significant because  it suggests that
hydraulic fractures filled with  grout and usedas barriers to flow (Huck and others,
                                   21

-------
1980; Brunsing, 1987) could be plagued by imperfections that compromise the
integrity of the barrier. Furthermore, the performance of fractures filled with sand
and used as drains could be reduced if some of the sand is isolated in discontinuous
lobes (Pollard, 1978).
Experiments

       Experiments conducted in the laboratory or the field have contributed to
useful applications of hydraulic fracturing. Laboratory experiments are designed
either to measure material properties, to test theories, or to illustrate effects of
processes that are difficult to analyze theoretically. Fracture toughness2 is a
material property, which was introduced to predict fracture in metals (Invin, 1957),
but has been applied the prediction of fracture in rock as well. Traditionally,
fracture toughness has been ignored (set to zero) because the energy required to
overcome viscous forces in large fractures is several orders of magnitude greater
than the energy required to overcome material toughness (Geary, 1980a).
Recently, however, it has been recognized that fracture toughness can affect
aperture and shape in cross-section (Spence and Turcotte, 1985; Nilson and
Griffiths, 1986; Abe and others, 1976), and that natural variations in toughness can
affect the shape of the leading edge of a hydraulic fracture (Thiercelin and others,
1989).

       Methods of measuring fracture toughness of rock differ widely in specimen
geometry and type of loading, but they all nave one feature in common: the use of a
large notch, or starter slot, cut in the sample. The starter slot nucleates failure at a
pre-existing fracture of known dimension, thereby avoiding the problem of
characterizing effects of naturally-occurring cracks or flaws. Some methods are
adaptations of ASTM techniques, which apply external loads to measure fracture
toughness of metal specimens (Klepaczko and others, 1984; Ouchterlony, 1982;
Geary, 1978b; Schmidt, 1977), whereas others are patterned after the hydraulic
fracturing process and inject fluid to measure fracture toughness (Clifton and others,
1976; Abou-Sayed, 1977).

       Early theories (e.g. Haimson and Fairhurst 1967; Kehle, 1964; Harrison and
others, 1954) predicted that hydraulic fracturing would initiate when the
circumferential stress on the wall of a borehole exceeded the tensile strength of the
formation.  This theory was tested by Haimson and Fairhurst (1969) by injecting
fluid into cylindrical samples of hydrostone, a gypsum cement exhibiting properties
similar to natural rock. They observed that during injection the fluid pressure
increased and then abruptly decreased, and they assumed that fracturing, or
breakdown, initiated at the maximum pressure. According to the elastic solution of
Haimson and Fairhurst (1967), breakdown pressure increases linearly with confining
pressure, but both the slope and intercept of breakdown as a function of confining
pressure decrease if injection fluid penetrates the sample prior to fracturing.
2 The l&nn fracture toughness is used in this work only as a nominal criterion of
    measuring resistance to fracturing, without any implication to specific criteria,
    such as the critical stress intensity factor, critical strain energy release rate, /
    integral (Lawn and Wilshaw, 1975), or the fracturability index (Daneshy,
    1976b).
                                  22

-------
Results of their experiments (as well as those of early studies by Harrison and
others, 1954; and Scott and others, 1953) showed that fluid penetration reduces the
breakdown pressure, similar to the theoretical results. In detail, results of the elastic
theory predicted slightly less breakdown pressures than were actually observed by
Haimson and Fairhurst (1969), a difference they attributed to plastic deformation in
the walls of the borehole that was omitted by their analysis. Medlin and Masse
(1979) performed similar experiments. Their results show that the elastic theory
predicts breakdown pressures at relatively low confining stress, but tends to
overestimate breakdown when confining pressures exceed some critical value, which
depends on rock type. Medlin and Masse (1979) attribute this deviation to plastic
deformation induced by the confining stresses, and they offer a modified theoretical
analysis that is consistent with their results.

      Haimson and Fairhurst (1969) also reported that the breakdown depended
on the rate of pressurization; it increased as the rate increased. Zoback and others
(1977) offered an explanation by using an increase in acoustic emissions (AE),
rather than a change in slope of the pressure record, to indicate the onset of
fracturing.  According to Zoback and others (1977), during relatively low rates of
pressurization an increase in AE occurred at the same time as the breakdown
pressure. As the rate of pressurization increased, the pressure of fracturing (as
indicated by AE) remained roughly constant, although the breakdown, or maximum
pressure increased dramatically.

      Theoretical analyses of fracture propagation have been tested in the
laboratory since the first analyses were proposed.  Harrison and others (1954)
created hydraulic fractures in photoelastic gelatin to verify an analysis of hydraulic
fracturing based on elasticity theory and originally published by Sneddon (1946).
Fractures they created experimentally had ratios or length to width that were similar
to those predicted by the simple theory (Harrison and others, 1954; fig, 6).
Moreover, photographs of the experiments revealed fractures that were elliptical in
cross-section, the shape predicted by Sneddon's theory. More recently, Medlin and
Masse (1984) created fractures in rectangular blocks fitted with ultrasonic
transducers to sense the fracture during propagation. Their results show that the
injection fluid lags slightly behind the leading edge, resulting in an unwetted zone at
the tip of the hydraulic fracture. The existence of an unwetted tip was long
suspected on theoretical grounds (Khristianovich and Zheltoy,  1955; Barenblatt,
1962). Rates of growth, dilation, and the pressure record during fracturing in the
experiments were shown to be consistent with two-dimensional, plane-strain
theories of fracture propagation (e.g. Khristianovich and Zheltov, 1955; Geerstma
and de Klerk, 1969; Spence and Turcotte, 1985).

      The orientation of hydraulic fractures normal to the least principal stress is a
result obtained nearly universally, from the early studies of Hubbert and Willis
(1957), or Haimson and Fairhurst (1969) to the more recent work of Hanson and
others (1979), and Medlin and Masse (1979). Daneshy (1973) showed that the local
state of stress induced by a cylindrical borehole can cause the initial orientation of a
hydraulic fracture to differ from that favored by far-field conditions.  In most of
Daneshy's experiments, fractures contained the axis of the borehole where they
were adjacent to the hole, but the fractures curved or twisted out  of their original
plane until they were normal to the direction of least far-field compression at some
distance from the hole. Similar results were obtained by Medlin and Masse (1979),
                                       23

-------
 who report that the initial orientation of fractures would contain the axis of the
 borehole and ignore the presence of shallow notches cut in the wall of the hole at
 orientations favorable to the far-field conditions.  Deeper notches overcame the
 effects of the borehole, nucleating hydraulic fractures in the plane of the notch
 instead of parallel to the axis of the borehole. The results or those studies are
 important to field applications, such as the tests described in later chapters, where
 horizontal fractures are created from vertical boreholes.

       Model studies have contributed to the understanding of how hydraulic
 fractures behave at interfaces between different materials. This behavior is
 important to applications in the petroleum industry where vertical fractures can be
 contained within a thin formation by overlying and underlying units that inhibit
 propagation. Daneshy (1976) describes results of some of the  first experiments
 where hydraulic fractures were created in layered media in the laboratory. He
 found that the shear strength of the interface was important, with hydraulic fractures
 cutting across well-bonded interfaces but unable to cut across poorly-bonded ones.
 Anderson (1979)  confirmed this result, presenting qualitative evidence that it is the
 relative tensile strengths of the materials on either side of the interface, as well as
 the shear strength and normal loading on the interface itself that determine whether
 a fracture will cross an interface.  Other experiments describing the behavior of a
 hydraulic fracture approaching an interface are described by Papadapolous and
 others (1983), Biot and others (1983), Hanson and others (1978 a and b, 1979), and
 Pollard (1973).

      The distribution of in situ stresses play an important role in the forms of
 hydraulic fractures, according to the results of several laboratory investigations.
 Warpinski and others (1982) designed a test cell that allowed the  radial load on a
 cylindrical rock sample (20 cm in diameter and 20 cm long) to be  varied along the
 axis of the cylinder.  Hydraulic fractures were created in homogeneous samples
 where radial loads at the ends of the^cylindrical axis were 2.0 to 2.75 MPa greater
 than loads at the center of the axis. They found that a stress contrast of that
 magnitude alone was sufficient to inhibit growth of a hydraulic fracture out of the
 zone of diminished confining stress. Similar results were obtained by Ahmed and
 others (1983), who created fractures in large (1.0 m on a side)  blocks of cement
 grout to reduce effects of exterior boundaries.

      The experiments cited above were conducted to examine effects of
 nucleation and growth; other experiments address processes occurring after a
 fracture has been  created.  Of particular interest are processes that could cause a
 reduction in permeability of the fracture, either by embedment, or crushing of
 proppant grains (Howard and Fast, 1970).  A propped fracture mav tend to close if
 the normal stress on the fracture is great relative to the strength of the formation,
 leading to embedment of the proppant grains. Results from experimental studies
 using either chalk or diatomite (Hartley and Bosma, 1985; Strubhar and  others,
 1984; Strickland, 1985) indicate that embedment is limited to roughly one-half a
grain diameter. Permeability is seriously reduced by embedment when the proppant
layer is one to two grains thick, but it is only moderately reduced when the proppant
 is three grains thick, and fracture permeability is unaffected when proppant is
greater than five grains thick (Hartley and Bpsma, 1985). Studies of embedment of
proppant grains into soil have not been published.
                                     24

-------
       Proppant crushing can occur when the confining stress on the wall of the
 fracture exceeds the strength of the proppant particles.  According to results of
 experiments reported in Howard and Fast (1970), for example, quartz sand can be
 crushed at pressures equivalent to depths of two km.  At greater depths a high
 strength proppant, such as aluminum oxide, is generally recommended. Crushing of
 quartz sand should be of no concern to the relatively shallow applications of
 remediation.

       Experimental investigations of hydraulic fracturing are by no means limited
 to the laboratory. Field experiments of most of the applications described in the
 previous section have been conducted to test new ideas or verify new theories.  In
 most field experiments, data are obtained from wells intersecting the fractures, but
 in a few experiments data has been obtained from the fractures themselves. The
 pioneering paper by Clark (1949) begins with a description of a shallow hydraulic
 fracture exposed by excavation. A photograph of the  excavation showing a wellbore
 intersecting a horizontal fracture propped with sand provided Clark with irrefutable
 evidence that fractures could be created using the process he describes.

       An extensive suite of field experiments were conducted at the DOE Nevada
 Test Site, where hydraulic fractures were created in ash-fall volcanic tuff and
 excavated using mining procedures (Northrop and others, 1978). The objectives of
 this program were to evaluate proppant distribution, examine characteristics of a
 fracture intersecting an interface between different formations, evaluate results of
 small-volume fractures as a tool for measuring in situ  stress, and compare the size
 and geometry of actual fractures to those predicted by theory.  Results are described
 by Warpinski (1983), Northrop and others (1978), Tyler and Vollendorf (1975), and
 references cited therein.
Analyses

       Hydraulic fracturing was proven as a technique of increasing the yields of oil
wells long before the physics of the process were understood. As the cost of the
fracturing operation increased, however, it became clear that analyses predicting
fracture geometry (size, shape, orientation) could be used to maximize recovery
performance and minimize expenses.  Khristianovich and Zheltov (1955) recognized
that the basic physics of hydraulic fracturing involves two processes: the flow of
liquid within the fracture, and the dilation of the fracture walls due to deformation
of enveloping material. The processes are coupled, though, in that the pressure
distribution resulting from viscous losses during flow is strongly dependent on the
fracture aperture, but the aperture depends on the amount of dilation caused by the
pressure distribution. Dilation of the fracture resulting from material deformation
is calculated, with few exceptions, using elastic, or poro-elastic theory.  One of the
exceptions, a study by Medlin and Masse (1982), showed that effects of plastic
deformation during fracturing are detectable in the laboratory, but are undetectable
in field applications of hydraulic fracturing in sedimentary rock. Effects of plastic
deformation are generally ignored. Fluid flow within the fracture is commonly
assumed to be laminar and governed by a linear viscous behavior. Many fracturing
fluids are thixotropic, so nonlinear flow laws have been incorporated into solutions
of hydraulic fracture propagation (Pascal, 1986).
                                    25

-------
       Two-dimensional analytical solutions to the problem of a linear viscous fluid
driving a fracture in an elastic medium have been derived that yield length,
aperture, and driving pressure as functions of time for the following geometries:

    1.) Vertical, circular (Perkins and Kern, 1961; Geerstma and de Klerk, 1969
        Abe and others, 1976);

    2.) Horizontal, circular (Perkins and Kern, 1961);

    3.) Vertical, rectangular with large ratio of length to height (Perkins and Kern,
        1961; Nordgren, 1972);

    4.) Vertical, rectangular with small ratio of length to height (Khristianovich and
        Zheltov, 1955; Geerstma and de Klerk, 1969; Geerstma and Haafkens,
        1979; Nilson, 1981; Spence and Turcotte, 1985; Nilson, 1986; Nilson and
        Giffiths, 1986).

       Analytical solutions show that the growth of an idealized hydraulic fracture,
that is the characteristic length (half-length or radius), L, aperture, A, and fluid
pressure, P, as functions of time, t, can be expressed as simple power functions


                                   L = Ci f"                             (1.9a)
assuming loss of the injected fluid by flow through the walls of the fractures, or
leakoff, is negligible. The constants C\, €2, €3 depend on details of each solution
(e.g. references cited above), but in general they depend onproperties of the
injection system and the material containing the fracture. The superscripts b, c, and
d, are constants that depend on fracture geometry (Table 1.3). The lengths and
apertures of both types of vertical fractures, for example, are predicted to growth
faster than those of the circular fractures.

       It may be possible to detect the mode  of growth of a fracture using Table 1.3.
A horizontal, circular fracture is expected to grow by increasing its radius and
compressing adjacent material early in its propagation history, and thus behave as
Case A.  With continued pumping, however, the fracture could cease to grow
radially and increase in volume by lifting its overburden (Case B, Table 13), either
because of a mechanical advantage achieved  with increasing radius (Pollard, 1973)
or because leakoff near the tip increases the sand concentration preventing flow. In
either case, the change in mode of growth could be detected because the pressures
would begin to increase approaching a slope of unity for Case B in Table 1.3.
Similar applications of principles derived from analytical solutions of vertical
fractures have been described by Nolte and Smith (1981; 1987).
                                   26

-------
TABLE 1.3. SOME POSSIBLE CONSTANTS IN EQ. (1.9)
Geometry                             b              c          d

Circular
       Case A'                       2/5            1/5        -1/5
       CaseB"                         0               1           1

Vertical, Rectangular
       small L/H                     2/3            1/3        -1/3
       large L/H                     4/5            1/5         1/5
* Vertical fracture, or deep horizontal fracture. Assumes dilation only by elastic
       deformation of adjacent material.

"  Shallow, horizontal fracture. Assumes volumetric growth only by lifting
       overburden, no radial growth.


       Analytical solutions require that the fracture geometry is known, and that it
is a simple geometric shape. Furthermore, details ofthe fracturing procedure, such
as effects of nonlinear fluid Theologies, transport of proppant grains, leakoff, back
stress, fracture toughness, material interfaces, and nonuniform stress distribution in
the host rock cannot be investigated by most analytical solutions.  In practice, the
fracture geometry is unknown and effects of those and other details can be crucial in
a design strategy. As a result, sophisticated  computer programs have been
developed that offer predictions of the shape of the leading edge of a  fracture, and
are able to estimate now a wide range of design parameters affect the length,
thickness, and distribution of proppant within a hydraulic fracture. There are many
dozen, or perhaps more, published descriptions of computer programs used to
simulate hydraulic fractures. The capabilities of those descriptions are reviewed
and compared by Cleary (1988), Palmer and Luiskutty (1986), Advani and others
(1985), Mendelsohn (1984), and Veatch (1983 a and b), so a detailed review seems
unnecessary in the present work.  Contributions to the development of computer
codes that predict fracture geometries in three dimensions have been made by
Luiskutty and others (1989), Vandamme and others (1988), Settari (1988), Bouteca
(1988), Acharya (1988), Morita and others (1988), Settari and Cleary (1986), Settari
(1985), Abou-Sayed and others (1984),  Settari and Cleary (1984), Cleary and others
(1983 a and b), Settari and Cleary (1982 a and b), Cleary (1978a), Daneshy (1973).
Carter (in Howard and Fast, 1957) presented the first analysis of leakoff, which was
refined by Williams (1970), and generalized to include nonlinear effects by Settari
(1983).  Analyses of relevant effects, including fluid rheology (Settari and Price,
1984; Cleary, 1980 a and b), back  stress due  to pressure of pore-fluids (Keck and
others, 1984; Cleary, 1980 a and b; Ruina, 1978), heat transfer (Meyer, 1989; Keck
and others, 1984; Griffiths and others, 1983; Settari, 1980), proppant transport
(Settari and Price, 1984; Daneshy, 1978; Novotny, 1977), and fracture toughness
(Theircelin and others, 1989; Morita, and others, 1988; Settari, 1985; van Eekelen,
                                   27

-------
 1980) are commonly included in computer programs analyzing fracture propagation.
 One of those programs, described by Boone and others (1989), represents fracture
 shape, pore-fluid pressure, and stress in the vicinity of a vertical fracture as a color
 image that changes with time on a video monitor.

       Interactions between a hydraulic fracture and the ground surface, a topic of
 particular interest to the proposed applications, have been analyzed by Pollard and
 Holzhausen (1979), and Narendran and Cleary (1983).  In related work, interactions
 between multiple hydraulic fractures have been analyzed by Narendran and Cleary
 (1984), and Hanson and others (1979), and consequences of fracture interactions
 have been utilized in design concepts by Warpinski and Branagan (1989), Ernst
 (1980), and Huck and others (1980). Theoretical solutions of a fracture
 encountering interfaces between materials of differing properties are described by
 Lam and Cleary (1984), Weertman (1980), van Eekelen (1980), and Hanson and
 others (1978 a and b).

       In addition to forecasting the geometry  of hydraulic fracture, the results of
 theoretical analyses are matched against field measurements, such as injection
 pressure or deformation of the ground surface, to estimate geometry after the
 fracture has been created. Methods of inverting pressure records were developed in
 a series of papers by Nolle and Smith (Nolte, 1988 a and b, 1984,1982,1979; Nolte
 and Smith, 1987,1981), and other investigators (e.g. Crockett and others, 1989) have
 made contributions as well. The inversion of measurements of surface
 deformation-comrnonly obtained using tiltmeters-to estimate the location,
 orientation, and dimensions of a hydraulic fracture was  described by Davis (1983),
 and applied by Holzhausen and others (1985 a and b).

       Another application  is the optimization of fracture designs, where the output
 of a fracture simulator is used as the input of a program that evalutes flow to the
 fracture (Anderson and Phillips, 1988) and possibly economic considerations, such
 as costs of fracturing materials and rate of return (Howard and Fast, 1970; Veatch,
 1983 a and b). A simple method of optimization, based on the cumulative volume
 of recovery and described by Elbel and Sookprasong (1987), is relevant to the
 application of hydraulic fracturing during remediation.

       The work cited above includes some notable contributions, but it is by no
 means an exhaustive survey of papers that address useful applications of hydraulic
 fracturing; there are many more. With few exceptions, those papers are targeted to
 applications where hydraulic fractures are created in rock at depths in excess of
 several hundred meters. Use of the significant body of existing information for the
 purposes of this research, therefore, will depend on the  extent to which hydraulic
fracturing of rock at great depth resembles hydraulic fracturing of soil at shallow
depths.


The Hydraulic Fracture as a Problem

       Hydraulic fracturing of soil at shallow depths has, in the past, generally been
regarded as a problem. This is because many soil engineering techniques, which
require increasing ambient pore pressures, will be ruined if they induce hydraulic
fracturing-with the exception of hydraulic fractures created to measure in situ
                                    28

-------
stresses in soil (Leach, 1977; Massarsch and others, 1975; Tavenas and others, 1975;
Bjemim and Anderson, 1972).


Applications

       The collapse of a dam is perhaps the most severe consequence of accidental
hydraulic fracturing. One example occurred at the Teton Dam, Idaho, where
excessive rates of seepage were noted downstream of the dam shortly after the
reservoir behind it was filled in 1976. Seepage rates increased and soon a muddy
flow appeared at the downstream toe of the dam. The flow eroded a gully, which
cut through the dam from the downstream toe upward to the crest. Fourteen people
died and 400 million dollars of property damage resulted from the subsequent
failure. An investigation of the tragedy concluded that hydraulic fracturing, due to
increases in pore pressure accompanying filling of the reservoir, probably
contributed to the dam failure (Jaworski and others, 1979; and references therein).
Elsewhere, hydraulic fracturing was attibuted as the cause of leaks that ultimately
led to the failure of 14 dams in Oklahoma and Mississippi, and other darns in
California, Brazil, and China, according to Sherard (19/2), and Jaworski and others
(1979).

       Ironically, hydraulic fracturing was known to workers in the oil industry as a
problem before it was used to stimulate wells. Yuster and Calhoun (1945)
recognized that a sudden increase  in the rate of inflow during a waterflooding
operation (a technique of injecting water in one well to sweep oil toward a recovery
well) without an increase in pressure could be caused by hydraulic fracturing, or in
their terminology, pressure parting. Hydraulic fracturing caused by excessive
injection pressures can reduce the  area swept out by the waterflood.  New methods
of estimating the maximum allowable pressure without causing fracturing during a
waterflood continue to be developed (Singh and Agarwal, 1990).

       In situ permeability of soil is commonly calculated by holding a constant
pressure and measuring the rate at which water flows into a borehole. Bjerrum and
others (1972) point out that hydraulic fractures can be created if the borehole
pressure used in a permeability test is greater than a critical value, which they relate
theoretically to the overburden load and these soil properties: poisson ratio,
coefficient of lateral pressure, tensile strength, and compressibility. They claim that
if hydraulic fracturing goes undetected dunng a test, values of permeability  •
calculated from the resulting data can be as much as three orders of magnitude
greater than the actual in situ permeability.

       In some applications, hydraulic fracturing can be either an asset or a liability,
depending on the details of the application. The disposal of liquid waste by
injection into wells, outlined in the previous section, is one example.  Injection
grouting is another. Some designs require grout to uniformly permeate pores in the
vicinity of an injection point (borehole), and the specifications of these designs
cannot be met if grout flows preferentially into hydraulic fractures (Wong and
Farmer, 1974).  Moreover, hydraulic fracturing during grouting could result in
problems, such as deforming a neighboring foundation, lifting a concrete cut-off
from its seat, or markedly increasing the amount of grout required to complete a job
(Morgenstern and Vaughn, 1963).  Interestingly, however, hydraulic fracturing can
                                    29

-------
also play a useful role in grouting practices. Zhang (1989) indicates that
intentionally creating hydraulic fractures during grouting increases the rate of
penetration of grout in some formations, and it effectively seals formations by
creating an interlocking network of grout sheets. The technique is especially useful
in formations, such as karstic limestone cut by clay- and sand-filled caves, where
conventional grouting techniques are ineffective. Moreover, the load-bearing
capacity of large diameter piles in silty sand is increased by creating grout-filled
fractures following pile driving, according to Zhang (1989).


Experiments

      The goal of many experimental investigations of the applications described
above is to predict the pressure at which  hydraulic fracturing occurs under a
particular set of laboratory conditions. A noteworthy set of experiments is described
by Jaworski and others (1979 and 1981), who examined hydraulic fractures  in the
vicinities of model wellbores and dam faces. They injected water into cubic blocks
of soil, slowly increasing injection pressure and recording flow rate until a marked
increase in flow rate indicated the onset of fracturing. This technique is common
for  other investigations of hydraulic fracturing of soil, although it differs from
studies of hydraulic fracturing of rock (e.g. Medlin and Masse,  1979; Zoback and
others, 1977; Haimson and Fairhurst, 1967,1970; Harrison and others, 1954; Scott
and others, 1953), which hold flow rate constant and monitor pressure. Jaworski
and others (1979) cut slots in the faces of samples and applied pressure to the slots
to simulate the role of joints or pre-existing cracks as stress concentrators.  Cutting a
slot in a sample to nucleate a fracture is common practice in tests designed to
measure fracture toughness, as mentioned above.

      Significant results of Jaworski and others (1979 and 1981) are as follows:

       1.  The pressure required to induce hydraulic fracturing from a cylindrical
borehole is linearly related to confining stress. A similar relation was observed for
the pressure required to induce hydraulic fracturing in rock by  Medlin and  Masse
(1979), and Haimson and Fairhurst (1967,1970).

      2.  A large amount of variation was observed in the fracturing pressures,
suggesting that fracturing pressure from an open borehole would be difficult to
predict. Fracturing pressure was affected by many factors, such as pre-existing flaws
in the soil, that could neither be eliminated nor characterized.

      3.  The fracturing pressure depended on soil properties, increasing with the
number of blows used to compact the sample. It increased as the water content
decreased.

      4.  A discontinuity, such as a slot, will nucleate a hydraulic fracture in soil,
and the presence of the discontinuity reduces the pressure required to induce
fracturing.

      In a later study, Mori and Tamura (1987) created hydraulic fractures from
cylindrical holes in cylindrical soil samples, using much the same technique as
Jaworski and others (1979).  Their results confirm the linear relation between
                                     30

-------
 fracturing pressure and confining stress.  Mori and Tamura conducted tests using
 various rates of injection, concluding that the fracturing pressure increased as the
 rate of pressurization increased. Hydraulic fracturing pressures in rock also
 increase with rate of pressurization, according to Hairnson and Fairhurst (1967 and
 1970), and Zoback and others (1977).

       In one suite of tests, Mori and Tamura applied an axial load that was less
 than the radial load. Hydraulic fractures were formed either inclined or normal to
 the axial hole, leading Mori and Tamura to conclude that shear failure controls
 hydraulic fracturing of cohesive soil. This conclusion contradicts a large body of
 solid evidence for hydraulic fracturing being a tensile phenomena in other materials.
 It seems likely that the peculiar orientations of fractures observed by Mori and
 Tamura resulted from boundary conditions on their test samples. As Medlin and
 Masse (1979) point out, a cylindrical hole will tend to create a hydraulic fracture
 containing the axis of the hole. The fracture will change orientations, however, as it
 grows away from the hole and into a region of different stresses-such as would have
 been the case under the applied stress conditions used by Mori and Tamura.  Large
 changes in orientations during propagation are accomplished by curving, twisting, or
 breaking into segments (Pollard and others, 1982), resulting in irregular or intricate
 fracture forms. The results that are briefly described by Mori and Tamura are
 consistent with behavior as tensile, or Mode I fractures (Lawn and Wilshaw, 1975),
 propagating in an abruptly changing stress field; their conclusion that hydraulic
 fractures in cohesive soil behave as shear failures is unconvincing.

       An apparatus used to create hydraulic fractures is described by Sun and Ting
 (1988), who claim that it offers advantages over apparatuses described previously.
 Their device is designed to create a hydraulic fracture from a cylindrical hole in a
 cylindrical sample, which is a standard configuration. The principal novelty appears
 to be that the pressure acting on the cylindrical hole can be controlled
 independently from pore pressure in soil adjacent to the hole. This is done by lining
 the hole with a permeable layer and then inserting a flexible bladder along the
 length of the lined hole; inflating the bladder controls pressure on the wall of the
 hole, whereas wetting the liner controls pore pressure in the soil.


Analyses

       Studies seeking methods of preventing fracture initiation commonly employ
analyses of stresses in the vicinity of pressurized holes. Hydraulic fracturing is
assumed to occur when tensile stresses exceeds the tensile strength of the material.
Analyses of this type were first based on elastic solutions and used to predict
hydraulic fracturing of rock (e.g. Harrison and others,  1954; Schiedegger, 1960;
       ' ~>64; Haimson and Fairhurst, 1967),
Kehle, 1964; Haimson and Fairhurst, 1967), but have been applied directly to
predict hydraulic fracturing of soil (Mori and Tamura, 1987; Jaworski and others,
1979 and 1981). Jaworski and others (1979) point out some discrepancies between
theory and data, not the least of which is a large amount of variability in the data.
This observation leads them to conclude that the criteria represented by the simple
theory were insufficient to reliably predict the pressure required to cause fracturing
from an open borehole in soil.
                                 31

-------
       Bjemim and others (1972), who were concerned with inhibiting hydraulic
fracturing during field permeability tests, included in their analysis the effect that
driving a piezometer would have on state of stress prior to fracturing. They assume
that when a piezometer is inserted, the soil adjacent to it yields, and the radial and
circumferential stresses change by an amount that depends on the compressibility of
the soil.  The radial stress always increases when a piezometer is inserted; by a
factor of 1.5 for highly compressible soil to as much as 5.2 for relatively
incompressible soil. The circumferential stress at the wall of the piezometer,
however, will decrease by a factor of as much as 0.6 for highly compressible soil, or
increase by as much as 2.1 for relatively incompressible soil. Deformation resulting
from pushing a piezometer into the ground, thus, could either increase or decrease-
depending on the compressibility of the soil-the pressure required to initiate
fracturing, according to the analysis of Bjemim and others (1972).

       An analysis of the conditions required to create a hydraulic fracture
specifically in unconsolidated sediment is presented by Horsrud and others (1982),
who use coupled elastic and plastic solutions. Fracture initiation pressures can be
reduced by plastic deformation to roughly 10 percent less than that predicted by
elastic theory, according to their results. Medlin and Masse (1979) presented both
experimental and theoretical evidence confirming that plastic deformation around a
borehole reduced the pressure required for fracture initiation.

       Predicting the maximum pressure achievable without causing a hydraulic
fracture during injection grouting has been an elusive task. The limiting pressure is
recognized to increase with depth; values of 22.6 to 102 kPa/m (1.0 to 4.5 psi/ft) of
depth are used in regulations defining maximum allowable pressure (Dickinson,
1988). Those gradients apparently are determined empirically. Morgenstern and
Vaughn (1963) suggest that a Mohr-Coulomb failure criterion could be adopted to
predict hydraulic fracturing during grouting, although the mode of failure-shear-
implied by that criteria is unfounded for hydraulic fractures. The analysis of Wong
and Farmer (1973) assumes tensile failure and uses an elastic analysis, which
includes pore-pressure effects, to examine the initiation of a hydraulic fracture
during grouting. This approach is consistent with established methods (e.g.
Haimson and Fairhurst, 1967), although it is unclear how well it predicts grouting
pressures.

       Wong and Farmer (1973) present both a stress analysis and an energy
balance approach to analyze the propagation of hydraulic fractures formed during
injection grouting. Their analyses assumes that many hydraulic fractures radiate
outward and are closely spaced within a certain distance from the borehole. In
contrast,  the majority of published analyses as well as field and laboratory evidence
suggests that only one, or at most a few, fractures grow away from the borehole.
Wong and Farmer (1973) indicate that their analyses predict pressures during
propagation that greatly exceed those observed in the field.
                                32

-------
                               SECTION TWO
    LABORATORY EXPERIMENTS OF HYDRAULIC FRACTURING OF SOIL

       The purpose of studying hydraulic fracturing of soil has traditionally been to
 provide methods of preventing fracture initiation during geotechnical engineering
 operations. This is because hydraulic fracturing can result in failures during, for
 example, injection grouting, permeability testing, deep-well injection, or dam
 construction. As a result of the focus on fracture initiation little is known about the
 physical appearance, mechanical behavior, or methods of analyzing the growth of a
 hydraulic fracture in soil.

       In contrast, a wide range of useful applications has been developed for
 hydraulic fracturing of rock.  These applications have prompted a broad scope of
 investigations; consequently a great deal is known about hydraulic fractures in rock.

       The goal of the following section is to apply methods of studying hydraulic
 fractures in rock to aid in understanding hydraulic fractures in soil. Laboratory tests
 were conducted using a bench scale apparatus designed to create hydraulic fractures
 in soil. Observations from the laboratory tests, including the appearance of the
 fractures and injection pressures as a function of time and moisture content, are
 described in the following section.

       During the laboratory tests, hydraulic fractures were created by injecting
 fluid into samples confined within a triaxial loading cell. The fractures were
 observed directly by looking through a transparent loading plate in the experimental
 apparatus,  and phenomena related! to fracturing were observed indirectly by
 measuring  the pressure of injected fluid as a function of time. The direct
 observations, combined with descriptions of fracture surfaces from samples split
 open after  testing, yield the important details of the appearance of a hydraulic
 fracture in  clay.  Records of injection pressure as a function of time are used to
 obtain critical pressures at which fracturing takes place. The critical pressure varies
 over an order of magnitude, depending on the moisture content of the clay and
 other factors. Moreover, shapes of the records also depend on moisture content,
 and the nature of this dependency will be described in the following section. Before
 the observations can be described, it will be necessary to explain the design of the
 apparatus,  the techniques of preparing a sample, and the procedures of conducting a
 test.


 EXPERIMENTAL DESIGN

      The experiments were designed to create hydraulic fractures by injecting
glycerin into rectangular blocks of soil confined in a triaxial pressure cell.  An
experimental apparatus and testing procedure were developed to reveal physical
characteristics of the fractures ana to yield data describing the fracture toughness of
soil. Rhodamine dye was added to the glycerin to highlight areas of the fracture
surface wetted by the fluid. The pressure of the injection fluid was monitored as a
function of time so that fracture toughness could be calculated.  Details of the


                                   33

-------
 design of the apparatus and the method of sample preparation are described in the
 following sections.


 Apparatus

       The experimental apparatus consists principally of a pump system, a fracture
 cell, and a data acquisition computer (Fig. 2.1). The pump system is used to inject
 fluid at a constant rate into a sample contained in the fracture cell. Typically, the
 pressure of the injected fluid increases until fracturing occurs, then decreases during
 fracture propagation.  The computer is used to monitor injection fluid pressures as a
 function of time and to control the flow rate of the pump.

       The fracture cell is a rectangular chamber (inside dimensions are 10 cm by
 10 cm by 39 cm) with one moveable side that is used as a loading plate (Fig. 2.2).
 The loading plate is transparent, so that the interior of the cell can be inspected
 during a test. The other five sides of the chamber are lined with neoprene bladders.
 The three principal stresses on the sample are controlled independently by adjusting
 air pressures in the bladders.

       A hole in the loading plate allows access to a soil sample inside the chamber.
 A spacer plate, placed between the loading plate and the sample, contains a
 pressure-tight fitting for tubing three mm in diameter.  The tubing extends from the
 pump through the loading and spacer plates and into a thin slot cut in the sample


       The pump used to inject fluid into the sample consists of two hydraulic
 cylinders dnven by a threaded rod attached to a stepper motor. Hydraulic fluid is
 forced out of the cylinders and into the lower chamber of a pressure interface
 device, which thereby forces dyed glycerin out of an upper chamber and into the
 fracturing cell (Fig. 2.1). The flow rate of the pumping system is controlled by
 regulating the rate of rotation of the stepper motor. Other details of the design and
 operation of the pumping system are given in Murdoch and others (1987).

      A flow rate of 0.225 ml/min was used for all the tests described in the
following pages. The Reynolds Number associated with flow of glycerin in the slot
at that flow rate is 0.01, and in  hydraulic fractures it was less because the aperture
of the fractures was less than that of the slot. Laminar flow was assumed to occur in
the fractures based on the small Reynolds Number.

      Pressure of the injection fluid as a function of time was the primary data
recorded during each test. A transducer, accurate to 0.3 kPa, positioned roughly 20
cm upstream of the slot was used to measure pressure once every 200 msec.
Pressure losses due to flow from the transducer to the tip of the starter slot are on
the order of the accuracy of the transducer, according to calculations using standard
methods (Streeter, 1971), so the pressure measured at  the transducer will be taken
to be the pressure within the slot.

      The injection pressure was recorded as a disk file and  it was displayed as a
function of time on a video monitor. Accordingly, during a test it was possible to
                                  34

-------
 000
  o
                                        000000000
                                        ooooooooo
                                        000000000
Figure 2.1. Apparatus used for hydraulic fracturing experiments. A. Fracturing cell;
         B. Pressure transducer; C. Gauges and regulators of confining
         pressure; D Fluid pressure interface; E. Pump; F. Computer for data
         acquisition and control
                         35

-------
            bladder
          Loading
          Plate
                                to  pump
Figure 2.2. Cut-away sketch of hydraulic fracturing cell.
                            36

-------
observe the pressure record on the monitor, and to see the trace of a hydraulic
fracture on the edge of the sample.


Physical Characteristics of the Soils

      Two soils were used in the laboratory experiments. Most experiments were
conducted using a yellow-brown colluvial clayey silt derived from a pit adjacent to
the Center Hill Research Facility, Cincinnati, Ohio. That soil will be termed the
Center Hill clay. Approximately one half m3 of moist Center Hill clay was
excavated and stored in air-tight drums, which served as a repository of sample
material during testing. Several experiments were conducted using a different soil, a
light-brown alluvial silt obtained from an exposure overlooking the Ohio River near
the River Downs Racetrack, Cincinnati, Ohio. That material will be termed the
River Downs silt.

      All samples were molded into shape for the fracturing experiments. A suite
of standard analyses was conducted on remolded Center Hill clay and River Downs
silt to determine their basic physical characteristics. Two samples of each soil were
analyzed, and they are nearly identical in physical characteristics.

      The Center Hill clay is a type CL soil, according to the USCS classification
based on Atterburg Limits (Table 2.1). It behaves as a plastic material in Atterburg
tests over the range of moisture contents used during the fracturing tests (moisture
contents were between the liquid and plastic limits). The River Downs silt is type
ML, and it shows plastic behavior over a much narrower range of moisture contents
(Table 2.1).  Moisture contents during fracturing tests of the River Downs silt were
less than the plastic limit of that material.

      Analysis of grain sizes indicates that the Center Hill clay is  dprninantly silt
and lesser amounts of clay (Table 2.1; Fig. 2.3).  Trace amounts of limestone
fragments occur in the soil naturally, but they were removed before the soil was
used in the fracturing tests. The  River Downs silt is mostly silt, with minor amounts
of clay.  Siderite concretions occur in the silt. These were removed by sieving.

Results from Proctor Tests (ASTM D698) indicate that the maximum density of the
Center Hill clay occurs when the moisture content (wt. water/wt.solid) is!9.7%.
The moisture contents of samples used in the fracturing tests ranged from 19 to
33%; typically greater than optimum moisture content. Bulk densities of samples
used in the fracturing tests are either equivalent to or slightly greater than densities
from the Proctor Tests (Fig. 2.4).


Sample Preparation
      Techniques of preparing soil samples were developed to yie
rectangular samples of a desired composition, water content and c<
                                                          ield uniform
                                                          [consolidation
history. Samples were prepared in rectangular molds the same size as the inside of
the fracturing cell.  A mold resembled a rigid box with a top that moved like a piston
to consolidate the sample.  Geotextile along two sides of the sample facilitated
drainage during consolidation.
                                      37

-------
TABLE 2.1. CHARACTERISTICS OF SOILS USED IN THE STUDY
 CENTER HILL CLAY

 Atterburg Limits                CHI        CH2       AVE
 (wt. water/wt. solid)
      Liquid Limit            0.429        0.438        0.433
      Plastic Limit            0.198        0.200        0.199
      Plastic Index            0.231        0.238        0.234
      Shinkage Limit          0.188

 Grain size
      Gravel                 0           0
      Sand                   0.03         0.03
      Silt                    0.61         0.62
      Clay                   0.36         0.35

 Proctor Test (ASTM D698)
 Moisture Content of Greatest Density: 0.197
 Maximum Dry Density: 1.68 gm/cm3 (104.7 lb/ft3)
 Maximum Wet Density: 2.01 gm/cm3 (126.0 lb/ft3)

 RIVER DOWNS SILT

 Atterburg Limits                RD1        RD2       AVE

     Liquid Limit             0.221        0.234        0.227
     Plastic Limit             0.208        0.200        0.204
     Plastic Index             0.013        0.034        0.023

 Grain size
     Gravel                   0.02         0.0
     Sand                    0.05         0.06
     Silt                     0.81         0.82
     Clay                    0.12         0.12

Proctor Test (ASTM D698)
Moisture Content of Greatest Density: 0.165
Maximum Dry Density: 1.73 gm/cm3 (108.2 lb/ft3)
Maximum Wet Density: 2.06 gm/cm3 (128.7 lb/ft3)
                               38

-------
I
£
h
~
«
0
cm

1UU
90
80
70
60
50
40
30
20

10
ft
& Center Hill clay
J*
i
	 r^
•



                 0.001
0.010         0.100
     Grain size (mm)
                                             1.000
            A
             ao

100
 90
 80
 70
 60
 50
 40
 30
 20
 10
                      River Downs silt
                0.001
0.010         0.100
     Grain size (mm)
                                             1.000
Figure 23. Grain-size distribution of soils used in the lab experiments.
                                    39

-------
              o
              o
                                                           • Proctor test

                                                           O Fracture test
2.05
                 1.95
                 1.85
                 1.75

                   0.16    0.20   0.22   0.24   0.26   0.28    0.30   0.32


                                 Moisture (wt. waler/wt. solid)



                 2.10
             a  2.05
                2.00


             i
             Q
             11.
                          River Downs silt
                                                            Proctor test

                                                            Fracture test
 95
                1.90 ' • •  • •
                   0.14  0.15   0.16   0.17   0.18   0.19  0.20   0.21   0.22



                                 Moisture (wt. water/wt. solid)
Figure 2.4. Bulk density as a function of moisture content from Proctor tests and

              fracture tests.
                                     40

-------
       Two methods were used to prepare a sample; one that involved compaction
and consolidation, and another that involved compaction alone. When samples
were to be consolidated, a layer of soil roughly 2 cm thick was compacted with 30 to
40 blows of a drop-weight hammer. The upper surface of the layer was scarified to a
depth of 0.5 to 1.0 cm and another layer of similar thickness was added and
compacted. That process was repeated until the soil in the mold was 10.5 cm thick.
Four sample molds were filled and then placed beneath the corners of a pallet upon
which there were four drums of water. The weight of water in the drums exerted a
vertical stress of 69 kPa on the samples.  Samples were consolidated under that load
for as many as 14 days.

       A pneumatic press (SoilTest Model CN-425A) was used to load samples that
were prepared only by compaction. The press was modified by attaching a special
rectangular load shoe (5.8 cm x 9.8), so that the entire width of the sample mold was
loaded by each blow. A layer of soil 2 cm thick was compacted by 80 blows from the
press, and a pressure of 139 kPa was applied during each blow.  The layer was
scarified and another layer added and compacted. That process was repeated until
the sample mold was filled to a depth of roughly 12 cm. The sample was removed
from the mold and trimmed with a wire saw to a height of 10 cm.

       A narrow slot (0.04 mm in aperture) was cut through the middle of each
sample (Fig. 2.2) using a special blade-like tool.  The slots were rectangular in shape
with the long axis of the rectangle spanning the width of the sample. The short axis
of the rectangle, or slot length, ranged from 12 to 72 mm depending on the size of
the blade used to cut the slot. A hole 3 mm in diameter was cut along the center of
the long axis.

       The purpose of the slot was to provide a starting fracture that was much
larger than existing flaws in the sample. The slot was necessary because
measurements of critical stress intensity require knowing the length of a fracture
when it begins to propagate (Tada and others, 1985). Natural discontinuities
resembling fractures several mm long were common in the samples, so a starter slot
of at least 12 mm in length  was used to nucleate hydraulic fractures.

       A film of silicone grease was applied to the surfaces of the sample to inhibit
leaking of the injected fluid where a fracture intersected the sample surface.  The
grease provided an adequate seal and, because it was nearly transparent, allowed
the trace of the fracture on the surface of the sample to be observed through the
plexiglass loading plate.

       The starter slot was filled with glycerin  prior to each test to ensure that air
was expelled from the slot. The loading plate was then secured and the cell
positioned so that the starter slot was horizontal, eliminating a static pressure
gradient normal to the slot (Fig. 2.2).
      Confining pressures were applied to the sample by inflating neoprene
bladders in the fracturing cell with air. Three sets of pressure gauges and regulators
provided independent control of each of the three confining stresses. Most tests
were conducted using confining pressures between 35 and 70 kPa (5 to 10 psi).
Typically, the confining pressure normal to the starter slot was 20 percent less than
the other two confining pressures, which were equal to one another. This loading
                                41

-------
configuration was used to promote the growth of a hydraulic fracture in the plane of
the starter slot. If all three confining loads were equal or if the minimum load was
in the plane of the starter slot, the hydraulic fracture would twist or curve yielding
complicated forms that were difficult to duplicate.


A Typical Test

       Events occurring during the fracturing tests followed a consistent pattern.
 ..„ r  illy, pressure increased nearly linearly with time early in the test. At some
point the slope of the pressure record began to flatten, reaching zero slope as the
pressure peaked and then becoming negative as the pressure decreased with
continued injection.

       A thin (on the order of 0.05 mm) fracture trace typically could be first seen
on the surface of the sample roughly at the time of maximum injection pressure.
Most traces were nearly straight, although in many cases they consisted of a family
of straight, sub-parallel segments arranged either en echelon or staggered. The
location of the fracture tip was difficult to establish because the aperture tapered
gradually until it became undetectable. Thus the tip could be located within
approximately one cm, but the trace was too thin to locate the tip precisely.

       The fracture trace grew at cm-long intervals, initially appearing as lines so
thin they could barely be seen and then slowly widening to narrow slits with a
recognizable aperture. Growth appeared essentially the same when magnified
roughly ten times; greater magnification was impossible using available instruments.
Although it was impossible to follow the tip of the fracture trace during propagation,
the average rate of propagation could be determined from the total fracture length
and the total time of injection; it was on the order of 0.5 to 1.0 mm/sec.  Medlin and
Masse (1984; fig. 9), who used an apparatus similar to the one described above and
injected at 0.5 ml/sec  (roughly twice the flow rate used here), report that the
propagation rate decreases with time from 2 mm/sec early in a test to 0.2 mm/sec
late in the test.

       A fracturing test was terminated by stopping the pump and opening a
pressure relief valve, which allowed the injection fluid to flow back into the injection
tube as the fracture  closed. The dyed glycerin was promptly removed from the
fracture to inhibit staining of the sample by flow unrelated to the process of
hydraulic fracturing.

       The sample was removed from the fracturing cell and pulled open along the
plane of the fracture.  A rough, finely-dimpled surface was formed where pristine
portions of the samples were pulled apart.  In contrast,  the surface of the hydraulic
fracture was relatively flat, or marked by angular steps. This contrast is important
because the leading edge of the hydraulic fracture was undyed, but could be readily
distinguished from pristine clay by the difference in textures.
                                     42

-------
 APPEARANCE OF A HYDRAULIC FRACTURE IN CENTER HILL CLAY

       More than five dozen experiments were conducted with the Center Hill clay
 using a variety of loading conditions, sample preparation techniques, moisture
 contents and durations of consolidation. Hydraulic fractures were created during
 every experiment, except in a few cases when the injection tube was plugged with
 clay as it was inserted.  Examination and description of the appearance of each
 fracture revealed features of the fractures that occur in virtually every case.  Most of
 those features exhibited slight variations, depending on the conditions of the test,
 but they were consistently identified in some form. The features were described in
 notes and they were traced onto transparent mylar. In the following section, the
 typical attributes will be used to describe an idealized hydraulic fracture in the
 Center Hill clay.

       In gross form most fractures were nearly symmetric with respect to the axis of
 the starter slot. The fractures created during the toughness experiments were
 roughly planar; other experiments resulted in fractures that curved or twisted into
 complicated forms that will be omitted from the following descriptions. The half-
 length of a fracture (measured from the hole along the axis of the slot to the leading
 edge) was roughly uniform along the width (measured parallel to the axis of the
 starter slot), although half-lengths were slightly shorter near the edges than in the
 center of most fractures.

       Some fractures were asymmetric with respect to their width, that is their
 lengths were several cm greater along one side than along the other. In some cases,
 the lengths were greater on the side of the injection tube, whereas in other cases the
 side opposite the injection tube was longer.  In virtually all cases, however, the
 asymmetry was mirrored across  the axis of the starter slot. Slight leakage at one
 side of the sample, either out of the slot or out of the fracture itself, apparently
 caused this type of asymmetry. Leakage in most instances occurred during
  Eropagation, according to direct observations through  the loading plate, so it had
  ttle effect on the early stages of nucleation and growth of the fracture.

       The typical test revealed  a continuous, parent fracture adjacent to the starter
 slot. The parent fracture broke into discontinuous, lobate planes with increasing
 distance from the injection hole. Dye staining formed irregular dendritic patterns
 near the ends of the lobes, but the leading edge of the  fracture was beyond the zone
 reached by the dyed glycerin and is unstained.  These features define four distinct
 zones,  arranged in increasing distance from injection hole: 1.) starter slot; 2.) parent
 fracture; 3.) fracture lobes; 4.) and an unwetted zone at the leading edge.


 Starter Slot

       The starter slot appears on the surface of a cleaved sample as a smooth strip
with a shallow trough, which is one half of the injection hole, along the axis of the
strip (Fig. 2.5). The width of the strip depends on the length of the blade used to cut
the slot.
                                  43

-------
                                                               J_aJ
                       0
10cm
Figure 2.5. Photograph and sketch of the surface of a hydraulic fracture, a.) starter
             slot; b.) parent fracture; c.) lobes; d.) unwetted tip. Lines on the
             sketch are prominent linear features on the fracture surface. Heavy
             lines represent overlap of fracture lobes, as shown in the inset.

                                  44

-------
 Parent Fracture
       The main, or parent hydraulic fracture appears as a continuous dyed surface
 that lies roughly within the plane defined by the starter slot (Fig. 2.5). Locally,
 parent fractures are twisted or curved, with most of the distorted regions located
 near the edges of a sample (top of Fig. 2.5). Those distortions are minor, however;
 the fracture rarely deviates by more than 10° from the orientation of the slot.
       The surface of a parent fracture is generally flat, but in detail it is marked by
 irregularities that have as much as one cm of relief. Some of the irregularities are
 rounded bulges or dimples.  More typically the surface is marked by elongate
 features; narrow ridges, grooves or angular steps (fig. 2.5).  Ridges and grooves
 typically have one mm or less of relief and the fracture surface on either side of
 them lies in the same plane. Steps are where the fracture surface abruptly changes
 height over a narrow interval.

       Steps range in size from less than one mm high and several mm long, to 5
 mm high and nearly as long as the entire fracture (Fig. 2.6). Lengths of ridges and
 grooves are on the order of several mm, typically shorter than the steps. The
 direction of the long axes of the linear features differ, depending on where the
 feature occurs on the fracture surface. Where they occur near the starter slot, the
 long axes are perpendicular 19 the axis of slot. At the other end of the fracture, near
 the tip, the steps are perpendicular to a tangent across the leading edge.  Within the
 parent fracture, the axes are sub-parallel to one another and within 20° of the line
 formed by the edge of the sample. As the edge of the sample is approached,
 however, the linear features curve outward at increasingly higher angles, and at the
 edge itself the axes are inclined 45° or more (Fig. 2.6).

       Steps, grooves and ridges on fracture surfaces are by no means unique to
 hydraulic fractures in soil; they have been described on the surfaces of tensile
 fractures in a variety of materials, ranging from glass to metal, plastic or rock
 (Williams, 1984; Pollard and others, 1982, and references cited therein; Pollard,
 1978; Pollard and others, 1975; Daneshy,  1973; Tetelman and McEvily, 1967). A
 common interpretation is that the steps indicate the direction of propagation when
 the leading edge of the fracture is at the location of the step (Pollard and others,


      This interpretation is consistent with the patterns of steps on the surfaces of
 fractures in soil. The typical pattern of steps is shown in Figure (2.7), from which
 the leading edge at several different times during growth is inferred. The inference
 suggests that propagation began near the center of the slot and the fracture grew
 toward the edges as it increased in length  (Fig. 2.7). Upon reaching the edge of the
 sample, the fracture grew roughly parallel to it. Propagation at the edges of the
 sample apparently lagged behind the center because the direction of growth there
 has an outward component. Thus, the fracture trace on the surface of a sample,
 viewed through the plexiglass loading plate, was oblique to the direction of
 propagation of the fracture contained within the sample. In some cases, as in Figure
 2.5, linear features suggest that propagation began near one side of the sample.

      Directions of propagation were slightly divergent within the fracture, and
strongly divergent near the edges.  As a result, all except the center-most of the
                                 45

-------
                                    5
                                    cm
10
Figure 2.6. Surfaces of three hydraulic fractures of various lengths.  Lines on
             fracture surfaces are linear features and hatch marks on the lines
             indicate the lower side of a step. Fractured, but undyed areas are
             stipled. Heavy lines indicate overlap of lobes, as in Fig. 15.
                                   46

-------
Figure 2.7. Idealized diagram of the position of the leading edge and the path of
             propagation during growth of a hydraulic fracture in the experiments.
             Inferred from the results of several dozen tests.
                                   47

-------
propagation paths eventually fan outward to intersect the edges of the sample (Fig.


       A few steps on fracture surfaces were nearly parallel to the axis of the starter
fracture-at a high angle to the steps that indicate propagation direction (Fig. 2.6a
and b). The axes of the high-angle steps are several cm in length and they are nearly
straight to gently arcuate.  Their location is variable; some of them occur in the
vicinity of the starter slot, whereas others are in the vicinity of the leading edge.  In
many cases, a high-angle step on a fracture surface is related to a smaller fracture
that either overlies or underlies the surface. The origin of the high-angle steps is
unknown, but they are similar to rib marks, which were observed by Daneshy (1973),
who used them to infer the position of the leading edge of the hydraulic fracture
prior to the termination of a test.


Lobes

       A continuous parent fracture breaks  into a family of fracture lobes as the
leading edge is approached from the starter slot (Fig. 2.6 and 2.7). The transition
from continuous fracture to discontinuous lobes is smooth, marked by slight twisting
or curving of lobes with respect  to the parent fracture. Lobes occurred in virtually
all the tests, although shapes and sizes of the lobes differ markedly between
samples.

       Lobes range from nearly equant to highly elongate, with aspect ratios from
1:1 to 1:10. In a few cases, fracture lobes originated at the starter slot and the entire
fracture consisted of a family of elongate lobes; a continuous parent fracture was
absent (upper right side of rig. 2.6c). The major axes of elongate lobes are within a
few degrees of parallel to the edges of the sample.

       Fine steps or ridges are present on the surfaces of fracture lobes. Axes of
steps are parallel to the major axis near the  center of a lobe, but curve outward as
the leading edge is approached.  At the leading edge of a lobe, the steps are nearly
perpendicular to the edge. The  elongate lobes, therefore, are inferred to have
widened as they increased in length.

       As many as several dozen lobes are present in any given sample, and they
range in size over more than an  order of magnitude.  The width of the largest lobe is
typically on the order of several  cm, roughly one third of the width of the sample.
Intermediate-sized lobes, whose widths are on the order of several mm, occur along
the leading edges of the larger lobes. Other  lobes that are even smaller, roughly one
mm in width, fringe the edges of the intermediate-sized lobes.

       Several of the larger lobes, the ones bounded by unfractured soil, occur in
each sample.  Each one of these lobes is in close proximity to one or more
neighbors. Three types of spatial arrangement between neighboring lobes were
identified during the tests. They are as follows (Fig. 2.8):


    1. Neighboring: Lobes are coplanar, and their edges are separated by a narrow
        band of unfractured material.
                                    48

-------
Figure 2.8. Idealized configurations of commonly-occurring fracture lobes, a.
             Neighboring lobes; b. En echelon overlapping lobes; c. Staggered
             overlapping lobes; d. Superposed lobes.
                                  49

-------
    2. Overlapping: Lobes lie in slightly different planes, and their edges slightly
        overlap.

    3. Superposed: Lobes are roughly parallel. When viewed normal to the plane of
        the fracture, one lobe entirely overlaps the adjacent lobe.

       Neighboring and overlapping lobes are the most common, occurring in nearly
all samples. Typically, the wider lobes tend to be overlapping whereas the smaller
ones are neighboring. The sense of overlap between lobes was consistent in some
samples, so that traces of the lobes would appear en echelon on a surface normal to
the long axis of the sample. In other cases, however, the sense of overlapping
reversed itself, so that traces of lobes would appear staggered on the normal surface.
Typically, the sense of overlap was inconsistent, with some lobes en echelon and
others staggered in any given sample (Fig. 2.6).

       Superposed lobes, or pair of lobes with a large amount of overlap, occurred
most commonly when the magnitudes of the applied stresses differed by a relatively
large amount. Superposed lobes were rare in tests where the applied load normal to
the starter slot was 80 percent of the loads parallel to the slot (the relative loading
typical of all tests described in Section 3).  Highly overlapping lobes were more
common, occurring in roughly half the samples, when the normal load was between
30 and SO percent of the parallel load.

       Steps on the surfaces of superposed lobes were used to infer the path taken
by the lobe as it grew, and  some of the results are rather surprising. Typically,
superposed lobes appear to be overlapping lobes that widen laterally as they grow in
length (Fig. 2.8d). In some cases, however, the steps indicate that the lobe widened
laterally and then propagated back toward the starter slot. In a few cases,
superposed lobes are roughly circular features that are nearly isolated from both
other lobes and the parent fracture. Small  fractures, several mm in length, oriented
normal to the lobes are the only connection that could be identified between those
superposed lobes and their neighbors. Steps indicate that these isolated fractures
grew outward in a radial pattern from the small connecting fractures.

      The axes of boundaries between adjacent lobes typically can be traced back
toward the starter slot to steps on the surface of a fracture (Fig. 2.6). This was
observed on several scales; from boundaries between the larger lobes which can be
traced to steps on the surface of the parent fracture, to the boundaries between the
smallest lobes which can be traced back to fine steps on a fracture surface.

      Pollard (1978) has argued that linear features, such as steps, ridges or
grooves, on the surface  of a tensile fracture result from the coalescence of lobes
during growth of the fracture. The close spatial relationship between the
boundaries between lobes and the axes of the linear features in the experiments
described  above support Pollard's argument.  Moreover, Pollard cites the
development of lobes, steps, ridges and grooves over a wide range of scales (eight
orders of magnitude) and in materials ranging from glass to metal and rock. The
results of this study indicate that the appearance of hydraulic fractures in clay-rich
soil is similar to hydraulic,  or other tensile, fractures in general.
                                   50

-------
Unwetted Tip

       A thin band of fractured, but undyed soil ocurred at the leading edge of many
of the hydraulic fractures.  This undyed zone was identified on the surfaces of split
samples based on surface texture and color. The zone lacked dye staining, so it
could readily be distinguished from the surface wetted by rhodamine-laced
glycerine. The break created when the sample was pulled apart had a rough,
dimpled surface and lacked linear features such as steps or ridges (Fig. 2.9). In
contrast, the surface in the undyed zone was smooth and nearly planar, marked only
by fine steps. In some cases, a fine step in the undyed zone could be traced back
into the stained region of the hydraulic fracture (Fig. 2.9).

       The similarity in surface textures suggests that the undyed zone represents
part of the hydraulic fracture that has not been wetted by the injection fluid, and
unwetted tip.  Accordingly, the hydraulic fractures extended from the bole along the
axis of the starter slot to the leading edge of the undyed zone.

       The length of the undyed zones varied greatly from one sample to another.
In general among samples of similar moisture content and consolidation history, the
length of the  undyed zone was roughly proportional to the length of the dyed zone
(Fig. 2.10). This suggests that those lengths are related by


                              Luw = mLw + Lw                          (2.1)

where m is the slope dLuvt/dL>, and L^ is the intercept. First order regression lines
are shown in  Figure 2.6 for measurements taken from samples of different moisture
content.

       The length of the unwetted tip relative to the length of the wetted fracture,
the parameter m in eq. (2.1), depends on moisture content of the sample (Fig. 2.10).
Rhodamine stain was present over the entire fracture surface-an unwetted tip was
completely absent-for samples containing less 'than 21% moisture. The relative
length of the  unwetted tip m increases as moisture content increases from 21 to
28%.  Within that range, the relative length m is roughly linearly related to moisture
content, according to Figure 2.10. The relative length is greatest, roughly 25%, at
moisture contents between 27 and 28%, and it diminishes as moisture content
increases to values greater than 28% (Fig. 2.10). The trend at moistures greater
than 28% is based on only a few measurements.

      The measurements of undyed zones made on different fracture surfaces and
compiled in Figure 2.10 could be used to infer different stages in the growth of the
unwetted tip of a single, idealized fracture. We must make this inferrence with
caution, however, because it assumes that the glycerin stops moving when
propagation ceases. That assumption requires that capillary forces are unable to
cause flow after the pump is turned off. To crudely examine the effects of capillary
forces, droplets of dyed glycerin were  placed on narrow, flat-lying cracks exposed on
the surfaces of samples of Center Hill clay. The glycerin flowed 6 to 12 mm into the
dry cracks due solely to capillary suction. The dye produced dendritic patterns
similar to those at the leading edges of the hydraulic fractures.
                                    51

-------
                0
5cm
Figure 2.9. Leading edge of a hydraulic fracture. Lines on the sketch indicate linear
             features as in Figures. 2.5 and 2.6.

                                  52

-------
       2.5
         0.0        2.5       5.0       7.5      10.0

                         Dyed Length (cm)
12.5

0)
O
Q)
73
CJ

0.25
0.20
0.15
0.10
0.05
o.oof
073
O
0 :
o
0 o .
) O O •-•--••-•-•- • • :
'a ^^ oVo 0.22 0.24 0.28 0.28 0.30
Moisture Content
Figure 2.10. Length of undyed zone as functions of dyed length and moisture
             content (upper figure). Ratio of undyed to dyed length as a function
             of moisture content (lower figure).
                               53

-------
      Capillary forces will certainly affect flow in the vicinity of an unwetted tip
during propagation, and it seems likely that those forces could cause the glycerin to
advance toward the tip even after propagation has ceased.  We conclude that the
lengths of undyed zones shown in Figure 2.6 are the lower limits of the lengths of
the unwetted tips during propagation. Likewise, we must accept the possibility that
some of the variation in m with moisture content described above could be
influenced by changes in post-propagation flow, rather than by changes in the rate of
growth of the unwetted tip itself.


RECORDS OF DRIVING PRESSURE

      Driving pressure Pa, the difference between the internal fluid pressure and
the confining stress normal to a fracture, is a fundamental quantity in analyses of  a
hydraulic fracture (Kristianovich and Zheltov, 1955; Perkins and Kern, 1961;
Geerstma and de Klerk, 1969; Pollard, 1973, Nolle and Smith, 1981; Crockett and
others, 1989). As such, Fd is a valuable quantity in the diagnosis of fracture
behaviour and it is routinely recorded as a function of time during both laboratory
experiments and commercial applications in the field (Nolle, 1988 a and b; Medlin
and Masse, 1984; Zoback and Pollard, 1978; Zoback and others, 1977; Hubbert and
Willis, 1957). Records of driving pressure were made during all the experiments
conducted during this research to compare with similar records from experiments in
rock and to examine how various parameters affect the record. During the
experiments, driving pressure within the starter slot was obtained by measuring the
pressure of the injection fluid as  it entered the fracturing cell and subtracting the
pressure in the pneumatic bladder acting normal to the slot

      The records of driving pressure as  a function of time, hereafter termed
records, were obtained for a variety of conditions. They will be described in the
following section.


Reproducing the Records

      Early in the experimental program a suite of four tests was conducted to
examine whether the records could be reproduced. Four samples were prepared
from a common batch of soil and consolidated together beneath the loading pallet
described above. The properties of the samples were similar: densities were within
0.01  gm/cm3 and moisture contents differed by less than one percent.  Slot lengths,
applied loads and other factors were identical.

      Forms of the records for the four samples (Fig. 2.11) are similar,
characterized by the following periods:

      Period I: Nearly constant positive slope

       Period II: Slope diminishes, but remains positive

      Period III: Slope is negative
                                  54

-------
   ctf
           0      50     100    150    200
                   Time  (sec)
Figure 2.11. Records of driving pressure as a function of time from tests conducted
         on similar samples.
                         55

-------
       The records are nearly identical during Period I. The boundary between
Period I to Period II was identified by marking the location of the break in slope
typical of Period I. Three of the records in Figure 2.11. show smooth changes m
slope, whereas one of the records shows an abrupt step in the record. Pressures
marking the boundaries between periods are indicated by short markers on the
records.

       The driving pressures marking the change in slope are within 1.0 kPa for
three of the samples, and they range over 4.1 kPa, from 22.8 to 26.8 kPa, for all four
samples. The driving pressures average 25.4 kPa and show a standard deviation of
1.86 kPa, which is within seven percent of the average.

       The forms during Period II are similar for three of the records, showing a
gradual decrease in slope. The other record is marked by two short intervals during
which the slope changes drastically as the pressure decreases and then increases
abruptly. That record is similar to the others, however, if the two short intervals are
ignored.  Slight, abrupt changes in pressure were observed in other records, but their
occurrence was unpredictable.

       The maximum driving pressure marks the boundary between Periods II and
HI, and it ranges over 3.7 kPa (from 27.0 to 30.7 kPa), roughly the same range as the
boundaries between Period I and II. The record showing the least pressure between
Periods I and n yields the greatest pressure between Periods n and HI. The average
maximum pressure is 28.8 kPa,  and the standard deviation is 2.03 kPa.

       During Period III, driving pressure decreases and approaches a constant
slope that is similar for three of the records.  A constant slope is approached by the
record of the fourth sample, but it is steeper than the other three.  Slopes of two of
the records diminish with continued propagation during Period III.

       The records appear to be reproducible in both form and magnitude up to the
time of maximum pressure.  Variations in the pressures markingthe boundaries of
the Periods are within 10 percent of the total driving pressure. The records during
Period m are similar in form, but their magnitude at any given time ranges over 10
kPa or more.

       The forms of the records described above resemble the records presented by
Medlin and Masse (1984) and Daneshy (1976a), both of whom used apparatus
similar to the one used here but conducted their experiments using limestone or
sandstone.  The principal difference between the records from tests where rock is
fractured and those in Fig. 2.11  is the slope early in Period III; it is generally steeper
when rock is fractured. However, considerable differences are seen in the forms of
records of hydraulic fracturing tests using rock, and some of them are nearly
identical to those in Figure 2.11 (e.g. Daneshy, 1976a; fig. 6).

       Previous studies of hydraulic fracturing of soil (Mori and Tamura, 1987;
Jaworski and others, 1981; Jaworski and others,  1979; Bjerrum and others, 1972)
have measured the rate of flow  into a sample while keeping pressure constant, and
then slightly increasing injection pressure until fracturing occurs.  The records of
those studies are typically presented as injection pressure plotted with respect to
                                 56

-------
flow rate, and they are difficult to compare to the records from tests presented here
where injection rate was held constant.


Fracture Development with Respect to the Record

       Most tests were terminated sometime during Period III, but several tests
were terminated during Periods I and II in an effort to correlate the development of
a fracture to the form of the record. Samples similar to the ones described in the
previous paragraph were prepared so  that the pressures marking the boundaries
between Periods could be anticipated. The surfaces of fractures terminated at
different times are shown in Figure 2.6a through c. The times when each of those
tests was terminated are shown on an  idealized record (Fig. 2.12).

       Most tests terminated during Period I showed a starter slot enveloped by
unfractured soil; a hydraulic fracture was absent. In two cases, small hydraulic
fractures were observed even though the tests were terminated prior to a break in
slope in the pressure record. These incipient fractures were less than 1.0 cm longer
than the starter slot, and short unwetted zones occurred at their leading edge.

       Tests terminated during Period II, however, show well-developed hydraulic
fractures several cm in maximum length. The edges of the fracture during Period II
typically are contained within the sample, that is the fracture has not yet reached the
edge of the sample (Fig. 2.6a).

       Another test was terminated early in Period III, roughly five seconds after
the maximum pressure. The sample from this test contains a fracture that is longer
than the one terminated during Period II, and it has intersected the edges of the
sample (Fig. 2.6b).

       The third sample shown in Figure 2.6c was terminated several tens of
seconds into Period III and it shows a fully-developed fracture cutting from one
edge of the sample to another.

       Those observations indicate that the break in slope of the record (Fig. 2.11)
indicates that hydraulic fracturing is taking place. In some cases, incipient fracturing
apparently occurs prior to the break in slope.  Incipient fracturing cannot be
detected on the pressure records, at least by the methods of recording pressure used
for this work, so we will assume that fully-developed fracturing starts at the break in
slope. The driving pressure at the break in slope Pa is taken as the critical driving
pressure required to initiate fracturing.

       Following the onset of fracturing, driving pressure continues to increase
during Period II and then decrease dunng Period III. Increasing driving pressure
during propagation indicates a period of stable growth of a hydraulic fracture,
whereas decreasing pressures indicate unstable growth (Tetelman and McEvily,
1967; Zoback and others, 1977; and Zoback and Pollard, 1978). These terms are
used because if the driving pressure were held constant a stable fracture would be
held open but would not propagate, whereas an unstable fracture would continue to
propagate at constant pressure. Thus, the three Periods of the injection record
relate to three different processes during a test:
                                  57

-------
                e
                o
                d-
                           c
                           o
                           d
                         (DQO
  e
  o
               — o
do

(00.1
— d
J3CID
dd
**Ou
wo
et-
                          Time
Figure 2.12. Idealized record of driving pressure as a function of time. Letters
            correspond to times of termination of tests resulting in fractures
            shown in Figure 2.6a, b, and c.
                              58

-------
      Period I:  Inflation of the starter slot

      Period II: Stable propagation of a hydraulic fracture

      Period III: Unstable propagation

      In an earlier section linear features were used to infer that the fracture
initiated within the sample and then grew outward to the edges as it increased in
length. This interpretation is supported by the images shown in Figure 2.6, and
suggests that stable growth occurs from the time when the fracture initiates to when
it intersects the edges of the sample.

      Williams (1984) presents one explanation relating this behavior to the
geometry of the specimen. He shows that a notched rectangular specimen will be
under conditions of plane strain at its center, whereas it willbe under plane stress
near its edges. The resistance of a material to fracturing in plane strain is less than
in plane stress, because conditions of plane strain reduce plastic deformation at the
crack tip (Williams, 1984). According to this explanation, the increase in driving
pressure during Period II occurs as the fracture grows from the center of the sample
into material offering greater resistance to propagation. Williams' arguments are
based on a sample whose outer surface is unconfined, whereas the samples used in
this study were confined in a triaxial pressure cell. Friction between the outer
surfaces of a sample and the walls of the  fracturing cell would further inhibit
propagation as a fracture grew toward  the edge of the sample.

      Zoback and Pollard (1978) show that propagation of a hydraulic fracture can
be initially stable, regardless of the geometry of the sample. They attribute this
stability to viscous losses in driving pressure in the incipient fracture. It is possible
that a similar mechanism contributes to the stable propagation during Penod n.


Effects of Moisture Content

      Suites of samples were prepared by placing soil of differing moisture content
into the sample molds. Soil in the molds was consolidated for 10 to 14 days, until
displacement during loading was negligible.  The resulting samples differed in
moisture content and bulk density, but were otherwise similar.  It is possible that
moisture content varied slightly within .the samples, due to incomplete equilibration.
Drainage during consolidation was only possible through geotextiles on the two
sides of the sample that were parallel to the starter slot, however, so that moisture
contents would be uniform along the planes of the fractures. Fractures were created
in the samples and the results of one suite of tests, which is representative, is shown
in Figure 2.13.

      The form of the record of injection pressure is markedly affected by
increasing the moisture content of a sample.  During Period I, it causes a slight
decrease in the slope, but the  most striking effects are seen during Periods II and HI.
The pressure marking the transition between Periods I and II diminishes as
moisture content increases. Furthermore, the slopes of both Periods II and ID
become more gentle, resulting in a flattening of the curved part of the record.
                                  59

-------
        100  -
     ctf
    PH
             0        50      100      150
                        Time  (sec)
Figure 2.13. Records of injection pressure as a function of time for samples of
          various moisture contents.
                             60

-------
       Driving pressures required to initiate hydraulic fracturing decrease by an
 order of magnitude, from more than 100 kPa to 10 kPa, by increasing the moisture
 content of the samples by only six percent.  Once propagation begins, theperiod of
 stable growth, Period n, becomes longer as moisture contents increase. The
 transition from stable to unstable fracture growth, indicated by a change from
 positive to negative slope on the record, is well-defined for relatively dry samples
 but it becomes subtle for relatively wet samples.  Despite this subtlety, however,
 periods of both stable and unstable growth could be identified in all tests.

       Explanation of the effects of moisture content on the fracture process will
 require development of several theoretical models in Section 3, but an explanation
 of how the slope dPa /dt differs with moisture content during Period I can be readily
 explained. The starter slot is inflating during Period I and the different slopes
 appear to be related to an increase in modulus as moisture content decreases.
 According to Tada (1985) the volume of a pressurized slot in plane strain is


                             V = 2-ffPAwai(l -v*)/£:                      (2.2)
where w is fracture width, a is the fracture half length, v is Poissons ratio, and £ is
the elastic modulus. Differentiating with respect to time and assuming1 that


                                  dV=Qdt                            (2.3)

where Q is the rate of injection yields


                         d?d /dt = QE/[2* fl2w(l - v2)]                    (2.4)

neglecting the effect of the small hole along the axis of the slot. The slope during
Period I is linearly related to elastic modulus of the sample, according to eq. (2.4).
Uniaxial compression tests conducted on the clay show that the elastic modulus
decreases by at least an order of magnitude over the range of water contents of the
samples. We conclude that a decrease of elastic modulus could cause the observed
change in slopes of the records.  The changes in slope, however, are small compared
with the change in modulus. This probably results from volume changes in tubing
and the pressure interface of the pumping system, which contribute to the slope
during Period I but are overlooked in eqs. (2.3 and 2.4).


Effect of Slot Length

       The effect of varying the length of the starter slot is summarized in Figure
2.14, which shows the records from four experiments conducted on similar samples
  This assumption requires that none of the glycerin leaked out of the fracture into
    the clay.
                                       61

-------
                               Half-length: 0.61 cm
                          100    200    300   400
                             Time (sec)
Figure 2.14. Records of injection pressure as a function of time using various lengths
            of starter slots.
                               62

-------
 differing only by the lengths of their slots. Other experiments conducted on pairs of
 similar samples (one large sample split into two) confirm the features shown in
 Figure 2.14.

       A decrease in the pressure at which fracturing occurs Pat is the most striking
 effect of increasing slot length. The relationship between Pa and half-length of the
 slot a is characterized by the empirical function


                                  Pdr=Cfl»                            (2.5)

 in which C - 39.3 and n = -0.61. Many other materials, including rock, steel,
 ceramics and glass, show a relationship similar to eq. 2.5, with values of C related to
 fracture toughness of the material, and values of n varying slightly from -0.5
 (Williams, 1984; Lawn and Wilshaw, 1975; Daneshy, 1976b; Clifton and others,
 1976).

       The pressure at which failure occurs is by no means the only effect of
 increasing the length of the starter slot; the shape of the record is also affected.
 Records from tests using shorter slots show a period of stable growth, during which
             period of stable growth becomes longer and driving pressure increases
by a smaller amount-several kPa-during stable growth. During unstable growth,
the amounts of decrease of driving pressure are less for fracture started from the
longer slots than from the shorter ones (Fig. 2.14).


SUMMARY AND DISCUSSION

       Observations in the preceding section suggest a scenario for the development
of a hydraulic fracture in soil that is straightforward, yet more detailed than previous
works (Mori and Tamura, 1987; Brunsing and Henderson, 1984; Jaworski and
others, 1981; Bjerrum and others, 1972).  The early stage of injection at a constant
rate was characterized by inflation of the starter slot, which resulted in a roughly
linear increase in driving pressure. A hydraulic fracture formed when the driving
pressure reached some critical value, the magnitude of which diminished with either
an increase in the length of the slot, or a decrease in the moisture content of the
soil. Driving pressure continued to increase during the early stages of propagation.
At some point, either at the onset of propagation or after the fracture had reached
some critical length, the injection fluid lagged behind the leading edge of the
fracture leaving an unwetted zone at the tip. Once it began to form, the length of
the unwetted tip increased roughly in proportion to the length of the wetted zone of
the fracture.

      The slope of the pressure record decreases from roughly the onset of
propagation, resulting in a concave downward form. The forms of the records
differed in detail, however, depending on the length of the starter slot and the
moisture content. A short slot resulted in a record that was tightly concave, with the
pressure diminishing sharply following the onset of propagation. A longer slot
caused the concavity of the record to broaden, so that the pressure increased slowly
                                   63

-------
and then decreased slowly during propagation. Keeping the slot length constant, but
decreasing the moisture content resulted in a similar change in the shape of a
record: it was sharply concave when the samples were relatively dry, but it became
broader and flatter as the moisture content increased.

      Moisture content affected the development of the unwetted tip as well. An
unwetted tip was absent, that is stain covered the entire fracture surface in samples
less than 21% in moisture (roughly the plastic limit). As moisture content increased,
the length of the unwetted tip relative to the length of the dyed fracture increased,
until it reached a maximum length of 0.25cm in samples of 28% moisture.
Preliminary evidence indicates that the unwetted tip decreases in relative length as
moisture content increases to values greater than 28%.

      The growth of the unwetted tip appears to be a fundamental feature of
hydraulic fractures in Center Hill clay at moisture greater than the plastic limit
That feature is by no means unique to day; unwetted zones at the leading edge of
hydraulic fractures in rock have been observed in experiments by Medlin and Masse
(1984).  Their apparatus differed from the one used in this research in that theirs
was fitted with ultrasonic transducers which yielded data on aperture and fluid
content as a fracture was growing. They observed marked attenuation of the
ultrasonic signals near the tip of a growing fracture and inferred that the tip of the
fracture was filled with air. Results from one of their experiments (Medlin and
Masse, 1984; fig. 6) indicates that the unwetted zone first forms at the starter slot, it
grows to length Lw and then is roughly constant as the length of the wetted zone Lu
increases during propagation. This differs slightly from our results, which indicate
that Lw increases with Lw.

      Similarity between surface textures of the unwetted tip and the wetted
fracture, as well as the continuity of linear features from the wetted fracture to the
unwetted tip indicate that the unwetted tip is open and a pan of the fracture itself.
This is in contrast to so-called decohesion zones, or process zones, identified or
postulated from studies of fracturing of metal, polymer or rock. The process zone is
an interval over which the material becomes separated. In tensile fractures, of
which hvdraulic fractures are a special case, processes of decohesion involve the
nucleation and growth of voids or microfractures (Williams, 1984; Tetejman and
McEvily, 1967). The process zone is characterized by vestiges of material cohesion
resulting from the strength of intact material remaining between voids, and as such
it differs from the fracture itself where the material is completely separated and
cohesion is absent. Process zones composed of clusters of voids have been
described at the tips of fractures in metals (Shockey and others, 1979; Tetelman and
McEvily, 1967; Irwin, 1957) and polymers (Williams, 1984; Weidmann and Doell,
1979).  Ouchterlony (1982) postulates a process zone in rock that consists of dilating
microfractures, which produce acoustic signals that are commonly used to monitor
the fracturing process in rock (Ouchterlony, 1982;  Zoback and others, 1977). Time-
lapsed sequences of photographs by Knauss (1976) capture the coalescence of voids
resulting in the formation of fracture surfaces in polymers.

      Phenomena occurring in the process zone of a fracture in clay have received
little attention. In normally-consolidated kaolin, micron-sized voids occurring
between aggregates of grains were observed to grow by coalescing during the early
stages of shear loading (Smart and Dickson, 1979). To our knowledge, processes at
                                     64

-------
the tip of a fracture in clay under tensile loading have yet to be studied, but it seems
reasonable to assume that those processes involve microfracturing or void growth
ahead of the fracture itself. Whatever processes are involved, the zone in which
they occur appears to be ahead of, and distinct from, the unwetted tip.

       In cross-section, an idealized hydraulic fracture in clay is inferred to be
composed of a zone wetted by pressurized injection fluid, an unwetted zone near the
tip where fluid pressure is zero, and a process zone in front of the tip where
decohesion of the clay takes place. The idealized cross-section is shown in Figure
2.15, where lobes are assumed to be absent for simplicity.

       Hydraulic fracturing of clay was inevitable in the laboratory experiments; it
occurred in all samples that could be properly prepared. We tried to inhibit
fracturing by increasing the water content and softening the clay. The wettest
sample that could be prepared as described above, however, could be hydraulically
fractured and the appearance of the fracture in that sample resembled that of drier
samples. Samples or extremely high water contents were too soft to be handled and
we were unable to prepare them using the procedures described.
                                    65

-------
ID
C
O

m
(Q
O
O

i£i
<***">•

cu
•»*
E-
•o

**
^
o
^c
c
^

G

0
r*
^^
o
0.
H

1 Wetted Fracture 1 "°
. .,,,,,»,,,,f,,m,,/M/M/SS///ff///////M/^^^ •

TJ

O
O
PA
CO
CO

N
O
^^
l»l
•<".'.".»

Figure 2.15.> Wealized cross-section of a hydraulic fracture in clay.
                                      66

-------
                          SECTION THREE

        ANALYSIS OF HYDRAULIC FRACTURING OF SOIL


      Since the work of Kristianovich and Zheltov in the mid-1950s, the
 principles of elasticity theory have been used to analyze the onset and
 propagation of hydraulic fractures in rock.  Those principals have resulted
 in  a significant body of published papers describing  both methods  of
 formulating and solving for various fracture geometries and boundary
 conditions, and techniques of applying  those methods to design field
 applications.   Details of the methods vary considerably, but  the use of
 elasticity theory to determine deformation of material enveloping the
 fracture is virtually universal.

      The broad knowledge-base of published  analyses based on elasticity
 theory  would be an extremely valuable tool in designing hydraulic fractures
 under near-surface  conditions.  The application of such methods to analyze
 the hydraulic fracturing of soil, however,  is currently untested.  The
 appearance of fractures created in clay resembles the appearance of hydraulic
 fractures created in rock, as well as.that  of tensile  fractures in a range of
 other materials. This semblance is encouraging, but appearance alone  is
 insufficient to provide grounds for adopting a mechanical theory.  Previous
 investigators of hydraulic fracturing of soil (Mori and Tamura, 1987; Horsrud
 and others,  1982;  Jaworski  and others, 1981  and 1979; Massarsch, 1978; Leach,
 1977;  Wong  and  Farmer, 1973; Bjerrum  and others, 1972; Morgenstern and
 Vaughan,  1963) have developed analyses based  on theories of elasticity  or
 plasticity to provide predictions of pressures required to induce hydraulic
 fracturing from a cylindrical hole in soil.  Those analyses are  also
 encouraging, but they are limited to predicting  initiation from  a borehole
 and reveal  little about the behaviour of the fracture  itself.

      Is elasticity theory applicable to hydraulic fracturing of clay?  The
 purpose of the  following chapter is  to answer that question by  analyzing
 critical aspects  of  the lab experiments.   The method of analysis will be
 acceptable if it can predict  or explain:

      1. The onset  of fracturing
     2. Essential features of a hydraulic  fracture
     3. Forms  of pressure records  observed during propagation

     Those criteria will be  addressed using analyses  based on linear elastic
 fracture mechanics, a branch of elasticity theory.  Results of the analyses
 will be compared with the results of lab experiments, and the  degree to which
 they are similar will determine applicability of the approach.

     The analyses  that  will be developed are intended to predict effects,
principally the growth of an unwetted zone at the fracture tip,  that are
commonly ignored  by investigators who study hydraulic  fractures in rock.
Unwetted zones do  develop in hydraulic fractures cutting rock, however, so
the results of the analyses may be  relevant to hydraulic  fractures in both
                                  67

-------
     The growth  of unwetted zones at the tips of hydraulic fractures  in
rock is particularly well documented by Medlin and Masse (1984).  Those
investigators used an apparatus that was similar to the one used in this
work.   One  important  difference, however,  is  that their apparatus  was
fitted with ultrasonic transducers yielding information on fracture
length,  aperture, and fluid content during propagation.  The experimental
method used in this work was only  able to yield pressures during
propagation.

     The primary goal of the following chapter  is to explain results of
lab experiments of hydraulic fracturing of soil.  Certain experimental
results of Medlin  and Masse will also be used in the following  chapter
because they offer data describing propagation.
PREDICTING THE ONSET  OF HYDRAULIC FRACTURING
     The stresses in the vicinity of a straight fracture in an infinite
elastic medium  are concentrated at the fracture tip.  In the vicinity of
the tip, stresses are proportional to  r    where r is the radial distance
from the  tip.  That relation results  in a singularity in the stresses at
the fracture  tip itself, at r = 0.  The strength of the singularity is
characterized by the stress intensity factor K which is given by (Inglis,
1957; Tada and others,  1985)

                          K = <7/7uTflb)                           (3.1)

where  a is the  applied stress, a the half-length of the fracture, and f(b)
is a function of the geometry of the fracture and the enveloping material.
     The driving  pressure P   (Secor and Pollard, 1975) of a hydraulic

fracture is the difference between the internal fluid pressure p,  which
tends to open the fracture, and  the  confining stress 
-------
provide a means of predicting  the driving pressure required to  initiate
fracturing.

      Critical stress intensities  of soil were determined  through eq. (3.1)
by  obtaining a critical driving  pressure P  at the first change  in  slope
of the pressure record from the  lab experiments  described in Chapter Two.
Other accepted  methods, such  as those decribed by Ouchterlony (1982) for
fracture  toughness testing of rock,  pick the critical load at fracturing
based on a certain (5 percent)  deviation  from linearity  of the record of
load with respect to crack opening displacement (COD).  Measurements of
COD were  impossible to make using our apparatus, however, and our
observations suggested that fracturing occurs at, or perhaps  even slightly
before, the first change in slope of the pressure  record.

      In  general, the function f(b) in eq. (3.1) is equal to unity under
the ideal conditions of a straight slot of zero initial aperture in an
elastic medium  of infinite extent.  It  is less than unity and can be taken
as a  correction  factor for practical constraints such as  finite sample
size.   Tada and  others (1985)  give expressions for f(b) for a straight
slot containing an axial hole, and for a straight slot embedded in a
rectangular sample of finite size. Those conditions represent relevant
corrections  for the experimental  apparatus used in this  work.   However,
correction factors determined for the geometries of the lab apparatus are
minor (within five percent of 1.0),  so f(b) was set equal to 1.0 in the
determinations of K .  The critical  stress intensity was calculated using
                                                                       (3.3)

     Tests were conducted to determine the relation between AT,  and the
                                                               1C
length of the starter slot  (Octerlony, 1982).  These tests consisted  of
preparing sets of similar samples  and cutting slots of  a different length
in each sample.   Sixteen  sets of samples of Center Hill clay were
prepared, where  one set consisted of four full-length  samples, two  sets
consisted of three full-length samples, and  the  other sets consisted  of
one full-length sample cut in half to give two  half-length  samples.   One
set of three  full-length samples composed of River Downs silt  was also
tested.  The samples  were fractured  using  procedures  described in Chapter
Two and  K   was determined using eq. (3.3).


     According to the results of those tests (Fig. 3.1), doubling the
length of the starter slot  resulted  in  an average variation of 0.104
(variation was determined as the absolute value of the difference between
two measurements divided by their average).  That average value includes
the results of several sets of unusually large variation (one value of
1.0, and two others greater than 0.4).   Nevertheless,  the average
variation caused  by doubling the length  of  the  initial fracture is similar
to the variation of 0.081  observed among results from four identical
samples described in the previous chapter.
                                   69

-------
         200
         180
      ^120
      PH.100
      ^08°
      w~eo
          40
          20
           0
    O—O Center Hill clay
                 ,o	
                                             D—Q  River Downs silt
0.0
                          1.0
2.0           3.0
 &i   (cm)
4.0
Figure 3.1.  Critical stress intensity as a function of half-length of
           starter slot for similar samples of Center Hill clay and River
           Downs silt.
                              70

-------
      The critical stress intensity of rock and metal is observed  to
increase slightly as a function  of slot length, if the slot length is less
than  a  critical value that depends on £fc and a  tensile yield strength

(e.g.  Ochterlony (1982); Schmidt (1977); Kaufmann and Nelson  (1974), and
references therein).  As the length of the starter slot  increased  in the
experiments using  soil,  the critical stress intensity  decreased for ten of
the sample sets, it  increased for four of the sets, and it both increased
and decreased  for two sets of  samples (Fig. 3.1).  Although K^ decreased

for more sample sets than it increased, there appears  to be no systematic
relation  between K  and a.  It is unclear whether  K  is independent of a
                   1C                                K
for the Center Hill clay, or whether the critical length for that
dependence is  less than the slot lengths used in this study.

      The results of the previous two paragraphs indicate that  K^ can be
used to predict (within roughly  10 percent) the driving pressure required
to propagate a hydraulic fracture in Center Hill clay.   Likewise, the few
tests conducted using River Downs silt suggest that ATfc can be used to

predict hydraulic fracturing in that material as well.   It is apparent
from  Figure (3.1), however, that the value of K^ varies over more  than  an

order of magnitude for  samples of the Center Hill Clay.  Thus, some method
of anticipating K  will be required to make a useful prediction  of the
                 1C
onset of propagation.

     Moisture content and duration of consolidation varied from one sample set
to another, and they both  have  a marked  affect on the  critical stress
intensity.  Among samples that were prepared by compaction alone, K^ is
greatest, roughly 200 kPa  cm172, at moisture  contents  between 0.17 and 0.21.
The critical stress  intensity decreases  abruptly to roughly 35  kPa cm  ,
however, as moisture content increases from  0.21 to 0.22.  The moisture
content of this sharp change in K  is slightly wetter than the plastic limit

of the soil.  Further increases  in moisture slightly decrease  the  average K^,

although the decrease is  small  and K^ is  practically independent of moisture

content over the range 0.22 to 0.30.

     Critical  stress intensity is  neglible for samples of moisture content
greater than 0.32.   This means that the break in slope  of the pressure record,
from which the onset of fracturing was inferred, occurs when the pressure of
the injection fluid  is equal to the confining stress.  Hydraulic fractures
produced during  these tests are  physically indistinguishable from other
fractures produced when ATfc is greater than zero.


     Consolidation appears to toughen the Center Hill clay (Fig 3.2b).
Critical stress intensity of samples that were consolidated  for four days
using procedures described in the previous chapter are greater than they  are
for the compacted  samples of similar  moisture.   Moreover, increasing the
                                   71

-------
duration of consolidation to 14 days further increases K,  according to

Figure 3.2b.  The forms of the three  curves in Figure 3.2b are similar,
however; they all show  K^  decreasing markedly (from roughly 200 kPa cm   to

between 30 to 50 kPa cm  ) over a few  percent of moisture.


      Limited tests conducted on the River  Downs  silt show that K is four or
                                                                  Ie
five times less than K  of the Center Hill  clay at similar  moisture content

(Fig. 3.2).  This  result  suggests that fractures will develop in the  silt at
lower driving pressures  than in the clay,  if the two materials contain  the
same weight percent water.  Lab experiments to confirm this suggestion  were
not conducted, but it  is consistent with the common behavior of the two
materials;  the silt readily fractures and crumbles, whereas  the clay tends to
resist crumbling  when pieces of the material are worked manually.


Discussion

The critical stress intensity is a material  property  of  clay that depends
on  moisture content and duration of consolidation,  according  to the results of
this investigation.  Driving pressure marking the onset of propagation of a
hydraulic fracture in the Center Hill clay could be roughly predicted by
estimating K^ from water  content and consolidation history and measuring the

length of a pre-existing  fracture.   The effects of other parameters, such  as
compaction effort, will also affect K  (e.g. Jaworski  and others, 1979)

although the present study  was limited to effects of moisture  content and
duration of consolidation,   the requirements of an  acceptable  theory of
hydraulic fracturing of soil, the ability to predict propagation  of a
pre-existing fracture, appears to be satisfied by the critical stress
intensity factor.

     The use of K   as  a fracture criterion generally  requires that  the

fracture process zone, where inelastic behavior such as microcracking  or  void
coalescence occurs,  is small compared with the  dimensions of the fracture and
the specimen (Tada and others, 1985;  Williams,  1984; Ouchterlony, 1982; Rice,
1968). Schmidt  (cited in Ouchterlony,  1982) describes a method of estimating
the size of the process zone based  on a maximum  normal  stress criterion.  That
method tacitly  assumes  that the dilation of microcracks is  the dominant
phenomena occurring in the process zone, an assumption that is accepted for
rocks (Ouchterlony, 1982)  and seems reasonable for clayey soil as  well.  The
maximum  length of the  process zone L is  given by


                     L  = 0.269 [K l(a   + a )]2                      (3.3a)
                      P             Ie  y     c

where a  is the mean confining stress, and a  is the  tensile  yield  stress

required to open microcracks in the process zone.  Several investigators  have
                                  72

-------
a
o
cfl
0.15
                     0.20          0.25          0.30
                     Moisture (wt. water/wt.solid)
0.35
 a
 o
 (d
CU
                                           — Compacted
                                            A Consolidated 4 days
                                            O Consolidated 14 days
       0.15
               0.20          0.25          0.30
               Moisture (vi. -water/Trt.solid)
 0.35
 Figure 3.2  Critical stress intensity of Center Hill clay as functions  of
             moisture content  and duration  of consolidation.   Filled circles
             on lower plot are for River  Downs silt.
                                   73

-------
described methods  of measuring tensile yield strength of soil (Snyder and
Miller, 1985; Snyder, 1980; Vomocil and Chancellor, 1967; Farrell and others,
1967), but all of those methods make use of special equipment that was
unavailable during  this study.   The tension of pore-water, however,  could  be
taken as a tensile  yield strength because dilation of microcracks in the
process presumably must overcome the pore-water tension.   In a  sample of 0.20
moisture under load  in the fracturing cell, the pore-water tension  was 60  kPa,
according to measurements made using a tensiometer  (SoilMoisture Model  2100).
Measurement of  a  sample containing 0.27  moisture yielded inconclusive results,
perhaps because  the  sample was saturated  and the gauge attached  to the
tensiometer was  only able to measure suction; the tensile strength of the 0.27
material will be  assumed to be zero.  The critical stress intensity for  0.20
moisture is roughly 180 kPa cm   , whereas for 0.27 moisture it is 50 kPa
cmre (Fig. 3.2).

     The maximum length of  the  process zone ranges from 0.47 cm for 0.20
moisture down to 0.13 cm for 0.27 moisture,  according to eq. (3.3a) and  data
in the proceeding paragraph.  The length of the  initial slots ranged  from  1.22
to 5.05 cm,  and  the length  from the end of the  slot to the end of the sample
was greater  than 5 cm.  The  process zone in  the Center Hill  clay was always at
least several times smaller than both the initial starter slot and the  sample.
This suggests that  K^ should be a reasonable predictor of fracturing,

according to Tada  and  others  (1985)  and Williams  (1984),  who state that  the
accuracy of  K improves as L diminishes to lengths  significantly  less than

the length of the starter slot or the sample.  It  is apparent from  eq. 3.3a
that L  decreases as either tensile strength or  confining stress increase, so
      p
that the accuracy of K  should improve as confining  pressures increase to

values greater than the 69 kPa used  in the lab.  Confining pressures of most
anticipated field  applications, for example, would exceed 69 kPa.  The
relation between  the size of the process zone and the accuracy of JTfc  also may
explain why  the  drier samples, which have the larger process  zones, exhibit
the greatest  variability  of K   (e.g. Fig.  3.1).
                           Ie
     We  recognize that in some cases incipient fracturing may occur at driving
pressures  less than those marked by a break in slope in the record.  A change
in slope of the  pressure record was  the only method available to detect the
onset of fracturing during the present  work.  Values of critical  stress
intensity presented here should be taken as apparent intensities for the onset
of fully-developed hydraulic fracturing and they may overestimate the critical
stress intensity  obtained from other  types of tests, such as a compact tension
test or  three-point bend test  (Tada and others,  1985).

     The method described above  is by no means the only one  that uses
principals of linear elastic fracture mechanics to predict the onset of
tensile fracturing  of soil.  Briones and Uehara  (1977b) cracked rectangular
beam-shaped samples of soil  in a  bending  apparatus and proposed a material
characteristic  (their eq. 9)  that is  equivalent to
                                     74

-------
                                                                      (3.4)


where a  is assumed  to be  the tension of pore fluid, and a  is an effective
        t                                 f
length of pores, or microcracks in the soil.  Their expression is  based on the
classic works of Griffith (1920) and Irwin (1957), and it is  identical in form
to eq. (3.1).  The tension of pore fluid  can readily be determined using a
tensiometer, but Briones and Uehara (1977b) were unable to determine
appropriate  effective  sizes of pores in  their soil samples, and thus were
unable to determine K  through eq. (3.4).
                      Ie
     Snyder (1980) and Snyder and Miller (1985) use methods of linear elastic
fracture  mechanics to develop another analysis predicting the tensile fracture
of soil.  They recognized  the importance of pore size and shape in nucleating
tensile fractures, and  they were fully aware of the practical difficulty of
measuring critical shapes  and sizes of pores.   To address this problem,  Snyder
(1980) introduces a parameter based on  the degree of saturation  that is
intended to characterize the role  of pore size and shape in the fracturing
process.   According to Snyder (1980), this approach  to characterizing  the  role
of existing  flaws yields results  that can  explain both his own data and
published data on tensile  cracking of soil.  Various other investigators
(Rogowski and others, 1968; Farrell and others,  1967; Kirkham  and others,
1959) have used methods  based on elasticity theory to estimate  tensile
strength  of soil without explicitly considering the effects of pores or
microcracks.

     The approach used here differs  only slightly from  that of previous
workers.  The principal difference—we cut a large slot to nucleate a  fracture
whereas  they omitted  the  slot-is significant because  our  method  explicitly
addresses the role of  existing cracks  in  a sample.

PROPAGATION OF A HYDRAULIC  FRACTURE IN SOIL


     Predicting or explaining essential features of both the hydraulic
fractures  themselves and the pressure records during propagation are two  more
requirements of the analysis.  Meeting those requirements will be accomplished
by including effects of propagation in the analysis.   The propagation of a
hydraulic fracture involves two basic processes, the flow of liquid within  the
fracture  and the dilation of the fracture  due to deformation of the enveloping
medium.  The processes are coupled in  that  the  pressure distribution resulting
from viscous losses during fluid flow is  strongly dependent on fracture
aperture,  and the aperture depends on the pressure distribution.   Flow within
the fracture will be treated  using methods of fluid mechanics, and dilation of
the fracture will be treated  using methods of elasticity theory.

     Another condition, one that ensures equilibrium propagation, must  also be
considered.   In keeping with the  approach of linear elastic fracture mechanics
we will require that the criteria for onset of fracturing  is maintained
                                   75

-------
 throughout propagation.  Accordingly, during equilibrium propagation the
 pressure  distribution within the fracture must  always result in the critical
 stress intensity.  The fracture is considered to be out of equilibrium if
 the  stress intensity exceeds the critical value,  and propagation will cease
 if the stress intensity is less  than the critical  value.   Other methods of
 maintaining equilibrium propagation are described  by Geertsma and Haafkens
 (1979).


 Conceptual Model


      A conceptual model of fracture growth is the sequence of events
 describing the progression  from an inflated slot to a typical fracture  in
 the  lab samples.  This model of growth will be inferred from lab
 observations and  from simple reasoning, and it will  used as a framework for
 formulating theoretical analyses of propagation.

      The model begins with a narrow slot cut in  the material (Fig. 3.3).
 The slot  is inflated at a constant volumetric rate,  resulting in an increase
 in fluid  pressure  and an increase in the intensity of stresses at the edge
 of the slot.  All  the inflow during inflation results in dilation of the
 slot, so  the flow rate diminishes along the fracture  and  must be zero at the
 tip.   The fracture begins to propagate when the stress intensity equals the
 critical stress intensity for the material, according to the principles
 described earlier.  Separation  of material at the fracture tip during the
 onset of  propagation occurs in response  to  the  intensity  of  stress  within
 the  enveloping  material.  The stress intensity depends only on the magnitude
 and  distribution of pressure,  and it is independent of properties governing
 the  flow  of the fluid in the fracture.  Indeed,  separation of material at
 the  fracture tip will occur  when the critical stress intensity is achieved
 even if the fluid  in the fracture is totally immobile  and unable to fill the
 newly formed region (e.g.  Clifton and others,  1976).

      This concept is important  because it suggests that  the newly-formed
 region at the tip  can be unfilled by injected fluid  immediately following
 the  onset of propagation.   In  other words,  the  tip of the fracture could be
 unwetted  from  the start of fracture growth. The  lab experiments  conducted
 for this work were unable  to  yield conclusive information about the fracture
 tip  at the onset of propagation; however, Medlin and Masse (1984, fig. 6)
 clearly show that the unwetted region forms at the onset of hydraulic
 fracturing of rock.

      The lengths of unwetted zones appeared to be  proportional to the
 lengths of the  wetted parts of fractures  in  the  lab experiments conducted on
 clay samples of similar moisture content.  In relatively moist samples, the
length of the unwetted zone appeared to be linearly  related to the length of
 the wetted zone from the start of the fracturing, whereas in drier samples
we inferred that a critical  fracture length was  required before the  linear
relation was observed.   The latter observation  is consistent with our  model
if we postulate  that a short unwetted zone  forms at  the onset of  fracturing
in the drier samples and grows slowly until the critical  fracture length is
                                     76

-------
                                                     1.
                                                      2.
                                                        3.
Figure 3.3.  Conceptual model of the growth of an idealized hydraulic
           fracture in lab samples.
                              77

-------
reached.

     The conceptual model  is summarized in four steps shown in Figure 3.3.

     1.  Pre-existing starter  slot cut in a sample

    2.  Inflation of the slot leading to an increase in stress intensity,

          depicted by stipling at the tips of fractures in Figure 3.3.

    3.  The inflated slot  begins to grow when the £, = £fc.   Part of the

          newly-formed fracture is unfilled  by injected fluid.

    4.  The fracture increases in  length  and partly fills with  fluid.   A

          region at  the tip of the  fracture remains unfilled, and the

          length of  the unfilled region increases during propagtion.  Lab

          measurements suggest that the  length of the unfilled region is

          linearly  related  to the length of the wetted fracture, although

          that relation is  only approximate.


Analysis of Propagation

     Two analyses  will be used to examine the conceptual model  and to predict
the records of driving  pressure as a function of time.  Both analyses are
based on  linear elastic fracture mechanics,  although they differ markedly in
level of complexity.  The first analysis that is presented solves the  problem
of coupled fluid flow and elastic dilation and  requires numerical procedures
implemented in a lengthy computer program, whereas the other one neglects the
effects  of fluid  flow and  results in a simple analytical  expression.

     Both analyses  treat  the fractures in two dimensions normal  to the axis of
the starter slot.  The lengths of fractures produced in the lab were roughly
uniform with respect to their width,  so that the appearance of the fracture in
a cross-section normal to the axis of the slot is nearly independent  of
location along the width.  The fractures  gapped open at the sample edges, but
the aperture at  the  edges was probably less than within the sample  due to
friction between the sample and the apparatus.  That edge effect, as well as
effects  related to  the development of lobes at the leading edge of the
fracture are  assumed to be  negligible in  this formulation.   The hole along the
axis of the slot will be ignored because including it will markedly increase
the complexity of the  analyses while adding little insight.

     The experimental apparatus was  designed to restrict  deformation parallel
to the axis of the starter slot, placing  the entire sample in conditions of
plane strain  deformation.  It will  be assumed that plane strain conditions
exist at the  leading edge  of the fracture, and those  conditions  will be
adopted for  the analyses.  Shear stresses induced  at the edges of  the sample
by  the  dilation  of the  fracture will be  ignored, and  consideration of plane
stress conditions at  the edge of a notched sheet as described by Williams
                                      78

-------
 (1984; p.  99) seem unwarranted in light of the edge constraint applied by the
 experimental apparatus.

      The  numerical analysis is designed to obtain a distribution of pressure
 within the fracture that  simultaneously  satisfies equations describing the
 elastic dilation,  viscous fluid flow,  and  stress intensity. An  iterative
 procedure, which converges in less  than ten iterations  for most cases, was
 developed  for this purpose.  The procedure makes use  of the Schwarz-Neumann
 alternating technique  introduced  by  Pollard  and Holzhausen  (1979) to analyze
 static hydraulic  fractures,  and  similar iterative methods are  used  by Nilson
 and Griffiths (1986), Cleary and Wong  (1985), and Narandren  and Cleary (1983)
 to  analyze propagation.

      A variety of other  methods of analyzing  a propagating hydraulic fracture
 whose  geometry resembles the one described here are available (e.g. Nilson,
 1986; Spence and Turcotte, 1985; Biot  and others, 1982;  Geerstma and de Klerk,
 1969).   The method of solution  used in this work was adopted because it is
 straightforward,  yet versatile enough to  accomodate a  wide range of
 modifications that may be required  in the future.


 Overview  of the Approach of the  Numerical  Analysis


      The  analysis begins by assuming fluid is  injected  at  a  constant rate at
 the midpoint of  a pre-existing  fracture (Fig. 3.4).  Symmetry about the
 midpoint is assumed and  allows  us  to only  treat one  half of the fracture.   The
 fracture is embedded  in  an elastic medium of infinite  extent along cartesian
 axes  x and v, and it  is of width w along the z axis.   The fracture, which  is
 of  half-length a{, lies on  the x axis  and the v  axis is  normal to the plane of
 the slot.   Displacements  along the z axis are assumed  to be zero, reducing the
 problem to one of plane  strain in the x-y plane.

      An initial period of injection inflates'the  pre-existing fracture but
 does  not result in propagation.   Pressure within the fracture increases during
 inflation and propagation  begins  when K  =  K .  Propagation is  assumed to

 continue such that the equality K  = K  is always maintained.


      The  history of growth is  analyzed  by treating a series  of fractures  each
 of  which contains a slightly larger volume of fluid than the last.   The time
 since injection began, which yields a growth history,  is obtained by dividing
 the volume of a  fracture by the  volumetric  rate of injection. The volume can
 only be determined after selecting a  fracture length and solving for a
 pressure distribution,  so different methods of changing  fracture  volume are
 needed  for stationary  or  moving  fractures.   During inflation of a  pre-existing
 fracture, volume is increased by  incrementing the driving pressure at  the
 fracture tip, whereas  during propagation volume is increased by incrementing
 fracture length.

      This  method of  solution is  termed  quasi-steady (Nilson, 1986) in that  it
ignores dynamic  effects in the  elastic solutions; transient terms only enter
                                      79

-------
                                                                 X
Figure 3.4  Geometry used  in analyses.
                                 80

-------
 in the  mass balance used to  solve the flow equations.   Dynamic effects  are
 insignificant, however, when  the velocity of a fracture is  much slower than
 the  speed of elastic waves in the solid enveloping the fracture.  As Nilson
 (1986)  points out,  viscous effects in  the  fluid filling a hydraulic fracture
 will limit propagation velocity to much less than the wave speeds of elastic
 solids,  so the quasi-steady approach seems  to be justified  for  hydraulic
 fractures.


 Inflation of a Pre-existing Fracture


      The steps used  to analyze  inflation  of a pre-existing hydraulic fracture
 are  outlined below.

 1. Assume an initial  length for  the pre-existing  fracture and assign the
      locations of nodal points starting at the midpoint and ending  at the tip.

 2. Assume a uniform  pressure over the fracture that is  less than the critical
      pressure for propagation.

 3. Determine apertures at nodal points from elastic  solution.

 4. Use  an expression for flow in the fracture to determine a  new pressure
      distribution  at the nodal points.  Apertures from step 3  yield effective
      hydraulic conductivities  along  the fracture.   Mass balance requires that
      all the flow into  the fracture  is accomodated by dilation, so that the
      flow rate decreases along the  fracture length and is  zero at the tip.  At
      the midpoint,  second-type boundary  conditions  are  used to satisfy the
      constant inflow  rate.  At the  tip, a first-type condition satisfies the
      stress intensity requirement, that K  < K  .

 5. Use  the new pressure distribution  to determine new apertures and iterate
      from step 2 through 4 until convergence criteria are satisfied.  The
      criteria  for convergence used in this work  include: a.) the maximum
      difference between pressures on two successive iterations is less than a
      certain value,  typically two percent  of the  maximum pressure, b.) the
      inflow  rate determined by  finite differences is  within a certain value,
      typically two  percent, of the the specified  inflow rate, c.) flow rate at
      die tip is essentially zero,  d.) mass balance at an  arbitrary point along
      the fracture is satisfied  withiil a certain value, typically  five percent.

 6. Determine fracture volume by integrating aperture over the length of  the
      fracture, and  determine  the time since injection began  by dividing
      fracture volume by  the  volumetric rate of injection.  Save the variables
      of interest.

7. Increase the pressure at the tip and repeat steps  2 through  6.

 8. Repeat step 7  until the stress intensity exceeds the critical stress
      intensity.  Estimate the  pressure at  the tip  so  that K = K .   This

      estimate is achieved by  interpolation between the previous two solutions
      and the observation that stress intensity varies nearly linearly with
                                      81

-------
      small  changes in pressure  at the tip.

9. Repeat steps 2 through  6 for the  estimated critical pressure at the tip and
      check  whether K = K^.   If this  check is  satisfied (within two percent)
      then the  current time and  pressure  distribution mark the onset of
      propagation of the pre-existing  fracture.  Additional iteration is
      required if the check is not satsified, but this was rarely needed.

Propagation of an Inflated  Fracture


      Two schemes are used  to  analyze propagation, one that assumes the fluid
      reacheshe crack  tip and another that tracks the leading edge of the
      fluid and results in an unwetted tip.  The former is the simpler of the
      two and will be described  first.
1. Increase  the length of the fracture by a small increment  and reassign
      locations  of nodal points so that the new fracture is covered by the same
      number of points as  the old one.   Assign a uniform pressure distribution.
      Convergence is improved by assuming  the  initial pressure satisfies the
      stress  intensity requirement for a uniformly distributed pressure in a
      fracture of the new half-length.

2. Determine apertures  from an elastic solution.

3. Use  an expression  for flow in the fracture to determine a new pressure
      distribution.  Apertures  from step 3 yield effective hydraulic
      conductivity along the fracture.  The difference between the current
      apertures and those at  the previous time step yields the mass balance in
      the flow  equation.  At the midpoint, second-type boundary conditions  are
      used to satisfy the constant inflow  rate.  At the tip, a first-type
      condition is  used to obtain the  stress intensity requirement, that KI -
5. Use the  new pressure distribution to determine new apertures and iterate
     through step 4 until the convergence criteria are satisfied.
     Oscillations between successive iterations are common in this iterative
     sheme, as illustrated by Narandren and Cleary (1983).  The  rate of
     convergence is improved by using a weighted average of the previous two
     iterations until the solution is  close to convergence  (Nilson and
     Griffiths,  1986).  The final two iterations lacked  a weighting factor to
     ensure that the maximum difference criteria (criteria a. above) is
     satisfied.

6. Determine stress intensity and compare with  critical stress intensity.  If
     the stress intensity is within two percent of the  critical value, then
     the solution for the current fracture length is complete.  If the
     critical stress  intensity is not satisfied, another solution is  obtained
     using  a slightly  different pressure at the tip.   A third solution is then
     obtained using a  new pressure at the tip.   Selection of that  new pressure
                                      82

-------
      is based on the observation that stress intensity is nearly  linearly
      related  to the specified pressure at the tip.   Thus, the previous two
      solutions are used  to estimate a new tip pressure  that  results  in the
      critical  stress intensity.  In many cases,  the  critical stress intensity
      was satisfied after convergence  of the first solution, and it was always
      satisfied by the linear  scheme described above  when the increment ,pf
      crack growth was  less  than five percent  of the crack  length and fi  <  1.0,
      where n = qi*ElK*e with q the  volumetric inflow rate, // dynamic viscosity,

      and E elastic modulus.  Larger  values of n  resulted in numerical
      instabilities and an alternate method based on  adjusting  the input  *
      pressure was used  to obtain convergence.  Those larger  values  of n
      exceed the range of lab experiments, so  the alternate  method was
      unnecessary for the present work.


 7. Save variables of interest  (e.g. time, driving pressure, length, aperture)
      and increase fracture length by  a small  increment.  Reassign the
      locations of nodal  points and determine  an initial  pressure distribution.
      Repeat  steps 2 through 6.

 8. Repeat step 7 until the desired fracture length is  reached.
      Injected fluid completely filling  the fracture is  a tacit  assumption of
      the analysis described above; however, results of that  analysis show that
      the velocity of fluid near  the fracture tip is less than  the velocity of
      the fracture tip itself (Fig. 3.5). This suggests that the fluid is
      unable to keep pace with  the fracture and the  the assumption of a
      completely  filled fracture is incorrect (Abe and others, 1976).  The
      analysis was modified to track  the position of  the wetted  front by
      integrating the fluid velocity at  the tip over each  time step.  The
      integration  was  done using a finite difference scheme  that is  backward in
      time,  which amounts to using the velocity over the previous time step to
      predict the location of  the wetted front at a new time step.  As a
      result, during the first  increment of crack growth  the  fracture front
      moves forward a small increment but the fluid front remains stationary
      because its velocity at  the  previous time  step was  zero.
      The method used to assign locations of nodal points was modified during
      tracking of the wetted  front.  The modified method generated one set of
      points that were spaced evenly  between the midpoint of the fracture and
      the wetted  front, and another set of points that were  spaced  evenly
      between the wetted front and the fracture tip.  Driving  pressures were
      positive and resulted from  the pressure of the injection fluid  over the
      first set of points,  whereas they were negative  and assumed to be equal
      to the confining load over  the second set. The  first set of points was
      used to  solve the flow  equation  and  both sets were used to solve for
      apertures and stress intensity.

Outline of  the Analysis

      Implementing the method of solution requires the fracture  to  be
discretized into many small segments  defined by the  nodal points.  Apertures
and  stress intensities  are obtained by  integral  methods cast  as summations
                                   83

-------
    1.000
>
>
    0.100
    0.010
    0.001
                                                       fracture
         0.0      0.2       0.4      0.6       0.8      1.0   j   1.2

                          Fracture Length (a/a ^          ^P
   Figure 3.5. Velocities of fluid within a fracture and of the fracture  tip,
              scaled to the velocity of fluid entering the fracture, v .
                                     84

-------
 over values calculated at each segment, and pressures are obtained using a
 finite-difference method based on the mid-points of the segments.  Details of
 calculating apertures, stress  intensities and pressures are described below.

      Consider  a fracture loaded by a uniform  driving pressure P  over the
                                                                be
 segment b <,  x  <> c, where b and c are nodal points bounding the segment.  The
 aperture at any point x  on the fracture resulting from a load  on that  segment
 is  (Erdogan,  1962; Tada and others, 1985)

                               - sinlb/a)(a2- x2)in 1             (3.5)
 assuming plane strain deformation.

      A  driving pressure of arbitrary distribution  P(x) over the entire
                                                  d
 fracture can be approximated by a series of loads, each of which is uniform
 over  a segment and  equal  to P(x  ), where x    is the midpoint of a segment.
                               d nip          mp
 Summing the contributions of each segment yields the aperture at any point.
 Assuming  there are a total of n segments,  where  the ith segment is bounded  on
 the left  side by b and  on  the  right side by c  leads to
         4(1  -  y2)  r_    r             a2-ex               t
                          P.l(c.- x) cosh'1	L_ - (b- x) coslf1-
              E      *~   'L  '          a\x -
                   z =
                   + (suf'c./a - sin-1fci/fl)(a2- x2)1'2 1          (3.6)



An expression  for  the stress intensity is obtained using a similar  approach
                                    85

-------
              n
           i  = -n
                   - [1  -  (c/o)]-  [1 -
which  is based on  an equation for stress  intensity due to loading over a
single  segment given by Tada and others  (1985).

Pressures

     Pressures within the fracture are determined by satisfying  conditions
governing laminar  flow that is one-dimensional  along the fracture length.
Assuming the change in aperture is  small over  any  given segment and the  fluid
is linearly viscous, then the flow law within that segment is of the form
                        q = • (<53/C/0(aPd/ax)                        (3-8)
where  q is  the volumetric flow rate per unit width  of the fracture,  and n  is
the dynamic viscosity of the fluid within  the fracture.  In general, the
constant C  depends  on Reynolds number R  and wall roughness.  If the walls of
the fracture are perfectly smooth, GI decreases  from  1.5  for J?e of roughly one
or greater to  12 for  R,«  1-0 (Schlichting, 1960).  Wall roughness tends to
increase the value  of C  and the magnitude of  the  increase depends on Rf.
Detournay (1979) reviews experimental studies of the role of wall roughness  on
flow through  a  fracture.
     The Reynolds Number during flow through a  narrow slot is (Detournay,
1979)
                                                                      (3.9)
which yields R = 5 x 10"4 for conditions of the lab experiments.  The effects
of the roughness of the fracture walls will be ignored,  so a value of C^  12
will be used (Schlichting, 1960).
      Conservation of mass requires
                           dJ/ar = -Bqldx                           (3.10)
assuming that  the  fluid is incompressible, that  no fluid is stored within the
                                     86

-------
pumping system,  and that no fluid leaks off through the walls of the fracture.
All those assumptions can be relaxed by modifying the conservation equation.
Nilson  (1986)  shows  how to include  effects of compressibility of injected
fluid and storage within the pumping system.  A  simple  but  widely-used
analysis of leakoff was first described by Carter (in Howard  and Fast, 1957),
and more general analyses of the problem  are treated by Settari (1983), Dean
and Advam  (1984) and Pascal (1986), among others.

      Substituting eq.  (3.8)  into (3.10) yields


                 dS/dt  = BTIto dPIdx +  T tfPJdx2                  (3.11)
                                   o             d

where the transmissivity 7 is


                             T = S3/fiCl                             (3.12)


      Boundary conditions at the midpoint of the fracture require that
                                             x = 0                  (3.13)

where q is the total flow rate per unit width of the fracture  and 8  is  the
       ^o                     f                                   o
aperture at the midpoint.   The  pressure at the fracture tip must  also be
specified  to control stress  intensity and  ensure that K  = K  during

propagation, so

                         Pd = P..    * -  «                        (3.14)

      The flow equation was solved  for the  boundary conditions described above
using an  implicit finite difference scheme.   Nodal points in the finite
difference scheme were midpoints of segments  in the elastic solution.  Several
methods of spacing the points along the fracture  were evaluated, but
uniformly-spaced  points yielded the most satisfactory  results.   Other methods,
which were designed to decrease spacing in  the vicinity of the fracture tip,
proved  to be  less stable  than the uniform spacing even when the finite
difference scheme was written to account for non-uniform  spacing.  The
following description is valid only 'for uniformly-spaced nodal points,
although it is readily modified to accomodate non-uniform  spacing.

      A second-order finite difference approximation of eq. (3.11)  is obtained
using Taylor series expansion and the resulting expression  is written in
implicit form as
                                         ,- «                      <3-15a>
where
                                   87

-------
                                 - [(T.I-  T^MJ*2]                 (3.15b)


                                                                    (3.15c)


                                                                    (3.15d)



                                                                    <3'15e>
 in which the subscripts i are indices of the the nodal points and the
 superscripts j indicate an increment of time.   The total volume per unit width
 of the  fracture is  V,  and the nodal spacing is Ax.   The boundary condition at
 the midpoint of the fracture eq.  (3.13) requires

                               a= 0                               (3.16a)

                            ^=  -27/Jjc2                            (3.16b)


                            c=  2TJAX2                            (3.16c)


              m= (6\ - d\'l)/[(V j - V H)/?o] - qJAx               (3.16d)

 whereas the boundary condition at  the tip  is written

                               «n= 0                               (3.17a)


                              &n= 1.0                              (3.17b)


                               cn= 0                               (3.17c)
The resulting set of n equations was written in tridiagonal form and solved
using a commercially-available inversion  algorithm.

     Tracking the length of the wetted zone a  was done by integrating the

velocity at the tip over time.  The fluid velocity v  at   the tip is obtained

to first order by the backward difference
                                                                    (3.18)

-------
 where m is the number of segments wetted by fluid.  The change in length  of
 the wetted part of the  fracture a  during a time step is established by
 and GW is determined at any given time by summing the increments of growth

 during the previous time  steps.

      This approach assumes that pressure within the unwetted zone is constant,
 which is a reasonable simplification although we recognize that it tacitly
 ignores the matching  of pressures and velocities at the interface between the
 injected fluid and the fluid filling the tip.  It also ignores capillary
 effects at the boundary between  the fluid  front and the  unwetted tip, as well
 as contributions that  the rate of change of pressure distribution and crack
 length have on stress intensity (e.g. Cleary and Wong, 1985; eq. 16).   The
 analysis described above could be modified to consider all of those  effects,
 and such modifications could be  important in predicting  various details of
 hydraulic fractures.  However, we expect that the major conclusions drawn from
 the analyses will be unaffected, so those modifications will be deferred for
 future work.


 Results


      Numerical  analyses were  conducted using parameters  resembling those of
 the Center Hill clay at moisture  contents between 0.26 and 0.27 (Table 3.1).
 The driving pressure on the unwetted zone P    was assumed to be zero,
                                            daw
 implying  that the fluid  pressure equals  the confining stress along that zone.
 In  its present formulation, the analysis is unable to account  for effects such


 TABLE 3.1  PARAMETERS USED IN NUMERICAL  ANALYSIS
*«•'
I* :
C:
i
Klc:
ai:
0.00375 cm2/s
175 cp
12

50 kPa cm1/2
1.0 cm
u:
£:
P :
dnw
w:
m:
0.3
10 MPa
0

34
30
as capillary forces, and leakoff, both of which will affect the distribution
of pressure at  the tip.   Accordingly, P   should be taken as an effective
                                      daw
stress that crudely accounts for the factors  omitted in the analysis, and  its
value was  estimated based  on the  form of pressure records generated by  the
analysis.
                                   89

-------
     The results show that  an unwetted tip begins to develop at the onset of
fracturing (Fig. 3.6)  and then increases in length during propagation.  This
result essentially confirms the conceptual  model.  Moreover,  it shows that  the
difference between the velocity of the fracture and the velocity  of the fluid
near the tip could  be the process responsible for the development of a
fracture  with an unwetted tip.

     The numerical model predicts (Fig.  3.6) that the length of the unwetted
zone a  increases rapidly compared with that of the wetted fracture a^ during
an early period of  propagation  (a^  <  1.5).  Following this early period,  the
relative rate of growth of a   diminishes  and is roughly  constant with a  slope
of 0.2 for a   > 2.0.
     Measurements made  on fracture surfaces  created in the lab (Fig. 2.10)
show that  the average  slope a  la  for samples of moisture content 0.26 to

0.27 is between 0.19 and  0.21; similar to that predicted by the analysis (Fig.
3.6).   Lengths of undyed  zones are a  few mm less than that  of unwetted zones
predicted by the numerical analysis.   Those differences  are considered small,
however, and could be due to advancement of the dyed fluid by capillary
forces,  a process ignored in the  analysis.

     Other results of modeling conducted using the numerical analysis indicate
consequences of the development  of an unwetted tip on driving  pressure,
fracture length and aperture as functions of time.  These consequences are
highlighted by comparing results  from two analyses;  one in which an unwetted
tip was  allowed to develop as a result of differences between  fluid and
fracture velocity, and  another in  which the development of an unwetted  tip  was
prevented by requiring that  the fracture  remaincompletely filled  with
fluid.   Driving  pressure, aperture, and fracture length,  have been normalized
to the values of those  variables at the onset of propagation.   The normalized
variables include
                              /*= tit,                              (3.20a)
                                    in
                              a*, a/a.                              (3.20c)


                             5'- 816 .                              (3.20d)
                                    ofnc
where the  subscript frx indicates the value  at the onset of fracturing, and
the subscript o indicates the value at the origin.  All the starred variables
have  values less  than 1.0 during inflation, and values equal to 1.0 at the
onset of fracturing according to this definition.

      Growth of  an  unwetted tip has a  marked effect on the driving  pressure,
                                     90

-------
           2.00
      cd
           0.00
                 .0  1.0  2.0  3.0  4.0  5.0  6.0  7.0  8.0
                                      ~*
Figure 3.6  Normalized length of unwetted tip with respect to length of
           wetted fracture, from numerical simulation (solid line) and from
           lab experiments (data points).
                                91

-------
aperture and fracture length as functions of time (Fig. 3.7).  The  driving
pressure increases linearly during inflation and  then  abruptly decreases at
the onset of propagation when growth of the unwetted tip is prevented (Fig.
3.7).   Similar results were obtained using a wide range of fluid viscosity,
elastic modulus,  and critical stress  intensity if development of the  unwetted
tip was prevented.  The analysis that tracks die fluid front and results in  an
unwetted  tip shows  that driving pressure continues to increase after
propagation has beguo^.   Indeed, the slope of the pressure record very  early in
propagation (1.0 <  f    < 1.25) is  only  slightfy less  than during inflation.
Driving pressure continues to  increase until t = 2,  when it reaches  a maximum
and then  decreases at later times (Fig. 3.7).

     Those results have several implications with respect to the lab
observations.  They indicate that growth of an  unwetted tip can cause the
early  propagation of a  fracture to be stable, and that unstable propagation
takes place after the fracture  has grown to some finite length.   Records of
driving pressure  from the lab  experiments consistently show an early period of
stable  growth followed by a period of unstable  growth.   We conclude  that this
behavior results  from  the early growth of the unwetted zone that was  observed
on  the surfaces of fractures created in the lab.

     Clifton and others  (1976) performed experiments where they  prevented
fluid from entering  a  hydraulic fracture by lining a  starter  hole with a
flexible membrane.  Stable propagation  was observed when  the membrane  was in
place,  and unstable  propagation when the membrane was omitted.  Their results,
which  are supported by theoretical  analyses, confirm the conclusion that lack
of fluid penetration results in stable  propagation.         The  numerical analyses
also indicate that fracturing may occur before a noticable change in  the slope
of the pressure record.   This  is because the onset of fracturing itself may
cause only a minor  change in slope,  and it is only after the fluid  enters the
fracture  that the slope of the  record  changes markedly.  That  could  explain
the observation of short, incipient  fractures in  experiments  terminated
slightly before a change in slope of the pressure record.  It also could
explain the observations of Zoback and  others (1977), who measured acoustic
emissions from rock samples to detect the beginning of hydraulic fracturing
during lab experiments.  They observed increased acoustic emissions while the
slope of their pressure records was constant and concluded  that the  beginning
of fracturing was undetectable on the pressure  records.

     Fracture length and aperture  are affected  by the growth of the unwetted
tip (Fig.  3.7),  according to  the analysis.  The  rate of growth of the fracture
is reduced, whereas the  rate of dilation is increased when the unwetted tip is
allowed to grow. Analyses ignoring  the formation of the unwetted tip
therefore apparently will overestimate length  and under estimate aperture.
This has  important  implications in  designing  field applications of hydraulic
fractures  where length and aperture are important variables  affecting the
performance in delivery  or recovery.

     The actual magnitude of the  propagation velocity predicted when the
unwetted tip is 0.58 mm/sec,  whereas that predicted by the other  analysis is
0.81 mm/sec (determined from slopes of Figure 3.7  and /,_=  4.72 sec, which is
                                                         tn
                                    92

-------
                  1.25
                  0.00 ^
                     0.0  1.0  2.0   S.O   4.0   5.0   6.0  7.0   6.0
                   3.5


                   S.O
              
-------
 determined from Table 3.1  and eq. 3.2Sc).   Propagation velocities in the lab
 experiments,  determined by dividing half-length by duration of propagation,
 ranged widely from 0.2 to  1.05 mm/sec (Fig. 3.7a).   Both theoretical analyses
 predict velocities  that  are within the range of observed  velocities.  A linear
 regression of the  lab  data,  however, indicates that the average propagation
 velocity was  0.57  to 0.59 mm/sec  for moisture  contents between 0.26 and 0.28;
 the  analysis considering an  unwetted tip predicts this  velocity with
 remarkable accuracy,  whereas the other analysis overestimates it.        The  rate of
 growth of the unwetted tip  itself varies during propagation (Fig. 3.7).  The
 tip grows rapidly while  driving pressure increases during stable  propagation.
 Growth rate diminishes  during the transition from stable to unstable
 propagation, and it is  roughly constant later  in  the period of unstable
 growth.  Our experiments were unable to yield  data on  growth  rate of the  tip,
 however, Medlin and  Masse (1984; fig. 6) show that  the tip grows rapidly  at
 the  onset of propagation, and then the growth  rate diminishes and is roughly
 constant during further propagation.

      Driving pressure and aperture were calculated along the length of the
 fracture during inflation, at the onset of propagation,  and during propagation
 (Fig. 3.8).  During inflation driving pressure varies roughly linearly along
 the  fracture,  except in  the vicinity of the tip where the slope flattens to
 zero gradient and thus  zero flow at x = a..  At the onset of propagation

 driving pressure is nearly constant, varying  a few percent over the length of
 the  fracture.  As the fracture lengthens during propagation a zone of low
 driving pressure, corresponding to the unwetted  tip, appears and increases in
 length.

      Stable propagation early in the growth of a hydraulic fracture has been
 attributed to large viscous losses during flow in a narrow region at the tip
 (Zoback and others,  1977; Zoback  and Pollard,  1978).  The results of this  work
 show that,  at least for conditions  of the lab  experiments,  we do not expect a
 large decrease in pressure due  to viscous dissipation in the tip.  This is
 because dilation of the  fracture reduces  the  flow' rate  along the  length of the
 fracture,  according to eq. (3.10).   As the tip is approached  the  decrease  in
 aperture is  offset  by a  decrease in flow rate, so that  the pressure gradient
 determined  through eq. (3.8) is gentle.  A steeper gradient is obtained  if the
 dilation of the fracture is ignored.

      The pre-existing fracture has  a nearly elliptical form during inflation
 (Fig. 3.8).  During propagation the form is grossly elliptical, but it  is
 marked by a  slight upward  concavity centered on  the  leading edge of the
 pressurized  fluid.   Similar fracture profiles were obtained by Nilson  and
 others (1985; fig. 3).

 Analytical Solution


      The numerical solution described above is  cumbersome to use because it
requires as  much  as an hour of execution time on a desk-top computer (IBM 386,
 16 Mhz)  to complete  an analysis.   Making two  simplications in  the problem will
allow a simple analytical expression to be developed that captures much of  the
                                  94

-------
                  0.20         0.25

                            Moisture
0.30
0.35
Figure 3.7a.  Average propagation velocities from lab experiments.  Linear
          regression line suggests a slight dependence on moisture
          content.
                              95

-------
         1.3
         1.0
   *,0.8

   DH

        0.5
        0.3
        0.0
        3.0



        2.5


        2.0



        1.5


        1.0


        0.5
        0.0
                                 3.4
                                                    13.8
           0.0       0.5      1.0      1.5      2.0      2.5      3.0
t = 13.8
          0.0      0.5      1.0      1.5       2.0       2.5       3.0
Figure 3.8.  Driving pressure and apertures as functions  of fracture length

            during inflation  (t   <  1.0)  and propagation.   The symbol x marks
            the end  of a fracture.
                                    96

-------
 essence of the numerical solution but requires  only a few seconds to compile
 and execute.  The results of the numerical  solution showed that  driving
 pressure varies only slightly within the fracture, so that  one of the
 simplifications will be to assume  that the driving pressure within the  wetted
 fracture is constant.  This is the same  as ignoring the pressure losses due to
 viscous  dissipation during flow.  As a result, the analytical solution will be
 unable to  predict the growth of the unwetted tip.  Both  the  numerical solution
 and the lab data, however, suggest that the length of the unwetted tip is
 roughly linearly related to the length  of the  wetted zone.  The other
 simplification, therefore,  is to assume that the length of the unwetted dp is
 proportional to the  length of the wetted pan of the fracture.

      The  geometry of the problem is the same as that used  for  the numerical
 solution and described in a previous  section (Fig. 3.4).   The fracture is
 filled  with fluid to a point b along the x axis, and the pressure  p of  the
 fluid is  uniform along 0  * x £  b.   Fluid  pressure along an unwetted tip, the
 interval b ^ x  £ a, is zero.  During inflation the entire fracture is  filled
 with fluid, so a =  b before propagation begins.   When the unwetted  tip
 develops during propagation b will be less than a, with the length of  the
 unwetted zone equal to o  - ft.  The relationship  between a and b is assumed to
 follow that observed in the experiments and  expressed in eq.  2.4.

      Fluid pressure within the fracture  is opposed by a confining stress a
 acting normal to the plane of the fracture.  Accordingly,  the  driving pressure
 is (p - o) on  0 ^ x £ b, and -a on b <  x  s  a.

      Static analyses will  be used  to obtain driving pressure,  aperture and
 half-length.  Those analyses will be made functions of time by treating  a
 series of fractures of slightly different volumes, and then determining  the
 time since injection began by dividing fracture volume by volumetric rate of
 injection. This quasi-static approach is the same  as  the one used in the
 numerical  analysis.

      The fracture will be assumed to be in  equilibrium,  so that  the stress
 intensity must equal the critical stress intensity K  throughout propagation.

 In the analyses of Kristianovich and  Zheltov (1959), and Geertsma and  de Klerk
 (1969) equilibrium propagation is obtained by adjusting either the length  of
 the unwetted tip, or the driving pressure to eliminate stress concentration at
 the tip;  they  set K   to zero.


      A  solution to the problem can be obtained by  superimposing solutions for
two  different loading conditions (Fig.  3.9): a.)  the load equals p  along 0  s  x
 £ b and 0 along b  <  x s a; and b.) the  load equals -a along 0 £  x £  a.   A
solution  to similar boundary conditions was obtained by Barenblatt (1962).
The  mode  I  stress  intensities  K^ volumes  V, and  apertures at the origin  5 of
fractures for conditions a and  b are (Tada and others, 198S)
                                   97

-------
                             nun
                        TT! I f TT! t f
                                              illllllllk
                                              Illllllllt
Figure 3.9.  Loading conditions used to develop analytical model.
                            98

-------
                     K^= pi na   (2/n) sm'lb/a                     (3.21a)
            V=  (4pa2w/£) [sin'Wi  + (bla) /I - (bla)2           (3.21b)







              5  = (ZnpalnE) [sm'lb/a  + bla cosb'la/b]             (3.2Ic)
               CU






                           JT=  -av IE                         (3.21e)
                            D
                            $  =  ^aaiE                            (3.21f)
                             OD



where the E, the plane strain modulus, is £'/(!  - u2), and £' is the elastic

modulus and u  is Poisson's ratio.   Adding the two  solutions  and introducing

the driving pressure P  from eq (3.2) yields
                   K = / na  [PO - a (1 - 0)]                   (3.22a)
                     I             d






                                                                   (3.22b)
                                                                   (3.22c)




where the terms 6, y and 0 depend only on  the length of the unwetted zone




                        0 =  (2/7T) sin'1 bla                        (3.22e)
             \p = (2ln) [smlbla  + bla / 1  -  (bio)* ]             (3.22f)
                    0 = sm'b/a +  Wa cosha/^                   (3.22g)



The rate of injection Q is constant,  and assuming that both the fluid and  the

pumping system are incompressible,  and that none of the fluid flows out of the

fracture, then


                               V =  Qt                              (3.23)


where t is the time since pumping began.  Substituting eq. (3.23) into eq.
                                  99

-------
 (3.22b)  yields



                  / = (2ita2w/QE) [Pw - (7(1  -  \i/)]                  (3.24)
                                     d


 For convenience driving pressure, time, half-length and aperture at the origin

 will be  normalized  in terms of those variables  at the onset of fracturing, so
                                                                    (3.25a)
                              a .  = at                              (3.25b)
                               char    i                              x     '




                        I fc  =  2rra2 wP  IQE                        (3.25c)
                         char      c    dc *                         *     '




                           d.  =  APalE                           (3.25d)
                            char     dc c                                   '



 Dividing eq.  (3.22a) by eq. (3.25a) and solving fpr normalized  driving

 pressure P  in terms of normalized half-length a
                              -in ,   _ /*    A, 14                      £ jg)
Similarly
                                                                     (3.28)
                       ^t


                      ff = ff/P. = ff (nay  IK                       (3.29)
                               dc         i     b                          '
                              fl  = fl/«.                              (3.30)



     In the plots that  follow normalized driving pressure, half-length  and   *

aperture were obtained as functions of  dimensionless time by incrementing a

and using eqs. (3.26) through (3.30).




Effects of Growth of an  Unwetted Tip




     According to the  solution derived  above driving pressure,  half-length,

and aperture as functions  of time depend on the growth of the unwetted tip,

manifested as the ratio b/a, and  on the magnitude of dimensionless confining
                                     100

-------
 stress.  Observations described in Chapter Two suggest that the length of the
 unwetted tip is linearly  related to the wetted length after the fracture has
 reached some critical length a  .  Rearranging eq. (2.4)
                              cr
                     b/a  =  (ma + al(a + ma)                     (3.31)
to
 where m is  the  ratio of wetted to unwetted length.

      Results from  the lab tests  showed that m and a   are roughly constant  for
                                                     cr        °  '
 a soil of a particular moisture content, but they both increase as moisture
 decreases.  For  the purpose of this section, I  will assume that  m and a are

 material constants that govern* the growth  of  the  unwetted tip according to eq.
 (3.10).  The effects of m, a   ,  and a  on dimensionless driving pressure,

 half-length,  and aperture are  shown in Figures 3.10 through 3.12. In those
 Figures, the quantities used to form^he dimensionless groups require that
 inflation of  the  slot occurs during (/   <  1.0), and propagation occurs during

 <£ > 1-0).
      When m = oo the unwetted zone is  absent  and eqs. 3.26 through  3.28 reduce


                             P* = *-in                             (3.32a)


                              c*  = r"3                              (3.32b)

                              6* = r*1/3                              (3.32c)

 for values of /  >   1.0.  Driving pressure shows  a sharp decrease following the
 onset of propagation for m = o> (Figure 3.10).   The rate of growth of the
 fracture  (slope of Figure  3.1Gb)  is greatest at  the onset of propagation,  and
 diminishes slightly  with  time.  The rate  of dilation (slope  of  Figure 3.10c)
 is greatest before propagation begins,  when the slot is dilating.  Aperture
 continues to increase during propagation, but  the  rate at which it does so
 diminishes with  time.

      Finite values of m change the form of the pressure  record significantly;
 driving pressure  increases slightly early in  propagation,  and then decreases
 at later times  (Fie. 3.10).  Thus, the presence of the unwetted tip  results in
 an early period of stable  propagation  followed by  an unstable period.  The
period of stable  propagation is brief for large  values of m, but  increases in
duration as m decreases  in value.  The rate of growth of the unwetted  tip,
embodied in m,  therefore tends to stablize  fracture growth.

      Fracture  length and aperture are also affected by m  (Figure 3.10).   The
rate  of growth of the fracture diminishes as m decreases.  The decrease in
growth rate is offset by an increase in the dilation rate, so that
conservation of volume is maintained.  Accordingly, the fracture becomes
shorter and fatter as the relative  length of the unwetted tip increases.
                                   101

-------
                         0129466789 10
                         0123458789  10
                       3.5

                       3.0

                       2.5

                   *   2.0
                   *0
                       1.5

                       1.0

                       0.5

                       0.0
                         0123466789  10
                                       t*
Figure 3.10  The effect of parameter m on dimensionless driving pressure,
            half-length  and aperture. Other parameters are constant: a _
            1.0, a*  =  0.5.
er
                                  102

-------
      Experimental data suggest  that in some samples the unwetted tip develops
 after the fracture reaches a critical length a  .  Observations  early in the

 growth of fractures in those  samples are unavailable, so we will simply  assume
 that the unwetted tip  grows slowly from the  onset of propagation and more
 rapidly when the critical length  is reached.  This suggests that eq.  3.31
 should be modified for the period beginning at  the onset  of propagtion and
 ending when a = a  .   We will use
                    cr

                    b/a  =  (mf  +  1.0)/(fl + Wjfl)                    (3.33)

 for a.  < a  <  a , which is the same form as  eq. 3.31.  The parameter m  was
     i           cr                                                         1
 set equal to three times m to determine curves  in Figure  3.11.

      A flattening of the slope of the driving pressure record  is the principal
 effect of increasing a   to values greater than the half-length  of the slot.
                      cr
 The change in  slope ocqurs when the unwetted  tip begins to grow  according  to
 eq.  (3.31), so that as  a   increases the point when the slope flattens occurs
                        cr
 at progressively latter times (Fig. 3.11).  Increasing  the  critical length
 results in an increase in fracture length, which reflects early growth that is
 unimpeded by the constraining effects of the unwetted tip.  As we saw in the
 previous  figure, an increase in growth rate  is accompanied by  a decrease in
 dilation rate (Fig. 3.12).

      The normalized confining stress affects the forms of the  records, as
 shown  in  Figure  3.12, because it is  the driving pressure over the unwetted
 zone.   The duration* of the stable period early  in propagation increases with
 the magnitude of a  .  The rate  of growth, of the fracture is reduced, and the
 rate of dilation enhanced  by increasing a  (Fig.  3.12).

      The  same confining  stress was  used in all the lab experiments, so there
 are no data on the effects of confining  stress from this study.   However,
 dimensionless confining stress is a function  of  both K  and a, through eq.

 (3.29). We saw  in  a previous section that  K^ of  a  soil increases as moisture

 content decreases.  The effects of increasing a   shown in Figure 3.12,
 therefore,  could occur either by decreasing K^, say by increasing water
 content, or by increasing  the slot length.

      Results from the proceeding analysis indicate  that the unwetted tip makes
 the soil appear artificially  tough.  It causes driving pressures to increase,
growth rates to diminish and dilation rates to increase.   The magnitude of
these effects depend on the length of the unwetted zone  relative to the
fracture length, when the  zone begins to grow, and the magnitude of the
confining  stress acting normal to the zone.
                                 103

-------
                   0123466789 10
                   0123466789  10
*0
                 3.5

                 3.0

                 2.5

                 2.0

                 1.5

                 1.0

                 0.5

                 0.0
                   0123466788 10
                                 t*
Figure 3.11   The effect of parameter a   on dimensionless driving pressure,
                                       ei
            half-length and aperture. Other parameters are constant: m = 6,
            ml = 3m, a = 0.5.
                                  104

-------
                1.50 r—.	1	•
                   0123466789  10
                 3.5,,—,	,	,	,
                   0123458789  10
                                            1.0

                                      0.5  °-75  •
                   0123456760  10
Figure 3.12   The effect of parameter a  on dimensionless driving pressure,
          half-length and aperture.  Other parameters are constant: m  =  6,
          a* -  1.0.
           er
                                 105

-------
Comparison With Results  From The Lab


      Variations in forms of records  of  driving pressure  observed  in the lab
experiments can be explained using the  analysis described  above.  To do so, we
must first include effects of a defonnable pumping system on  the form of the
pressure record.   Volumetric distortion  of the pumping system causes fluid to
either be stored when pressure  is increasing or released  when  pressure is
decreasing, so that the flow rate into the fracture will differ from the flow
rate of the pump.   Distortion can be included in the analysis  by  modifying
eq. (3.22b) to
         V  = V + V= (2ita2w/E)  [Pw - a(\ - 
-------
         1.50
        0.00
             0
50
100
150
200
                       p  (kPa)
Figure 3.13  Volume stored in pumping system as a function of fluid
         pressure fluid pressure. Dashed line is second-order regression.
                         107

-------
                                                                    (3.39b)
                            v            i

which reduces  to  eq. (3.28) when C = 0.  Values of C^ ranged from 0.02 to 2.0

during the experiments, depending on  the modulus of the  soil and the slot
length.

      Dilation of the experimental apparatus affects the  forms of pressure
records by changing the time scale according  to eq.  (3.39).   The
characteristic time increases, reducing  the slope of the record during
inflation, as the apparatus  becomes more compliant.   The form of the pressure
record during propagation is especially affected early in propagation  when a
large value of  C  causes  the record to steepen (Fig.  3.14).  As a  result,  the

slope of the record during  early stable propagation may be nearly
indistinguishable from the slope during inflation of the starter  slot.   The
term C  also results in a steepening of the  record early in the unstable
period of propagation.  Steep slopes during  this period are commonly observed
in both field and  lab investigations of hydraulic fracturing of rock,  where
large moduli could result in large values  of C.
      The rates of growth and dilation of the  fracture both increase
dramatically, indicating an  increase in the volumetric growth rate, as
C increases (Fig.  3.14).  The volumetric rate  of growth of the fracture

increases, even though  the  pump is running at a constant rate, because fluid
is released from storage in the pumping apparatus when driving pressures
diminish during unstable  propagtion.  Accordingly, the flow rate into the
fracture  will be greater than the flow  rate of the pump during unstable
propagation.  This conclusion is confirmed by observations of minute bubbles
in glycerin flowing through a translucent tube leading to the fracturing cell;
their velocity was less  during inflation and  stable propagation than it was
during unstable propagation.

      The revised  expression for /  in  eq. (3.39)  was used to compare results
from the analytical solutions to  the forms of records from the experiments.
Results from the  lab experiments showed that the form of the record of  driving
pressure depended on both  the moisture .content and  the slot length.  As
moisture content decreases,  the driving pressure required to initiate
fracturing increases, the period of stable  growth early in propagation
shortens, and the  downwardly concave shape of the record becomes  tighter.
Similar changes in the  records occur when the slot length decreases.

      Records of driving pressure are similar to results from the analytical
solution  (Figs.  3.15 and 3.16) using parameters listed in Table 3.2.   The
parameters were estimated  from lab data, except m which was adjusted to  yield
the best fit to  the data—using m estimated from lab  data  yields driving
pressures that  greatly exceed the ones that were observed.  The general forms
of the records  for samples  of different moisture content are  predicted by the
analysis.  The  driving pressure at the  break in slope and  the duration of the
early period  of stable growth are within a few percent of those observed.  The
                                    108

-------
                          0.00
                             0.0    1.0    2.0    3.0   4.0    5.0
                           5.0
                           4.0
                           3.0
                           2.0
                           1.0
                           6.0



                           4.0



                      *   3.0

                      <0

                           2.0



                           1.0
                           0.0
                                    10.0
                             0.0    1.0    2.0    3.0    4.0   6.0
10.0
                             0.0    1.0    2.0    3.0    4.0    6.0
Figure 3.14.  Driving pressure,  half-length  and aperture as functions of
             time  and compliance of pumping  system,   m:  10; a  :  1.0;
                                      109

-------
    100
      75
 cti
DH
      50
      25
       0
                   -T	1	r
                              -i	1	"
          0
50       100       150
   Time  (sec)
   Figure 3.15.  Records of driving pressure as a function of time from
           experiments and from analytical solution (dashed) for samples of
           various moisture contents.
                            110

-------
    0   25  50   75  100  125  150
            t (sec)
the length of the starter dot.
                                 in
             111

-------
 slope of the records during unstable propagation steepens with increasing
 moisture  content in both the analysis  and the lab data

      The effect of changing the slot length is predicted by the analysis (Fig.
 3.16).  Driving pressure at the break in  slope increases, the duration of
 stable propagation  decreases, and the  slope during unstable propagation
 increases as the length  of  the  initial  slot decreases.

      The analysis  overestimates the driving pressure during unstable
 propagation in  all  the examples (Fig.  3.15  and 3.16).  Small amounts of the
 injection  fluid leaking out  of the edges of the sample, due to imperfect seals
 between the samples and the apparatus, probably caused the driving pressure
 during propagation to be less than  that predicted by the analysis.


 TABLE 3.2.   PARAMETERS  USED  I N ANALYSES
Curve

Moi s t u re m K
kPa cm1'2
/
c
( sec)
C a.
c i
(cm)
Fi



Fi


g.3.15
a
b
c
g.3.16
a
b

0
0
0

0
0

.29
.26
.23

.27
.27

75
30
50

30
30

22
41
186

63
63

28
50
150

100
95

0
0
5

0
0

.05
.30
.0

.088
.354

1
1
1

1
0

.22
.22
.22

.22
.61
      In  both Figure  3.15  and 3.16 the parameter m  was selected to yield
results that matched  experimental  data.  Values of m are an order of magnitude
greater than those obtained from experimental observations, so that the length
of unwetted  tip required for the analytical solution to resemble the observed
pressure records is considerably  less than the length  of the undyed zones on
fracture  surfaces.  This suggests either that the effective length of the
unwetted tip may be shorter than  the  undyed zone observed on fracture
surfaces, or  that the driving pressure acting over  the unwetted tip may be
greater than the negative confining stress.


Comparison with Other Solutions


      Several other two-dimensional analytical  solutions have been published
describing the growth of a hydraulic fracture  having  a geometry resembling the
one used here.  Khristianovich and Zheltov (1959) solve  the case where the
fracture  has  an  identical geometry, but the distribution of driving pressure
differs slightly from  the one used  here.  They assume that the driving
pressure is a maximum  at the center of the fracture  and diminishes due to
viscous losses to 0 at a point  b, which is less than the total length of the
                                    112

-------
fracture.  The variable driving pressure  is approximated by a statically
equivalent driving pressure that is uniformly distributed  and acts over 0 & x
£ 6' (b' is  obtained from Khristianovich  and Zheltov, 1959; eqs.  6 and 7).
Thus, the driving pressure they use is uniform but its magnitude depends on
fluid viscosity, whereas the pressure  used  here is uniform but independent of
viscosity.

     The analysis of Kristianovich and Zheltov includes a zone at the tip  of
the fracture  where fluid pressure is equal  to the fluid pressure in  the
confining material.  In contrast, the  analyses used here  assumes that the
pressure along the tip zone is zero, and is independent of the pressure  of
pore-fluid in the soil.  Our assumption is  certainly valid for
partially-saturated soil, where the  pressure of  the pore fluid is less than
zero.   The assumption seems reasonable for saturated  soils of low permeability
as well.  This is  because  the dilation of the crack itself must reduce the
pressure within the tip, and the tune required for those low pressures to
equilibrate with pore-fluid  pressures would be  longer than the few seconds
that the tip  remains unfilled by injection fluid.  In any  case, the assumption
used here marks one end  and the one of Khristianovich and Zbeltov  marks the
other end of the range of  possible pressures within the tip zone.

     Geerstma and DeKlerk (1969) solve a problem whose geometry  is similar   to
the one  used here, except  they provide for a distribution of pressure  within
the fracture  that varies from a maximum  at x =  0 to zero at x = b.  They
obtain the pressure distribution by integrating  over the interval 0  s x  b  an
equation for flow through  a narrow rectangular channel,  and assuming that the
pressure within the fracture is zero over b < x s a.      Both Geerstma and
DeKlerk, and Kristianovich and Zheltov introduce  an unwetted  zone remove the
stress singularity at the fracture tip.  This concept, which  is applied to
brittle fracture by Barenblatt (1962)  and to ductile fracture by  Dugdale
(1963), is mechanically different from the one used in this work.   We  show
that the unwetted tip can develop due to a difference  between the velocities
of the fluid  and  the fracture, and the length of the unwetted zone is
therefore related to differences between  the two velocities.   In the other
model,  the length of the unwetted zone  is related  to the magnitude of JTfc and

is independent of the relative velocities.  These differences are subtle, but,
inasmuch as the  size of the unwetted zone strongly affects the length and
aperture of a fracture, the mechanism for growth  of that zone plays  an
important role in predicting fracture  dimensions.

     Spence and Turcotte  (1985) built on  the  work of Geerstma and DeKlerk by
deriving an analytical solution  that includes effects of both viscous
dissipation of pressure in  the fracture and finite fracture toughness.  The
general solution of Spence and Turcotte  requires numerical integration;
however, for specific cases where the system  is dominanted by either fracture
toughness or viscous dissipation the solution reduces to  algebraic
expressions.   Predicting the migration of igneous sheet intrusions was the
primary goal of Spence and Turcotte, and they have omitted an investigation of
the pressure distribution at the tip of a  hydraulic  fracture.

     Perkins and Kern (1961)  present a solution for  a long, rectangular
                                      113

-------
fracture that is elliptical in shape in sections taken normal to the long
axis. They obtain a pressure distribution by  using an equation for flow
through a narrow elliptical channel, and assuming the driving pressure is zero
at the  end of the channel.  Nordgren (1972) modified the solution of Perkins
and Kern (1961) to include effects of dilation and leakoff on  the pressure
distribution in the fracture.  Both Perkins and Kern (1961)  and Nordgren
(1972) ignore the problem of stress concentration at the tip of the fracture,
so conditions of equilibrium propagation are not  explicitly satisfied.

     Medlin and  Masse (1984) and  Geerstma and Haafkens (1979) point  out that
the analytical solutions described above can  be written as  power law functions
relating driving pressure, half-length, and  aperture to time
                             P=  c f                            (3.40a)
                              0

                             a - d £                             (3.40b)


                             8 = e f                            (3.40c)

The solution derived  for this  work reduces to  a  set of power law
functions  by setting bla =  1.0, which  is the same  as setting the
length of  the unwetted tip to zero  (eq. 3.32).  Values of the
exponents given by the present  solution (using bla  = 1.0) are
identical to those derived by Geerstma  and DeKlerk, and by Spence and
Turcotte (Table 3.3), and they are similar to values from
Kristianovich and  Zheltov.


—TABLE 3.3.   EXPONENTS FROM VARIOUS ANALYTICAL SOLUTIONS
_ JL JL JL
Kristianovich and Zheltov**
Kristianovich and Zheltov
Geerstma and DeKlerk
Spence and Turcotte
Perkins and Kern
This work (bla = 1.0)
P
-0.3
-1/4
-1/3
-1/3
+1/5
-1/3
JL
0.6
1/2
2/3
2/3
4/5
2/3
£
0.4
1/2
1/3
1/3
1/5
1/3
* Values  cited by Medlin and Masse (1984)
** Based  on eqs.  (33), (36) and (45) of Kristianovich and Zheltov (1959)
The values  are markedly different from those given by Perkins  and Kern
(1961), however. In particular, the solution  of Perkins and  Kern
(1961) yields  a positive value  for /?, whereas  the others yield
                                     114

-------
 negative values.   Positive value of /? indicates that driving pressure
 increases throughout propagation, whereas pressure was observed to
 decrease in the experiments suggesting that a negative value of /S  is
 appropriate.

      Forms  of the constants c, d, e  from the present solution are
 similar to those  given by Spence and Turcotte for a fracture  whose
 resistance  to propagation due to fracture toughness is much   greater
 than  resistance to now (Table 3.4).   Fracture  toughness enters  as  the
 same power  for  both
   -TABLE 3.4.  COMPARISON OF CONSTANTS
 Constant         Spence and Turcotte                     This  paper

                                          1'3           r    .          ,  il/3
                                                         2K*w (1  - o2)

                                                                QE'



                            .1/405/4     'W
    d     0.5599
                                      3/41
    ,      Q.5685 P
-------
 time on a log-log scale.   They found  that apparent values of x and £
 decrease, whereas ft increases as confining stress increases  (Figure
 3.17).

      The experimental apparatus of Medlin  and Masse  resulted in
 conditions resembling two-dimensional plane strain, so the analysis
 described above  was used to examine  the effect of confining  stress on
 apparent values of the exponents.  The ratio bla is independent of a,
 but  increases with confining stress according to Medlin and Masse
 (1984; fig.  7).  Figure 3.17. is  based  on figure 7 in  Medlin and
 Masse, but values of bla  have been extrapolated for the purpose of
 this analysis.  A driving pressure at failure  of approximately  500 psi
 is apparent from figure 3 (Medlin and Masse,  1984), and this value
 will be used to determine a  for all the calculations--^,  and slot

 length are assumed  to be  constant.  The apparent value, of ft  jvas
 determined  by linear regression  of the logarithms of P.  and t , and

 the  other exponents were  determined in a  similar manner.

      The analysis is able  to predict the experimental values of ft
 throughout  the range of confining pressures.  At relatively low
 confining pressure, ft increases with confining pressure, but it
 remains  roughly  constant at intermediate to  high confining pressure.
 Figure 3.18 shows that  increasing a   tends  to  increase  ft, whereas
 decreasing the length of the unwetted  tip,  with a  according  to Figure
 3.17, tends to decrease ft.   At confining pressures greater than 14 MPa
 these effects apparently balance  one another according to the  data of
 Medlin and Masse and the results of the analysis.  We conclude that
 those changes in the form  of the pressure record, expressed as changes
 in ft, can be explained by  the effect of confining stress  acting on the
 unwetted tip.
     The analysis yields values  of x and £ that exceed those observed
in the experiments of Medlin and Masse.  Flow out of the fracture and
into the enveloping  material, or leakoff, is a possible explanation of
the discrepency.  The present analysis is unable to  account for
leakoff,  but the analysis of Nordgren (1972) shows  that leakoff will
decrease the values  of x and e.  Moreover as  Medlin and  Masse (1974)
point out,  conservation of mass  requires that £  + £ = 1.0 if no
leakoff occurs, and  that sum will be less than  1.0 if leakoff does
occur.   At equal confining stress, sums of x and £ determined
experimentally by Medlin  and Masse (Fig. 3.18) are all less than 1.0
and  those  sums decrease with increasing confining pressure.  This
information substantiates the conclusion that leakoff is a possible
explanation for why the data of Medlin and Masse  on x and £ are less
than the results of the analysis.


DATA QUALITY


     The objective stated  in the  Quality Assurance  Project Plan (memo
                                   116

-------
Ctf
    1.00
   0.98
   0.96
   0.94
   0.92
                   4.0
       (J
       8.0
          12.0
                10
20
30
40
                         a  (MPa)
             16.0
50
Figure 3.17  The ratio of wetted to total length of a fracture as a function
          of confining stress.  Data from Medlin and Masse (1984).
                            117

-------
X
  -0.35
0.70


0.65


0.60


0.55


0.50


0.45
(1)
0.40




0.40


0.35


0.30


0.25


0.20
                  10
                  10
20
30
                         20
           30
                               a  (MPa)
                                                                  16.0
                                                   .1
40
           40
                                                             50
                                                                   O
50
           60
    Figure  3.18.  Variation of exponents in eq. (3.40) as functions of confining
               stress.  Data from Medlin and Masse  (1984), solid line  from
               analytical solution.     1Q
                                    llo

-------
 submitted to David L. Smith, 3 October 1988), was to make measurements
 of Kie  that can be duplicated  to within  IS percent.   In Chapter Two,

 we  showed the results of four tests conducted on similar samples.  The
 three of  the four values of initiation pressure were  tightly
 clustered, they were within 2.5 percent  of their average value.   The
 fourth  measurement was less than the other three.  The average of all
 four measurements is 25.4 kPa and that fourth value is 22.8 kPa,  which
 is 10 percent less than the average.  Critical stress intensity is
 determined directly  from initiation pressure,  and so it will  have the
 same range of values as described above.

      Earlier in this chapter, results showed  that doubling  the slot
 length  caused variations of K  that  averaged 10.4 percent.   The

 critical stress intensity both increases and decreases as the slot
 length  increases,  indicating that the variations that were observed
 reflect  experimental error,  rather than a systematic  dependence  on
 slot length as  reported by  Ouchterlony (1982).

      Critical stress  intensity measurements can be duplicated to
 within  15 percent of an average value using the methods described  in
 the  text.

     The principal measurements,  the methods of obtaining them, and
 the  accuracies of those methods are presented in Table (3.5).
TABLE  3.5. EXPERIMENTAL MEASUREMENTS

Measurement              Method              Accuracy

Sample size                scale                ±0.1 cm
Length of slot             micrometer          ±0.001 cm
Density                    Weight/Volume      ±0.01 gm/cc
Confining pressure         Gauge               ±0.005 kPa
Injection pressure          Transducer          ±0.002 kPa
Injection rate              pump calibration     0.00003  cc/sec
Ar                         calculation           0.007 kPa cm"2
      Pressure gauges and transducers were purchased at the beginning
of the project, and were maintained  in good  working order.   Gauges
used to measure confining pressure were calibrated using a pressure
transducer.  The transducers were calibrated  at the factory.

     The precision of the critical stress intensity measurements is
impossible to  assess because published values of that parameter are
unavailable for soil.

     All the tests that we set out to conduct were accomplished during
the study.  Completeness was  100 percent.
                                  119

-------
      The results are expected to represent K  determined using

 similar techniques of sample preparation, experimental apparatus, and
 method of loading.  Experimental methods that facilitate  detecting the
 onset  of  fracturing,  perhaps by acoustic or optical methods,  probably
 will show that fracturing occurs slightly prior to  the break in slope
 of the injection record, which was used in this study to infer
 fracturing.  As  a  result, critical stress  intensity measurements
 obtained  by those methods  will be less than those cited here.


 DISCUSSION


      A satisfactory  analysis of hydraulic fracturing of soil should
 yield a method of predicting the  driving pressure at the onset of
 fracturing as well as the forms of pressure records during
 propagation, and it should explain the development of  essential
 features of a fracture, according  to requirements  put forth in the
 introduction of this  chapter.  Elements of linear elastic fracture
 mechanics were used to develop various analyses  to test  against those
 requirements.

      The critical stress intensity  at the  onset of fracturing is
 independent of the length of the  initial slot, and  appears  to  be a
 material property of the soil  used in  this study that will serve  as an
 acceptable predictor  of fracturing. Moreover,  theoretical estimates
 suggest that the size of the process zone at the tip of a fracture in
 the Center Hill  clay ranges from  one to a few mm, which is small
 compared to lengths of initial slots and  further supports  the
 applicability of K^ as a predictor  of  fracturing.


      The critical stress intensity  of the  Center Hill clay ranged over
 an order  of magnitude for various moisture contents and durations of
 consolidation, so that that lab data will  only  yield rough estimates
 of conditions in the  field.   An in-situ method of  determining K

 would yield  values more suitable  for  field applications.
 Nevertheless, the lab results are significant  in that they show that
 ATie is  a meaningful  property describing fractures in clay.

     The development of a hydraulic fracture in the Center  Hill clay
 at moisture contents less than 0.21 includes the growth of an unwetted
 zone at the tip of the fracture.  A theoretical analysis of coupled
 fluid  flow and elastic deformation shows  that the velocity of fluid
 near the  tip of a fracture is less  than the velocity of the fracture
itself,  thus providing a mechanism  for the development of the unwetted
tip.  Tracking the position  of the  fluid front  by the theoretical
model yields lengths of the unwetted  and wetted parts  of the fracture
 that are similar  to lengths observed on the surfaces of fractures
created in the lab.   Detailed  analyses of the rate  of growth  of  the
                                120

-------
 unwetted  tip as a function of moisture content (e.g. Figure 2.10) are
 the subject  of  future investigations.

      An early  period of stable propagation, characterized by
 increasing driving pressures,  was observed  in nearly every lab test.
 Stable  propagation  early in the growth of a hydraulic fracture can be
 explained  by an increase in resistance to propagation resulting from
 growth of an unwetted tip.   This explanation is confirmed by results
 of both numerical and analytical  solutions to fracture growth.   An
 alternate explanation of stable propagation  resulting  from large
 viscous losses due to flow  in a narrow region near the tip of a
 growing fracture (Zoback and others,  1977; Zoback  and Pollard, 1978)
 could also explain the lab data but it  is inconsistent with the
 results of the numerical analysis.

      Results of the numerical solution using parameters typical of the
 lab experiments show that  the fluid  pressure within  a  fracture is
 nearly  uniform.  This suggests  that  a  simple analytical  analysis,
 which  assumes uniform pressure, could provide insight  into the general
 behavior of  the fractures.   Pressure  records from the lab can be
 reproduced usin| the analytical model  by selecting an appropriate
 function describing the growth of the  unwetted zone.  The analytical
 model  also provides an explanation of the dependence of the form of
 pressure records on  confining stress, which was observed in  hydraulic
 fracturing  experiments conducted  using rock specimens  (Medlin and
 Masse, 1984).

      The  methods of linear elastic fracture mechanics  meet the
 requirements of an acceptable model of hydraulic fracturing of soil.
 Modifications of the analysis to account for large-scale  plastic
 yielding at the  fracture tip appear to be unnecessary.  The present
 methods could readily be adapted to account for appreciable plastic
 yielding during hydraulic  fracturing (e.g. Biot and others, 1982),
 although measurements that are more  detailed than those used here
 would be required before such methods seem warranted.

      Much of the published analyses on hydraulic fracturing of rock
 appear  to  be generally applicable  to  hydraulic fracturing of soil.
 Analyses that track the position of the wetted front, such as those by
 Abe and others (1976), Roegiers and others (1982), Griffiths and
 Nilson  (1986), or Nilson (1986), appear to  be particularly well-suited
 to predicting hydraulic fracturing  of  soil.

      Analyses that  assume  that the fracture is completely filled with
 fluid will overestimate lengths and underestimate apertures as
 compared with  analyses that allow for the development  of an unwetted
 tip.   Inasmuch  as fracture length  and  aperture control the performance
 of a fracture in delivery and  recovery,  it is clear that understanding
processes  at  the fracture tip will improve the ability  to design field
 applications of hydraulic fractures  in either soil or rock.
                                  121

-------
                                 SECTION FOUR

                 SETTING AND DESIGN OF THE FIELD TEST-1988

       The 1988 field test was conducted at a site 10 km north of downtown Cincinnati on
 the western side of the valley of Mill Creek, a southerly-flowing tributary of the Ohio
 River. The site is on the southeastern side of an area owned by the ELBA Company, who
 currently uses it as a municipal landfill.

       Investigation and preparation of the test site took place during several months in
 Spring 1988.  The actual test, however, only required 12 hours and was completed in one
 day, 15 June 1988. Ten hydraulic fractures were created during the test. Field evaluation
 of the results of the test took place during the six weeks following the test.


 SITE CHARACTERISTICS

       The vicinity of the test site is an area of gently sloping ground bounded on the
 northwest by a vertical face and on the southwest by moderate slopes into an excavated
 trough (Fig. 4.1). The southern side of the trough is bounded by a north-facing vertical cut,
 the eastern end of which is shown on the southwestern corner of Figure 4.1. The ground
 surface at the time of the test had been excavated during operations at the landfill to
 several meters below the natural ground surface.

       The site itself is an elongate strip trending N20E and roughly 80 m long. The
 northern end of the site is on a gentle, south-facing slope, whereas the southern end of the
 site is on level ground (Fig. 4.1).  The slope of the ground surface is important because it
 apparently affected forms of the fractures.


 Geology

       Glacial till, probably of Dlinoian age, underlies the test site. The thickness of the till
 is unknown, but it is more than several tens of meters. Three stratigraphic units were
 identified in the till based on exposures in vertical faces near the site, on boring logs, and
 on exposures in trenches cut in the test site (Fig. 4.2 and 4.3). The upper till unit, Unit 1, is
 0.5 to 2 m thick and consists of massive light brown-grey clay and silt containing 5 to 15
 percent rock fragments. Coarse gravel, cobbles, fragments of limestone and organic matter
 are disseminated through Unit 1.

      The middle unit, Unit 2, is dominantly flat-lying beds that fine upward from coarse
gravel to clay. A typical bed consists of orange-brown cobbles and gravel at the base that
grades upward to orange-brown coarse and fine sand, to light brown silt, and to light
brown clay at the top. The graded beds range in thickness from 0.1 to 0.4 m, and in most
exposures are several m to several tens of m in lateral extent. Locally, the forms of the
beds are irregular and highly contorted with large changes in thickness occurring over
several m. An example of the irregular bed forms is shown in Figure. 5.9. Massive beds of
grey clav and silt are locally interfingered with graded beds. The total thickness of Unit 2
ranges from 3 to 7 m.
                                   122

-------
                     Scale
                    5   10
                    nrwtars
                             15
\
                o Well location

                13 Well I.D.

                '»n Vent

               	Strati graphic
                  contact


             Topographic contours  In
               meters above  MSL
                                                                        Location of
                                                                        hydraul Ic
                                                                        fracturing
                                                                        equipment
Figure 4.1. Topography and locations of boreholes, vents, and hydraulic fracturing
               equipment at the ELDA test site.
                                         123

-------
                Unit 3: Mouiv, iiity day

                o Well location
             -- Stratlorophlc
                  contact
            Topographic  contours In
               meters above MSL
Figure 4.2. Geology of the ELD A test site.
                                          124

-------
              UNIT1
                          Weathered till topsoil, 0 to 0.3 m thick.
                          Light brown to grey, massive clay and silt
                          containing 5 to 15 percent disseminated
                          gravel, cobbles and rock fragments.
                          Roots common.' One half to 2 m thick.
               Sharp, planar contact.
              UNIT 2
                          Light brownish-grey, massive silt. One m thick,
                          thins south of site.
                          Upwardly-grading beds and lenses of gravel, sand
                          silt and clay that range from 0.1 to 0.4 m thick;
                          and massive to laminated beds of grey silty-clay
                          0.2 to 1.0 m thick. Irregular, contorted bed-
                          forms common. Total thickness 4 to 7 m.
               Irregular contact.
              UNIT 3
                          Light grey, massive, silty-clay containing 5 to 20
                          percent disseminated gravel, cobbles and rock
                          fragments.  Irregular lenses of silt, or yellow-brown
                          sand present locally.  Lower contact not exposed, at
                          least 8 m thick.
Figure 43. Stratigraphic section from the vicinity of the ELDA test site.
                                   125

-------
       The upper surface of a light brown, massive silt bed several dm in thickness marks
the contact between Units 1 and 2. The contact is planar and nearly flat-lying, differing in
elevation by roughly 0.1 m over the study area.

       The lower till unit, Unit 3, consists of massive, grey silty-clay containing 5 to 20
percent disseminated rock fragments. Pods and beds of light grey silt, or brown sandy-
gravel occur locally. The lower contact is not exposed, so the thickness of Unit 3 is
unknown. The upper surface of the massive grey clay of Unit 3 marks the contact between
Units 2 and 3. The contact varies in elevation by several meters in the vicinity of the site.

       Hydraulic fractures were created in Units 2 and 3. Fractures at boreholes 2 and  13
were in Unit 2, the one at borehole 12 propagated upward from Unit 3 into Unit 2, and the
others were contained within Unit 3.
Hydrology

       Eleven of the 14 boreholes drilled during exploration of the site were dry when we
created the hydraulic fractures. In those boreholes that did contain water, the water depths
were less than 20 cm at the time of fracturing. Some other boreholes contained water
several days prior to fracturing, but the day before fracturing we found soft mud, but no
standing water, at the bottom of those holes (Table 4.1).

       A regional water table is present several tens of meters or more below the level of
the boreholes, but it was insignificant to the test.

       During testing the silty clay that comprises most of the till was unsaturated (ratio of
water volume to pore volume of 29%). Water that was observed in boreholes apparently
drained from perched zones within sand and gravel lenses in the till.

       Effects of the hydrologic conditions in a borehole on the results of hydraulic
fracturing were undetectable.

In situ State of Stress

       The state of stress in the till was measured in situ using a method of analyzing
injection pressure after creating a small-volume (10-20 ml) hydraulic fracture.  The
method of analysis of the pressure record is used routinely for research in rock mechanics
(Zoback and Pollard, 1978; Gronseth, 1979), but the apparatus used to create the small
hydraulic fractures in soil was designed for this project.

       The apparatus consists of a heavy-walled steel pipe (1/2" Nominal Schedule 80
pipe, O.D.: 2.13 cm; I.D: 1.42 cm) fitted with a conical point. The point consists of an outer
beveled ring attached to the pipe, and an inner, bullet-shaped point attached to a steel rod
extending the length of the pipe.  A cap fits on the end of the pipe and prevents the rod
and inner point from moving when the pipe is driven into the ground.
                                   126

-------
TABLE 4.1. CONDITIONS OF BOREHOLES PRIOR TO TESTING.

BOREHOLE      DEPTH OF             CONDITIONS
                  CASING (m) (6/9/88)     (6/13/88^    (6/14/881

     2            2.76              DRY       DRY       N/A
     3            1.39              N/A*       N/A        N/A
     4            3.65              DRY       DRY       DRY
     5            1.64              DRY       DRY       DRY
     6            1.86              DRY       MUDDY    DRY
     7            1.83              DRY       MUDDY    DRY
     8            1.89              N/A        1.45"       1.77
     9            1.86              0.88         MUDDY    MUDDY
    10            2.10              1.98         1.58         2.07
    11            1.98              DRY       DRY       DRY
    12            1.99              1.47         1.80         1.80
    13            2.01              1.18         MUDDY    MUDDY
    14            1.85              DRY       DRY       DRY
    15            2.23              1.24         1.71         1.77

 *: N/A, data not available.
 **: Water level in m below ground surface.
      To measure lateral stress the pipe is pushed into the ground using a post-hole
driver (Fig. 4.4); we were able to drive the pipe to depths of 2 m in both till and colluvium.
The rod and inner point is retracted exposing soil at the bottom of die pipe. A coring
point, shaped like a thin-walled rube 0.63 cm in diameter, is driven into the soil at the
bottom of the pipe and cuts a cylindrical hole roughly 10 on in length (Fig. 4.4).  Water is
injected into the pipe using a positive displacement pump and the pressure is monitored
using a pressure transducer and a data acquisition computer.  Lateral pressure of the soil
against the pipe seals the system and a hydraulic fracture is created from the cylindrical
hole at the bottom of the pipe (Fig. 4.4).

      The least compressive in situ lateral stress is obtained by assuming that it is equal
to the pressure of the injection fluid when a hydraulic fracture closes after pumping has
ceased. This pressure is commonly termed the instantaneous shut-in pressure, ?«.  When
the pump is turned off after a hydraulic fracture has been created, the pressure of the
injection fluid commonly shows two well-defined periods of behavior (Medlin and Masse,
1984). During the first period, the pressure decreases rapidly as fluid flows out of both the
borehole and the fracture and into the enveloping material. The rate of decrease of
pressure diminishes, however, as the fracture closes. During the second period, the
pressure decreases slowly, at a nearly constant rate, as the fluid leaks only out of the walls
of the borehole.

      There are a variety of methods used to infer P*; from the record of injection
pressure as a function of time (e.g. McLennan and Roegiers, 1982; Aamodt and
Kuriyagawa, 1982; Gronseth, 1979; Zoback and Pollard, 1978; Nolte, 1979; Muskat, 1937).
                                  127

-------
                                                      s^
                                          ®=-
                                                          •::•' Fracture from
                                                          : '.'• small diameter
Lance tip'-'-'
Small core::.:^
  cutter ~
      Figure 4.4. Method of creating small hydraulic fracture to measure the least
                 horizontal confining stress.
                                    128

-------
A relatively simple method, suggested by Medlin and Masse (1984), is to use the maximum
pressure during the second period of pressure decline. This pressure is obtained by fitting
a straight line to the record and recording the earliest point of intersection between the
line and the record (Fig. 4.5).

      According to the experiments of Medlin and Masse (1984), ?« obtained in this
manner can over-estimate the actual applied stress, and the amount of over-estimation
depends on the rate of flow out of the fracture and other factors. Roughly one dozen
experiments were conducted using the laboratory apparatus described earlier to test
whether this method of obtaining Pig would yield the confining stress in soil. Hydraulic
fractures were created from a cylindrical hole 0.63 cm in diameter and 10 cm long  (the
same size as in the field apparatus) cut in Center Hill clay. The results are encouraging;
Ptsi obtained from the method described above was typically within five percent of the
confining stress applied to the sample (Fig. 4.5a).

      In the field at the ELDA site, tests were conducted at a depth of 1.2 m in silty-clay
till (Unit 3). The injection pressure rapidly increased to 1200 kPa and then abruptly
decreased indicating the onset of fracturing (Fig. 4.5b).  We turned off the pump after 60
seconds, at which time the pressure decreased rapidly for roughly 10 seconds and then the
rate of decrease started to diminish. The rate of decrease became roughly linear at 65
seconds. A second test was conducted by turning on the pump until the pressure reached a
peak and started to diminish (indicating further propagation), and then turning off the
pump. The pressure record of the second test is similar to that of the first (Fig. 4.5b).
Fluctuations in pressure during the linear part of the second test probably resulted from
vibrations caused by heavy earth-moving equipment operating nearby.

      An instantaneous shut-in pressure of 320 kPa was obtained during the test at the
ELDA site (Fig. 4.5b). The transducer was 2.15 m above the fracture, so a correction for
the static pressure gradient was added to obtain an estimate of the lateral confining
pressure. The vertical confining stress is assumed to be the unit weight of the till times  the
depth of the test.

      Using those methods, the lateral confining stress is inferred to be 340 kPa (roughly
50 psi),  and the vertical confining stress roughly 26 kPa (3.9 psi). The ratio of vertical to
lateral confining stress is 1:13 at a depth of 1.2 m.


Hydraulic Conductivity

      Saturated hydraulic conductivity, K«, of till was measured in situ at several locations
using a Guelph borehole permeameter, a device manufactured by Soilmoisture Equipment
Company, Santa Barbara, CA. The device is designed to measure Kg  in boreholes
penetrating unsaturated porous media. In general, the device uses a Mariotte bottle device
to hold a constant level of water in a borehole, and allows flow rate out of the well to be
measured  as a function of time. A test is conducted until the flow rate becomes constant.
Details  of the operation of the Guelph Permeameter are available from the manufacturer.

      The saturated hydraulic conductivity is obtained  from the steady-state solution of
flow from  a cylindrical hole in an unsaturatedporous media (Reynolds and Elrick, 1985).
Two methods of solving for K« are available. The most accurate method requires
measurements of flow rate at two different heads (water levels in the borehole). This
                                     129

-------
      o
     Q_
          25
        100
         75
      to

      D
      (0

      s  50
      i_
     Q_

         25



           0
                 Injection pressure
         .  Applied
            Load
 a
CL
 
       1250




       1000




        750




        500




        250
                       Pump on
off
                   100     200     300     400     500
                              Depth:  1.2 m

                              ELDA Landfill
      Hmin
              Pump on |   off     | on   j     off
            0      50     100    150    200    250

                           Time (sec)
Figure 4.5a. Injection pressure as a function of time during a lab experiment.  4.6b.
            Injection pressure as a function of time at the ELDA test site using
            field apparatus designed to estimate lateral confining stress.
                             130

-------
 method requires a homogeneous porous media, and will yield erroneous values if layers or
 discontinuities are present.

       Another method of determining Kg requires measurements of flow rate at only one
 value of head, and it can be applied in layer media. The accuracy of this method is less
 than of the method described above because this one requires an estimation of the ratio
 of KS and the matric flux potential. Fortunately, however, the ratio can be estimated from
 descriptions of the till to yield values K« that are within 20 percent of Kg determined by
 other methods, according to Elrick and others (1988). This accuracy is considered
 sufficient for our purposes.

       Both methods of calculating Ks were used, but we found that the method requiring
 two measurements typically yielded unsatisfactory values of Kg (values were less than 0).
 As a result, the method requiring one measurement and an estimation of the ratio of Kg
 and matric flux potential was used for all estimates of Kg.
       In general, the saturated hydraulic conductivity of silty clay is between 1.5 x
cm/sec and 1.9 x 10-7 cm/sec, whereas those measured in silty sands and gravels are an
order of magnitude or more greater; between 1.0 x 10-5 cm/sec and 3.5 x 10-s cm/sec.
Measurements of Kg were made at five locations, and the values are tabulated below
(Table 4.2).


Physical Characteristics of the Till

       A suite of tests was performed in the laboratory on samples of the ELD A Till to
determine standard geotechnical characteristics. The tests were conducted by Patricia
Strube, a geotechnical soils analyst at the Center Hill Research Facility.  The methods
used to make the measurements are described in the ASTM Handbook on Soil Testing.
Samples from four shelby tubes pushed into the till at the ground surface were used for the
analyses.
TABLE 4.2. SATURATED HYDRAULIC CONDUCTIVITIES OF TILL
Location
1
2
3
4
5
Depth (cm)
40
45
150
30
42
Material
silty clay
silty clay
silty clay
silty sand
and gravel
silty sand
and gravel
K» (cm/sec)
7.5 x 10-7 to 1.5 x 106
1.0x10* to 1.4x10*
1.9 x 10-7 to 8.9 xlO-7
1.0xlO-sto2.4xl05
3.0 xlO-5 to 3.5X105
                                   131

-------
TABLE: 4.3 PHYSICAL CHARACTERISTICS OF SILTY-CLAY TILL
Atterburg
Limits
    Liquid         22.6         22.9         24.8
    Plastic         12.5         10.5         11.4
    Shrinkage      19.8         15.5         15.3
    P.I.            10.1         12.4         13.4

Grain sizes (by wt.)
    Gravel           0.13         0.18
    Sand             0.29         0.28
    Silt              0.36         0.31
    Clay             0.22         0.23

Void Ratio:          0.34
Porositv:             0.25
Moisture:          11.6         13.5
Degree of
Saturation:           0.29

Bulk Unit Weight:    2268kg/m3 (141.6 lb/ft3)
Bulk S.G.:                        22
Drv Unit Weight:     2031kg/m3 (126.8 lb/ft3)
             0.10
             0.34
             0.34
             0.22
                                                         ET4
                        25.0
                        15.7

                         9.3
0.11
033
0.37
0.19
                                    AVE
            23.8
            12.5
            16.8
            11.3
Specific gravity, solids:
2.68
      The characteristics of the ELD A Till are similar to those of other tills analyzed by
Strube in the vicinity of Cincinnati. It is a hard soil, requiring more than 30 blows during a
standard penetration test (Navfac, 1982), and it is relatively dense (bulk unit weight of
2268kg/m3 or 141.6 lb/ft3) compared with colluviai soil or alluvium.

      The till is well-graded from particles of clay size to sand size or larger. Cobble- to
boulder-sized rock fragments are common in field exposures, although they were absent
from the sample used for the grain size analysis (Table 43; Fig. 4.6).

Results from tests of the Atterburg Limits plot  above the A line and indicate that the till is
a CL type soil, in the nomenclature of the Unified Soil Classification System (ASTM
D248,
BOREHOLES

      Eleven boreholes were hydraulically fractured during the field test. Most borings
are approximately 10 m from their neighbors (Fig. 4.1), except at the southern end of the
site where a cluster of four borings are each 2.5 m from their nearest neighbor (Fig. 4.1).
                                     132

-------
u.i. HA*
tOO. 4-1




n


i "
* >•

r





0








































1



















DAB WVI ontwo M MOO U.I
m i v H « i « « in i
"



















t



















— -^^
1


















MO 1M M 10

COMUI
WMM



•"" —



sssa
r
.
^




















^
















nANDAASKVf NUMHH
U*» M 40 J070JCM
1 1 1 >


.

















—
*» 	
1 1














T

---
•s














1





X,


















i^













ri«o 71
I I






X.
^ ,











MAM S« HUWCTItt 
  • Figure 4.6. Grain size distributions in samples of ELDA Till. 133

  • -------
        U i HAMUV MVt OttMMO M NCMU    U.L llAMUJO WV1 NUXMB
    
    
    
    
    
    
    
    u
    
    
    
    
    
    
    
    
    
    
    
    
    
    Ml
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    4 1 1
    1
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    100
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    10
    H
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    1
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    V K H
    r^-<
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    10
    1 i
    
    •
    ^
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    *
    1
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    1 1
    1
    
    
    V.
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    0 1* M »
    1 I 1
    
    
    
    "**-^
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    • JC
    1
    
    
    
    
    * " N
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    0.1
    J
    
    
    
    
    
    
    ^
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    0 7(
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    *DO*I40 71
    1 '
    
    
    
    
    
    
    X
    ^
    
    
    
    
    
    
    
    
    
    
    
    O.I
    0
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    OOJ
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    J
    \
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    ^
    \
    
    
    
    
    
    
    
    
    0.01
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    1
    s
    
    
    
    
    
    4.00J
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    '
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    •^
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    S*^-
    
    
    
    DC
        u i II««UD im OKMNO M torn     u.i. SIUOAJO tun NUMHB
    i 1 1
    
    
    
    
    
    . _
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    —
    
    
    
    
    1 IK 1 » » ft
    
    ,)B
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    -
    
    
    
    
    1
    to-.
    
    
    
    
    
    
    
    
    
    
    
    -
    
    
    
    
    '
    
    ^
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    1 4 * IWUUK
    r
    
    
    " ,
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    N
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    1 '
    
    
    
    
    •x^
    '
    
    
    
    
    
    
    
    
    
    
    
    
    
    >OJC XI 70 100' I«O Ml
    
    
    
    
    
    
    
    *^
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    ^
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    ""^x^
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    '"-1
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    J
    (
    
    
    -
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    y
    s
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    ^v
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    .
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    —    H
                    «AD«10N CUIVU
                                           134
    

    -------
           Depths to the bottom of casing ranged between 1.64 m and 1.95 m at nine of 11
    boreholes.  The other two boreholes were deeper; 2.72 m and 3.81 m (Table 4.1).
    
    
    Design
    
           Boreholes were designed for the purpose of creating hydraulic fractures. In
    general, a borehole consists of a steel tube cemented into a boring, and open at both ends
    (Fig. 4.7). A basket is fixed to the lower end of the casing to prevent cement from plugging
    the bottom of the borehole. The open boring extends several dm below the basket ana is
    partly filled with fragments (cuttings) of till. A narrow notch, oriented normal to the axis
    of the borehole, is cut in the wall of the boring several cm below the bottom of the casing.
    
           Roughly horizontal hydraulic fractures were expected to develop, and the boreholes
    were designed to nucleate a horizontal fracture at the notch. Although the expectations of
    horizontal hydraulic fractures were correct, most of the fractures initiated at or above the
    basket, not at the notch.
    
    
    Specifications
    
           Borings were made with a Marmon-Herrington drill rig using an auger 7.62 cm (3
    in.) in diameter. The auger was pulled out of the hole after each 0.3 m (Ifoot) of drilling
    so that samples of the cuttings could be collected. Each boring required between 30 and
    90 minutes to complete.
    
           The borings were drilled several dm beyond the depth at which we intended to cut
    the notch because cuttings filled a zone several dm deep at the bottom of the hole. Tools
    to retrieve cuttings from the bottom of the hole were unavailable during drilling.
    
           Notches were cut in the walls of the boring using a mechanical device.  The device
    consisted of a talon-shaped blade (made from tool steel 0.3 cm thick) mounted on the
    bottom of a PVC tube which fit snugly into the boring. The blade was mounted on a rod
    attached to the inner wall and extending the full length of the tube.
    
           To cut a notch, the blade was brought flush with the tube and the apparatus
    inserted into the boring until it rested  on the top of cuttings at the bottom of the boring.
    The blade was then slightly extended by twisting the rod at the ground surface. The entire
    apparatus was rotated, cutting a shallow notch in the wall of the boring. Then the blade
    was extended further and the notch was deepened by continuing to rotate the apparatus.
    
           When the operation was complete, the wall of the boring contained a notch 3.75 cm
    deep and 0.3 cm thick (Fig. 4.7). Notches of those dimensions were observed when the
    boreholes were excavated.
    
           The boreholes were cased with steel tubing 3.75 cm in outer diameter and 3 mm in
    wall thickness (1 1/4" Nominal Schedule 80 steel pipe). The tubing was obtained at a local
    plumbing supply store.
    
           Baskets were constructed of annular disks of 3/4" (nominal) plywood, plastic foam
    and fine steel mesh. Three foam disks were placed on the end of the casing, and a steel
                                        135
    

    -------
                Bentonite
    
    
    
                  Gaskets
            Support ring
              '    '  ^
              0   5  10
                 cm
                                        Portland  cement
    JN;   Ijqi—Basket
    
               «-Notch
                                             Boring partly filled
                                             with  cuttings
    Figure 4.7. A borehole used to create hydraulic fractures.
                                 136
    

    -------
    mesh disk placed beneath them. The outside diameter of those disks was 8 cm, slightly
    larger than the inside diameter of the boring. The plywood disk served as a support for
    the other disks,  steel band was secured beneath the basket to prevent it from slipping off
    the end of the casing (Fig. 4.7).
    
          When the casing and basket assembly was inserted into a boring, the plastic and
    steel disks were flush with the boring walls.  Bentonite pellets were placed on top of the
    basket to seal the annulus between the casing and the wall of the boring.  Portland cement,
    which contained an additive that inhibits shrinking, was poured in the annulus to secure
    the casing into the borehole. The result was a cemented casing whose lower end was open
    to the till.
    METHOD OF FRACTURING
    
           Hydraulic fractures were created by workers from Halliburton Services, a company
    specializing in creating hydraulic fractures for the petoleum industry. Most of the methods
    used during the test are similar to methods used in the hydraulic fracturing of oil wells.
    We modified the methods slightly, however, because of the shallow depths of the
    boreholes.
    Equipment
    
          The fracturing equipment was mounted on five vehicles: a truck containing a
    blender and a positive displacement pump, two trucks containing sand, a truck containing
    water, and a van containing monitoring and control equipment.
    
          Three pumps were used in the operation: a Triplex positive displacement pump,
    and two centrifugal pumps, one upstream and one downstream from the blender. The
    positive displacement pump can generate a maximum of 41,000 kPa (6000 psi), whereas
    the centrifugal pumps can produce a maximum of between 400 and 550 kPa (60 and 80
    psi) and flow  rates of 2.4 to 2.8 m3/min (630 to 750 gals/min).
    
    
    The Injection Fluid
    
          Sand and several chemicals were added to water in the blender to improve the
    performance of the fracturing operation. To  increase the viscosity of the fluid, we added  a
    guar gum base gel at a concentration of 2.5kg/m3 (0.025 Ibs/gal), and a cross-linking
    chemical at a concentration of 0.05kg/m3 (0.0005 Ibs/gal). A buffer was used to reduce
    the pH of the fracturing fluid to between four and five, which is optimal for use of the gel.
    
          The guar gum gel increases the viscosity of the water from 1 to 15 cP. The cross-
    linking chemical causes the gel to become thixotropic, with an apparent viscosity of 100 to
    200 cP and an apparent shear strength of roughly 25 Pa, according to laboratory tests
    conducted with a rotational viscometer. The  high viscosity and shear strength of the gel
    facilitates the transport of sand into the fracture. Typically, another chemical is blended
    into the fluid to cause the gel to break down several hours after injection. The breaking
    chemical improves removal of gel, greatly increasing the permeability of the propped
    fracture.
                                          137
    

    -------
          Potassium chloride was added to the fluid at a concentration of three percent by
    weight.  The KC1 is an additive that is used routinely to inhibit the swelling or clay
    minerals in hydraulic fractures created by the oil industry. In addition, this material
    increases the electrical conductivity of the fluid and was used to improve the resolution of
    fractures by electrical geophysical techniques.
    
          Dye was added to the fluid to stain the fracture surfaces and make them easier to
    identify during excavation. Three types of dye were tested during the experiment:
    rhodamine, fluorescein, and fluorescent orange paint pigment. Dye was mixed at a
    concentration of 1:500 in the first few fractures but it was more dilute (1:1000 to 1:2000) in
    the last few fractures.
    
          Ottawa sand was mixed with the fracturing fluid to prop open the fractures. Two
    different sizes of sand, very coarse-grained, 12/20 mesh, and medium-grained, 20/40
    mesh, were used. Concentrations of sand ranged from 9% to 18% by volume (2 to 4 Ibs.
    sand/gal, fluid).
    
    
    The Fracturing Procedure
    
          During routine fracturing operations of oil wells, water is pumped from the water
    truck and mixed with sand and chemicals in the blender. A centrifugal pump pulls the
    mixture of water, sand and chemicals from the blender and creates input head for the
    positive displacement pump. Injection into the borehole is accomplished using the
    positive displacement pump. Flow rates and injection pressures are monitored in the van.
    
          During our tests, the procedure differed slightly from above in that we used the
    centrifugal pump at the downstream end of the blender to inject fluid into the borehole.
    The positive displacement pump was used intermittently, when pressures in excess of 550
    kPa (80 psi) were required to initiate fracturing.
    
          The fracturing procedure began by filling the casing with water, so the initial
    fracture would be created with water rather than gel.' Pipes were connected from the
    borehole to the trucks containing the fracturing equipment, which were parked on the
    northeastern edge of the site (Fig. 4.1).
    
          Injection using the centrifugal pump caused pressures measured at the borehole to
    increase to between 400 and 550 kPa (60 and 80 psi) within several minutes.  In some
    cases, a fracture was created soon after the pressure reached 60 to  80 psi, but in other
    cases 5 to 10 minutes were required before fracturing. Injection pressure was increased
    using the positive displacement pump if a fracture was not created  after approximately 10
    minutes, injection pressures of as great as 820 kPa (120 psi) were required to initiate
    fracturing at some boreholes (Table 4.4).
    
          The onset of fracturing was determined by an abrupt decrease in the injection
    pressure. Injection pressures decreased to between 69 and 275 kPa (10 and 40 psi) during
    the propagation of most of the fractures (Table 4.4). Records of the pressures and
    volumes as functions of time are presented in the Appendix.
    
          Injection was terminated at the shallow boreholes when fluid vented to the ground
    surface, and at the deeper boreholes injection was terminated after a predetermined
                                          138
    

    -------
     volume of injection. Each fracture was created in less then ten minutes. The pipes
     extending from the pump truck to the borehole was uncoupled at the borehole
     immediately after termination of injection. Roughly 100 liters (several tens of gallons) of
     injection fluid flowed back out of the fracture into the borehole and onto the ground
     following uncoupling.
    
           Flow rates of between 0.075 and 0.42 m3/min (20 and 90 gal/min) were used to
     create the fractures. In general, the flow rate was low during peak pressure and increased
     as the pressure decreased and the fracture began to grow.  Details of the history of flow
     rates, however, are poorly known because the flow meter used by Halliburton Services was
     often clogged with sand and non-functional. Clogging apparently occurred because the
     flow rates that were used were less than those designed to be measured by the meter.
    
           The total volumes of fluid and sand pumped  out of the blender were estimated by
     the operator of the pump truck.  Typically the fractures required 0.38 to 0.76 m3 (100 to
     200 gals) of fluid, but the volume ranged from as little as 0.076 m3 to as much as 1.51 m3.
    
           The volume of sand pumped to the fractures was typically 0.056 to 0.14 m3 (2 to 5
     ft3), but it ranged from less than 0.03 to as much as 0.85 m3 (30 ft3). These estimates are
     greater than the volume contained in fractures because they do not account for material
     that remained in the pipes after fracturing, or that flowed out during venting.
    TABLE: 4.4. SUMMARY OF DATA FROM FIELD TESTS
    Id
    
    FLUID
    Volume Dye
    m3 (gal)
    13
    11
    12
    2
    10
    4
    9
    5
    8
    7
    6
    1.51 (400
    0.11
    0.45
    0.46
    0.57
    0.57
    0.76
    0.76
    0
    0.38
    (30
    120
    110
    150
    150
    200
    (20
    (0
    (10
    0.57 (150
    rh
    rh
    rh
    rh
    rh
    fo+fl
    fo+fl
    fo
    fo
    fo
    fo+fl
    SA
    Volume
    m3
    0.28-0.42
    0.057
    0.08-0.14
    0.85
    0.22-0.24
    0.17-0.22
    0.08-0.14
    0.03-0.08
    0
    0.03-0.08
    0.14
    JSTD
    Gr.Size
    
    20/40
    20/40
    20/40
    12/20
    12/20
    12/20
    20/40
    20/40
    20/40
    20/40
    20/40
    PRESSURE
    Max. Propagation
    kPa(BSi) kPa
    620 90
    340 50
    410 60
    410 60
    410 60
    760 ( 10
    820 (120
    410 (60
    3700 (540
    550 (80
    550 (80
    140-340
    140
    140
    140-280
    340
    480
    200-275
    410
    ..
    550
    200
                       rh: rhodamine red
                       fo: fluorescent orange pigment
                       fl: fluorescein
    
    Shown in the order in which the fractures were created.
                                         139
    

    -------
           The amount of sand observed in some of the fractures was much less than the
     amount pumped from the blender.  For example, 0.122 m3 (4.2 ft3) of sand filled the
     fracture at borehole 13 (details of this calculation are described in Section 6), but 0.28 to
     0.42 m3 (10 to 15 ft3) were pumped from the blender. Approximately 0.1 m3  of the
     missing sand is inferred to nave been deposited in pipes extending from the pump truck to
     the boring, and roughly 0.05 m3 flowed out of the fracture during venting.
    
           Hydraulic fracturing was attempted at eleven boreholes, and fractures were
     successfully created at ten of those boreholes. Hydraulic fracturing did not occur at
     borehole 8.  Two attempts were made to fracture borehole 8; during the first injection
     pressures reached 1720 kPa (250 psi), and during the second injection pressures  reached a
     maximum of 3700 kPa (540 psi). Following both attempts, we discovered plugs of sand
     several dm from the bottom of the casing that presumably blocked the injection fluid. The
     cause of the sand blockage is unclear. One explanation is that the gel in the injection fluid
     broke down and was unable to adequately transport sand. The flow of the gel is sensitive
     to heat and could have been affected by the ambient temperature (which was 38 ° C on
     the day of the test), according to Mark Roberts, field engineer with Halliburton Services.
    
    
     Shortcomings of the Equipment
    
           Shortcomings of the hydraulic fracturing equipment used during the test are all
     related to scale: the equipment was designed to create fractures one to several orders of
     magnitude larger than the ones we created. As a result sensing devices such as flow
     meters and pressure transducers used by Halliburton were unable to accurately measure
     the low flow rates and pressures of the test.
    
           The blending system was unable to consistently deliver a quality mixture of
     injection fluid at the relatively low volumes required. This resulted in variations in fluid
     properties between fractures, according to observations of the fluid flowing back out of
     fractures. In some cases, especially during the last several fractures, the viscosity of the
     injection fluid was low and this could have been partly responsible for sand settling out
     before reaching the subsurface.
    
           The fracturing equipment was practically immobile, so pipes as long as 50 m were
     used to connect the injection pump to the boreholes. The volume of the long pipes was on
     the same order as the volume of the fractures, which reduced the control of the fracturing
     process. Moreover, in many cases it appeared that sand settled in the pipes before
     reaching the fracture.
    
           The team from Halliburton Services made every effort to adapt their equipment to
     the highly atypical conditions of our tests.  The shortcomings cited above in no way reflect
     the manner in which the equipment was operated. The shortcomings do indicate,
     however, that conventional equipment used to create hydraulic fractures at oil wells will
     be too large to create hydraulic fractures of the size used during this test. Future
     applications would benefit from equipment designed for the pressures, flow rates, and
    volumes required to create and prop open hydraulic fractures at shallow depths (tens of
     meters).
                                      140
    

    -------
                                   SECTION FIVE
    
        HYDRAULIC FRACTURES CREATED DURING THE FIELD TEST-1988
    
           These two observations, 1) an abrupt decrease in the pressure of the
     injection fluid,  and 2) venting of the fluid several m or more from the point of
     injection, were critical because they are strong evidence that hydraulic fracturing of
     soil occurred during the field test. There was little information at the time of the
     test, however, about fracture geometry, that is the size, shape, orientation, location,
     and thickness of the fractures. Those parameters largely control how the fracture
     will affect the recovery of ground water, and so they are of fundamental importance
     to the goals of this research.
    
           The geometries of the fractures were determined by excavation and mapping.
     The day after the fractures were created, we mapped the locations of the vents and
     prepared to excavate the vicinity of each borehole.  Two days later a backhoe was
     rented and brought to the site to begin excavation. The strategy used during
     excavation was to first cut a trench from the vent to the parentborehole. The
     fracture was identified at the vent and traced along the trench during excavation.
     The trench was extended beyond the borehole until the fracture could no longer be
     identified in the trench walls. Then another trench was cut perpendicular to the first
     one.  Subsequent trenches were placed to intercept the fracture at critical locations
     based on existing exposures.
    
           The locations of trenches were selected to give the maximum amount of
     information on the forms of the fractures, while keeping the amount of excavation
     to a minimum.  In general, the technique was successful, but in a few cases
     (boreholes 2,4, and 9) the fractures could not be identified during excavation, so a
     full array of trenches was not completed. Careful investigation following excavation
     did reveal the locations of fractures at borehole 9, but the backhoe was no longer
     available to complete excavation.
    
           Two boreholes that were  incompletely excavated (2 and 4) were several
     meters deeper than the others and they lacked a vent from which to start excavation.
     We discovered a fracture adjacent to borehole 2, but the fracture was poorly
     developed so excavation was discontinued. At borehole 4, the maximum depth that
     could be reached by the backhoe was 0.3 m above the  bottom of the casing.
     Monitoring techniques indicated that a fracture developed at borehole 4, but after
     considerable effort we were unable to locate the fracture.
    
           Traces of fractures on the walls of the trenches were identified and marked
     with brightly-colored flagging.  Many of the fractures lacked sand and could be
     identified only after detailed excavation revealed traces of dye staining.
    
           The appearance of the fractures on the walls of the trenches depended on
     the material filling them. Where fractures were filled  with sand, they appeared as
     thin white layers in the grey till.  Gel mixed with the fracturing fluid allowed some
     hydraulic fractures to be identified because the gel appeared as a mucous-like or
     rubbery film on the fracture surfaces.  Many of the fractures lacked sand or gel and
     could be identified only by trace  amounts of dye on the fracture surfaces.
     Rhodamine was the most effective dye, staining fracture surfaces deep purple (a
    color that was readily distinguished from natural iron oxide stain, which is dark red).
                                      141
    

    -------
    Fluorescent paint pigment yielded satisfactory results, although the distribution of
    the pigment was less uniform than the distribution of the rhodamine stain. The
    paint pigment fluoresced and could be detected in the field using a UV lamp.
    However, the UV lamp was not used routinely to detect the dye because the
    florescence could only be detected when the trenches were covered with a tarp to
    reduce ambient lighting. Both of those dyes effectively stained fracture surfaces
    when mixed with injection fluid at concentrations of roughly 1:500, and they could
    be detected on fracture surfaces at concentrations of 1:1000 to 1:2000. Fluorescein
    dye was ineffective at staining the surfaces of fractures, even though it resulted in a
    strongly-colored injection fluid.
    
    
    FORMS OF THE HYDRAULIC FRACTURES
    
          Hydraulic fractures were created at shallow depths so that they could be
    excavated and mapped in detail. The purpose of the excavation and mapping was to
    determine physical characteristics of the fractures, such as their size, shape,
    orientation, and thickness. This information is necessary to the design or hydraulic
    fractures in remedial actions, to infer processes of fracturing, and to highlight
    features of the fractures that are undesirable.  Identifying the undesirable features
    will allow us to develop techniques that will inhibit their formation.
    
          The geometry of each hydraulic fracture, that is its size, shape and
    orientation, was determined by mapping exposures of the traces of the fractures on
    the walls of trenches. Sections obtained! from the trench walls were compiled into
    maps showing the surface of the fracture in plan.
    
          The sections were made by measuring the location of the fracture traces
    relative to a horizontal datum, which was constructed by positioning fine string with
    a hand level and securing it to the trench wall. The location of a fracture trace
    could be mapped to within a few cm using this technique. Sections were mapped at
    a scale of 1:12 and reduced to the sizes shown in the following pages. The locations
    of most of the fracture traces are accurate to within one line width on the following
    sections.
    
          The fractures shown in the following maps and sections represent exposures
    that could be confirmed because they contained sand, dye, or gel.  The
    concentration of those materials decreased as the leading edge of a fracture was
    approached, and we determined the leading edge as the last exposure of injected
    material. Commonly, however, an unfilled fracture extended several dm or more
    ahead of the point that was mapped as the leading edge.  The unfilled fractures
    probably were caused by hydraulic fracturing, but they were omitted from the
    sections because their origin was uncertain.
    
          The forms of each of the hydraulic fractures will be described  in the
    following pages. An idealized form has been inferred from the descriptions and will
    be presented at the end of this section.
                                       142
    

    -------
    Borehole Four
    
           A hydraulic fracture could not be identified in two trenches cut adjacent to
    borehole Four. Monitoring of injection pressure and surface tilts indicates that a
    hydraulic fracture was formed, but we were unable to excavate deep enough to
    reach it.
    Borehole Five
    
           A hydraulic fracture occurs on the northern side of borehole five (the
    fracture will be referred to as HFS). In general, the fracture is continuous in the
    vicinity of the borehole, and it splits into two lobes several m north of the borehole
    (Fig. 5.1). One lobe vented on the northwestern side, and the other remained in the
    subsurface on the northeastern side of the borehole (Fig. 5.1).
    
           The northwestern lobe is shaped like a trough that plunges 20° to 30° toward
    the borehole (Figs. 5.1 and 5.2). Each side of the trough is roughly planar and the
    sides differ in stnke by 135°. The sides of the trough split near the ground surface,
    and they vent as separate fractures striking roughly tangential to the major axis of
    the fracture.  As the hydraulic fracture approaches the ground surface it flattens for
    several dm, and then curves abruptly upward one dm below the ground surface. The
    fracture is nearly vertical where it intersects the ground surface at the point of
    venting (Fig. 5.2).
    
           The northeastern lobe of HFS splits from the northwestern lobe roughly 1.5
    m north of the borehole. The dip of the northwestern lobe is slightly greater than
    that of the northeastern lobe, so that the northwestern lobe climbs above its
    neighbor with increasing distance from the borehole.  In plan view, the two lobes are
    separated by unfractured ground; no overlap was observed (Fig. 5.1).
    
           A vertical fracture striking N62E cuts the host till adjacent to the open
    interval and the bentonite-filled zone of the borehole. At 15 cm above the bottom
    of the casing (adjacent to the bentonite), the dip of the fracture flattens abruptly
    over severalcm. Within 20 cm of the bottom or the casing, the fracture dips 25° to
    the south, and this dip is typical of most of the fracture.  The vertical  fracture is
    apparently unrelated to the notch at borehole five.
    
    
    Borehole Six
    
           Hydraulic fracture six resembles HF5 in that it consists of a steeply dipping
    fracture adjacent to the borehole, a zone that is flat-lying in the vicinity of the
    borehole, and several shallowly dipping lobes (Fig. 5.1). The vertical fracture is
    relatively large compared with HF5; it is at least one m in length and it extends
    upward from the notch to a height of 1.25 m (Figs. 5.1,5.2 and 5.3). The strike of
    the vertical fracture is N23W, nearly perpendicular to that of the vertical fracture at
    HFS.
    
           The hydraulic fracture is nearly flat-lying within approximately two m of the
    borehole, a distance that is similar to the depth of initiation of the fracture. At a
                                       143
    

    -------
         	Elevation of frx surface
                   (m ibov» bonom of eulng *s)
    
         —	Wan of trench
    Figure 5.1. Hydraulic fractures HF5, HF6, and HF7. Structural contours are on the
                  fracture surfaces.
                                         144
    

    -------
    distance of one to two meters from the borehole, the fracture changes orientation
    abruptly, curving upward to dip shallowly toward the borehole (Fig. 52).
    
           A large fracture lobe extends south of the borehole to a vent roughly 8 m to
    the WSW. This lobe is roughly planar in form and dips toward the borehole at 14°
    (Figs 5.2 and 5.3).  In plan, the large lobe is elongate with an aspect ratio of roughly
    2:1.  The borehole is near the end of the long axis of the lobe (Fig. 5.1).
    
           The fracture trace along the major axis  of the large lobe is, in general,
    straight from where the fracture begins to turn upward to the vent. In detail,
    however, the fracture is gently-stepped, consisting of riser intervals that are 20 to 40°
    steeper than the average dip of the fracture. The riser intervals are a dm or less in
    length and they are spaced roughly one m apart along the fracture. Till cut by the
    fracture was studied carefully for variations in  texture or composition that could be
    related to the location of risers, but no such variations were found. The step-like
    form of the trace appears to be unrelated to variations in till detectable in hand
    sample.
    
           Two smaller lobes were identified; one  of them is at the same elevation as
    the large lobe and extends to the SSE, whereas the other is 1.5 m above the large
    lobe and extends to the north where  it vents roughly three m from the borehole (Fig.
    
    
           The upper, northern lobe apparently is  connected to the vertical fracture in
    the vicinity of borehole. There are several isolated fracture segments a few dm in
    length exposed between the lobe and the vertical (Fig. 5.2), but apparently a direct
    connection of the upper lobe and the vertical fracture was removed during
    excavation.
    
           The vent at the large lobe resembles the vents at HF5 in that the fracture
    trace flattens (Fig. 5.3) and then turns abruptly upward when it is within one dm of
    the ground surface (Figs. 5.2 and 5.3). The strike of the fracture at the vent is sub-
    parallel to the strike of the vertical fracture at  the borehole.
    
           The vent at the smaller, northern lobe of HF6 resembles the other vents, but
    the fracture flattens at an unusually shallow depth of 3 to 8 cm. The fracture cuts
    beneath the ground surface at this shallow depth for 0.5 m before turning upward
    and venting.
    
    
    Borehole Seven
    
           Hydraulic fracture seven is the smallest observed during excavation. The
    fracture is steeply dipping adjacent to the borehole and it flattens gradually with
    increasing distance from the borehole (Fig. 5.4).  The dip of the fracture decreases
    to 25° toward the borehole roughly at 0.5 m from the vent. The average dip of HF7
    was greater than the other fractures because the transition from vertical to gently-
    dippmgfracture was relatively gradual. This behavior is important because it shows
    that HF7 intersected the ground surface after growing a small distance from the
    parent borehole, compared with the lengths other fractures (Figs. 5.1 and 5.4).
                                        145
    

    -------
    Figure 5.2. Trace of hydraulic fractures HF5 and HF6 on the northern wall of the
                 longest east-west-trending trench shown in Figure 5.1. Boreholes are
                 projected onto the section.
                                       146
    

    -------
    Eut
          Notch
                                              South wu of tr«n<*
                                                                              M«t«f»
       Figure 53. Trace of hydraulic fracture HF6 on the southern wall of the same trench
                     as in Figure 52.
                                             147
    

    -------
    Silty clay till
    
         10-20%  Cobbles
      Gravelly sand
              limestone fragment
                     Detail from opposite wall
    Figure 5.4. Trace of hydraulic fracture HF7 on the western wall of the trench cut
                along the major axis of the fracture. The section was made prior to
                excavating the tranverse trench shown in Figure 5.1.
                                     148
    

    -------
           In other respects, HF7 resembles the other fractures. The steeply dipping
    part of HF7 cuts the open zone below the basket and contains the axis of the boring.
    It is apparently unrelated to the notch. In plan view, the fracture is slightly elongate,
    with an aspect ratio of 1.5:1. The borehole is at the end of the long axis (rig. 5.1).
    The vent at HF7 is typical: the fracture flattens as it approaches the ground surface,
    reaches a depth of 0.2 m and cuts abruptly upward venting as a vertical fracture
    nearly 1 m in length and striking parallel to the vertical fracture at the borehole.
    
           A lens of gravelly-sand and limestone fragments is cut by HF7. The fracture
    cuts beneath one limestone fragment, upward through a narrow space between
    neighboring fragments, and then flattens, assuming roughly the path it was taking
    before reaching the fragments. Limestone fragments appeared to affect the path of
    the fracture only locally.
    
    
    Borehole Nine
    
           The form of HF9 resembles the large lobe at HF6 in that it is roughly
    elongate, and the parent borehole is near one end of the major axis (the
    northeastern boundary of the fracture was not excavated and is inferred).  The
    fracture is fiat-lying in the vicinity of the borehole, but most of the fracture dips
    shallowly (17°) to the southeast toward the borehole (Figs. 5.5 and 5.6).
           The fracture trace from borehole to vent along the major axis of HF9 is
    nearly straight, marked only by gentle steps spaced roughly 1 m from each other
    (Fig. 5.6). The trace resembles those of HF5 and HF6 (Fig. 52 and 5.3).
    
           A vertical fracture cuts upward from the notch to the basket of HF9. At the
    basket, the vertical fracture flattens abruptly, rolling over and becoming flat-lying in
    the vicinity of the borehole. The exposures of the vertical fracture are marked in
    Figure 5.6, but the size of the fracture was impossible to determine because the
    edges were removed during excavation. The vertical fracture strikes N63W, which is
    at a high angle to the strike of the fracture at the vent.
    
    
    Borehole 10
    
           The hydraulic fractures at borehole  10 consists of two lobes extending from
    the borehole in opposite directions, to the northeast and southwest. The two lobes
    meet at a vertical fracture intersecting the borehole.  The vertical fracture is roughly
    2 m along strike and 1 m in height, which is similar in size to vertical fractures at
    HF6 and HF7. The lobes dip between 20 and 25° toward the borehole and are
    roughly planar (Fig. 5.7).
    
           Vents occur at three locations, two on the southwestern lobe and one on the
    northeastern lobe. At the vents, the strikes of the fractures are roughly parallel to
    the vertical fracture intersecting the borehole (Fig. 5.7).
    
           The northeastern lobe cuts across a hydraulic fracture created from borehole
    11 (Fig. 5.7), which was created prior to HF10.  The line of intersection of the two
    fracture planes was shallow, so it was excavated by hand and studied carefully. We
                                         149
    

    -------
      —~ Leading edge  of fracture
      ///// Vent
       O Borehole
     	Wall of trench
    Vertical
      fracture
      — Elevation of frx surface
                       bottom fff culn0)
    Figure 5.5. Hydraulic fracture HF9. Structural contours are on fracture surfaces.
                  The northeastern edge of the hydraulic fracture was fracture was
                  covered and has been inferred.
                                            150
    

    -------
                    Bor«hol« *9
                                                                         Silty clay 1111
                                                                          10-20% cobtl.i
           Vartlcal Iractwa
    Figure 5.6. Trace of hydraulic fracture HF9 on the southwestern side of the trench
                  that cuts the major axis of the fracture.
                                              151
    

    -------
                                                          Vertical  fracture
                                                    Fracture
                                                     Intersection
                               Vertical
                               fracture
              v  meter   '
    
    	 Leading edge  of fracture
    ///// Vent      l*Bh"1 *™ awrox-)
     @  Borehole
    	Wall of trench
    	 Elevation of frx surface
                      Im obovt bottom
                       of casing #10)
    Figure 5.7. Hydraulic fractures HF10 and HF11.
                                           152
    

    -------
    could find no indication that the presence of the earlier fracture (HF11) affected the
    shape, size, distribution of sand, or any other characteristic of HF10.
    Borehole 11
    
           The hydraulic fracture at borehole 11 consists of a main lobe extending to
    the southwest, and a smaller lobe that lies above and to the west of the main one.
    The main lobe is slightly elongate in plan with an aspect ratio of 1.5:1. The
    borehole is at the end of the long axis (Fig. 5.7). The fracture surface is roughly
    planar, dipping 20° toward the borehole. The main lobe cuts the borehole adjacent
    to the basket, 0.1 m above the notch.
    
           A vertical fracture occurs at the borehole and extends to the SSW. The
    bottom of the fracture lies on the notch, and it cuts upward through the main lobe.
    The vertical fracture breaks into segments above the main lobe, but it appears to be
    connected to the smaller lobe. The connection was removed during excavation.
    
           The strike of the vertical  fracture is at a high angle to that of the vents.
    Although this relationship between strikes was observed at HF9, it is atypical. It is
    possible, however, that vertical fractures oriented parallel to vents were present at
    HF9 and HF11 because the excavations would have removed fractures of that
    orientation at those boreholes.
    Borehole 12
    
           The form of HF12 resembles a plunging trough (Fig. 5.8) formed by two
    crudely planar sides and a narrow zone that curves sharply. Both sides of the trough
    dip shallowly toward the borehole, and the strikes of the sides differ by 60°.  The
    northwestern lobe of HF5 also has a trough-like form, but the angle between the
    sides at HF5 is more shallow than it is at HF12.
    
           The excavation of HF12 was unusual because the easternmost trench cuts
    roughly parallel to  the leading edge of the fracture-other trenches cut the leading
    edges of fractures at high angles.  Exposures in that trench show six lobes  fringing
    the edge of the fracture. The orientation of the lobes could be determined in a few
    locations at which the lines of their major axes intersected the parent borehole. The
    lobes range from several dm to 1 m in width and they appear to be longer than they
    are wide.
    
           The contact between till Units 2 and 3 occurs roughly 0.5 m below  the ground
    surface at HF12. The fracture initiates in massive silty clay of Unit 3 and  cuts across
    the contact into beds of gravelly sand and silt in Unit 2 (Fig. 5.9). The geology
    exposed on the east wall of the long trench cutting HF12 was mapped in detail to
    examine the relationship between bedding and the form of the fracture.
    
           The plunge of the trace of the fracture flattens slightly over a meter-long
    interval where HF12 cuts across the upper surface of a bed of gravelly sand slightly
    east of the center of the section shown in Figure 5.9. Bedding is irregular, however,
    and the fracture cuts partly through the gravelly sand and partly through the
                                         153
    

    -------
                     — Loading «dga of frortura
                     Illll Vent    w—"d •"" •*""•'
                      O Borahol*
                    ••" ™~ Woll Oi II 6fldk
                     — Elevation of frx sirfac*
                               frn BbuM • datun O.&n
                                tt»M batlon «r Mine)
    Figure 5.8. Hydraulic fracture HF12. Lobes inferred from discontinuous exposures
                   of the fracture on the walls of the trench.
                                            154
    

    -------
                                                 Orange-brown gravelly sand
                                                 local sill beds
    Figure 5.9. Geology and trace of hydraulic fracture on the eastern wall of the trench
                  cutting the major axis of the hydraulic fracture HF12.
                                         155
    

    -------
    overlying silty clay. To the east of this zone, the fracture cuts across several beds
    and there appears to be negligible effect on the form of the trace. As the fracture
    approaches the ground surface it flattens and then curves sharply upward, a form
    resembling most other vents.
    
          The trace of HF12 shown in Figure 5.9 is remarkably similar to traces of
    HF5, HF6, and HF9, where the fractures cut relatively homogeneous silly-clay.
    Apparently, bedding in the till affected the form of the trace of HF12 only slightly-
    flattening it near the center of Fig. 5.9. The sharp trough-like form of HF12 is
    unusual, however, and could be a result of bedding or other heterogeneities in the
    till. Exposures were insufficient to make that determination.
    
    Borehole 13
    
          The fracture in the vicinity of borehole 13 is the largest one that could be
    excavated.  It is also unusual in several respects other than size. In particular, the
    fracture is nearly flat-lying and the overlying ground is sloping, whereas at the other
    fractures the ground is flat-lying and the fracture slopes.
    
          At least four lobes and a main parent fracture compose HF13 (5.10 and
    5.11). Typically, a lobe splits and either overlies or underlies the main fracture so
    that the two overlap in plan (Fig. 5.10).  This differs from fractures at the other
    boreholes where lobes are adjacent and they rarely overlap in plan. The lobes at
    HF13 are roughly parallel and separated by roughly 0.3 m of unfractured till. As a
    result, the traces of two fracture planes were commonly found on the walls of
    trenches cutting HF13 (Fig. 5.12).
    
          Junctions between the  main fracture and lobes are well-exposed in two
    locations, at the eastern and southern sides of HF13. The junctions are several m in
    length and the lines they form intersect the borehole.  The other two junctions are
    covered, and  their locations and orientations are inferred (Fig. 5.12). Double lines
    are used to indicate junctions  in Figures 5.10 and 5.11, and short solid lines at each
    end of the double lines indicate lobes associated with each junction.
    
          The hydraulic fracture is crudely basin-shaped in the vicinity of the borehole,
    but it flattens and is nearly flat-lying at distances greater than several meters from
    the borehole. Most of the southern half of the fracture, for example, differs by less
    than a few cm from horizontal.
    
          In plan view, HF13 resembles the other hydraulic fractures. It is elongate,
    with an aspect ratio of 2:1, and the borehole is near one end of the major axis. The
    longest dimension of the fracture, as measured from the borehole, lies roughly in
    the direction  of the slope of the overlying ground surface (Fig. 5.10 and 5.11).
    
          The vent at F13 is down slope from the borehole, and it occurs where the
    elevation of the ground surface is slightly greater than that of the bottom of the
    casing-the vent is 0.75 m above the bottom of the casing. The fracture curves
    sharply upward at the vent (Fig. 5.12).
    
          Two vertical fractures  occur adjacent to the borehole; one is parallel to the
    long axis of the fracture, whereas the other differs in strike by 125°. Both vertical
                                          156
    

    -------
                                                            e-
                                                            Orientation of
                                                            faint surfaca
                                                   Leading edga or rracnra
                                               Hill Vont
    
    
                                                   BWBhole
    
    
                                               13 Well 1.0.
                                                   Elevation of frx surfoce
                                                                    0.13m
                                                          Ma* batum of oahal
    Figure 5.10. Hydraulic fracture HF13 showing structural contours fracture surfaces.
                  Double lines mark the intersections of fracture lobes.
                                             157
    

    -------
                                                        8"
                                                        Orl»ntotlon of
                                                        ground *urfoc«
                                               '.•:::.7;:.:. Froetu-* cutting tondy
                                               i*>3»S flrovtl Ixd
    Figure 5.11. Outline of HF13 showing area where fracture cuts a bed of upwardly-
                  grading gravel, sand and silt Fracture cuts massive silty-clay till
                  where unstipled.
                                            158
    

    -------
    South
             West wal of trench
            Figure 5.12. Trace of hydraulic fracture HF13 and graded beds on the western wall
                         of the trench cutting the major axis of the fracture. Graded beds
                         exposed above the fracture trace have been omitted.
                                                 159
    

    -------
    fractures cut till adjacent to the bentonite seal in the borehole, and are above the
    level of the notch. The main, sub-horizontal hydraulic fracture is contiguous with
    the vertical fractures, and is unrelated to the notch (Fig. 5.11,5.12).
    
           Till in the vicinity of HF13 consists of massive silty-clay and graded beds of
    silty and sandy-gravel (Unit 2). Most of the fracture is in massive silty clay, but it
    intersects a graded bed roughly midway between the borehole and the vent (Fig.5.11
    and 5.12). The bed is 02 m at its thickest and grades from gravel at the base to silt
    at the top. The fracture cuts relatively soft silt at the top of the bed.
    
           The flat-lying orientation of HF13 is atypical compared with the other
    fractures, which are dipping. The graded bed could have affected the orientation of
    HF13 because the silt appeared to be less tough than the enveloping silty clay.
    However, much of HF13 is horizontal where it cuts through massive silty clay,
    material that lacks stratification.  It is inferred that the general flat-lying orientation
    of HF13 results from factors other than stratification. In particular, the slope of the
    overlying ground surface could have affected the dip of the fracture.  This inference
    should be checked with a theoretical analysis, but here it will allow us to ignore the
    orientation of HF13 in developing an idealized model of a hydraulic fracture.
    
    
    DIMENSIONS OF THE HYDRAULIC FRACTURES
    
           The plan areas of most of the fractures were between 10 m2 and 30 m2 (Table
    5.1), according to measurements made from the maps described in the previous
    section. One fracture (HF13) was much larger, covering an area of 90 m2, and
    another fracture was smaller (HF7 was 2.2 m2) than most of the others (Table 5.1).
    TABLE: 5.1. DIMENSIONS AND DIPS OF HYDRAULIC FRACTURES
    
    FQC          Depth       Plan Area        Max. Length      Ave. Dip
                  £m)          (m2)              (m)
    
    HF2          2.77        unknown         unknown        shallow
    HF4          3.84        unknown         unknown        shallow
    HF5          1.64          13                3.6              25°
    HF6          1.85          28                6.4              15°
    HF7          1.83          2.2               1.8           variable
    HF9          1.75          20                5.5              17°
    HF10         1.83          12                3.3            22°,25°
    HF11         1.67          9                4.1              24°
    HF12         1.98          30                8.2              14°
    HF13         1.83          90               13.5           sub-horiz.
          The greatest length of all exposed fractures occurred between the parent well
    and the vent. The maximum length of most of the fractures, where length is
                                         160
    

    -------
     measured from the injection well to the leading edge of the fracture, is typically
     between 3 and 8 m. The longest fracture is 13.5 m, and the shortest one is 1.8 m
     (Table 5.1).
    
           Ratios of maximum length to depth of initiation typically range between 1.5
     and 4, with a maximum of 7.4 and a minimum of 1.0. Fractures HF2 and HF4 did
     not vent and were too deep to excavate completely, so their lengths are unknown.
    
           Many of the hydraulic fractures are approximately planar between the
     borehole and the vent. The average dips of the fractures are similar; all of them are
     within 5° of a 20° dip. The exception is HF13, which is sub-horizontal (Table 5.1).
    
    
     DIRECTION OF PROPAGATION
    
           Most of the fractures are asymmetric in plan with respect to the borehole.
     Thus, the fractures acquire a preferred, or dominant direction of propagation as
     they grow away from their parent boreholes. The dominant direction of
     propagation, the azimuth of the line between the borehole and a vent, ranges from
     northeast to west to southeast (Table 5.2). The direction of propagation of HF13 is
     nearly parallel to the direction of maximum slope of the overlying ground surface,
     which is 225° (Fig. 5.10 and Table 5.2). At the  other fractures, however, the
     direction of propagation is unrelated to known geologic or topographic features.
    
          The direction  of propagation is related to the location of the backhoe used to
     prevent movement of the casing during fracturing. In most cases, the fractures
     propagated away from the backhoe. The two exceptions, HF7 and the southwestern
     lobe of HF10, are small fractures that propagated toward the backhoe, venting near
     the front wheels of the vehicle (Fig. 5.13). No fractures propagated beneath the
     backhoe.
    
          Theoretical  analyses indicate that hydraulic fractures will propagate hi
     directions of decreasing confining stress. Apparently the weight of the backhoe
     resulted in vertical  stress gradients that were great enough to affect the propagation
     of the underlying hydraulic fracture.
    
          The dominant direction of propagation is related to the trend of the
     hydraulic fracture at the vent, and to the vertical hydraulic fractures adjacent to the
     borehole. Strikes of vertical fractures at the vent are nearly perpendicular to the
     dominant direction of propagation in every fracture (Fig. 5.13 and Table 5.2).  At
    most fractures, the  strike of a vertical fracture at the borehole is also nearly
    perpendicular to the propagation direction (Fig. 5.13 and Table 5.2). In a few cases,
    however, the strikes of fractures adjacent to the borehole (e.g. HF9, HF11) are at
    high angles to the direction of propagation, and in one case (HF13), two vertical
    fractures differing in strike by 125° were observed. The formation of those fractures
    is poorly understood, but it could  be similar to the results of laboratory experiments
    where hydraulic fractures were  created from cylindrical holes in clay.
                                        161
    

    -------
                                       r
                                           Locotlon of
                                           hydraulic
                                           fracturing
                                           equipment
    
                                            	|
                                                Scale
    L
                                                _L
       _L
                                                5   10
                                                meters
    J
            15
                                            o Well location
                                              Backhoe
                                           /v_^—-Borehole frx
                                         fv,
                                          H_V
    Hydraulic frx
    Figure 5.13. Outlines of hydraulic fractures and locations of a backhoe at the time of
                fracturing.
                                     162
    

    -------
     TABLE 5.2: AZIMUTHS OF FEATURES OF HYDRAULIC FRACTURES.
    
    HF5
    HF6
    HF7
    HF9
    HF10
    HF11
    HF12
    HF13
    Propagation
    Direction
    325, 282
    225,10
    150
    326
    206, 39
    205,240
    175
    200
    Vent
    Fix
    60
    165
    62
    35
    124
    130
    75
    105
    Borehole
    Frx
    62
    157
    58
    117
    128
    18
    _
    265, 20
          In some experiments, two co-planar fractures were created, whereas in other
    experiments, three fractures in a trigonal pattern (striking roughly 120° from
    another) were created from the cylindrical hole. Haimson and Fairhurst (1970) also
    describe laboratory experiments where both co-planar and trigonal patterns of
    hydraulic fractures were created from cylindrical holes in rock.  Our experiments
    and those of Haimson and Fairhurst (1970) show that the two patterns of fractures
    are possible, although additional work is required to determine the factors that
    actually cause these patterns to develop.
    
    
    SUMMARY: AN IDEALIZED HYDRAULIC FRACTURE CREATED DURING
    THE FIELD TESTS
    
          Forms of hydraulic fractures created during the field test differ in detail, but
    there are certain characteristics that are common to nearly all the fractures. The
    common traits were used to infer an idealized hydraulic fracture created in the field
    tests. The form of the idealized fracture consists of four zones, which are arranged
    in increasing distance from the parent borehole (Fig. 5.14):
    
        1. Zone One: a sub-vertical orientations adjacent to the borehole.
    
       2. Zone Two: a flat-lying orientation in the vicinity of the borehole (within
            several m).
    
       3. Zone Three: a planar to trough-like feature dipping gently toward the
            borehole. This zone composes most of each hydraulic fracture.
    
       4. Zone Four: a steeply-dipping orientation occurring within several dm of the
           ground surface.
                                        163
    

    -------
     Zone One
    
           Zone One is characterized by one or more vertical fractures containing the
     axis of the borehole and initiating along the open interval below the casing, vertical
     fractures were observed at all boreholes except HF12, where the bottom of the
     borehole was poorly exposed during initial excavation and the trench was flooded
     before we could complete excavation.
    
           Two vertical fractures whose strikes differed by roughly 120° occurred at
     HF13, whereas single fractures were found adjacent to the other borings. More
     than one fracture may have occurred adjacent to some of the other borings, and they
     were removed during excavation of the trenches.
    
           Vertical fractures typically flatten upward and become sub-horizontal at the
     bottom of the basket or adjacent to the bentonite-filled segment of the borehole,
     one to several dm above the notch.  The change in orientation typically is quite
     abrupt, occurring over a length of several cm and having a radius of curvature of one
     to two cm.
    
          There are several exceptions to the above generalizations.  One is at
     borehole 7, where a vertical fracture rolls over gradually over a length of 1  to 1.5 m.
     At two other boreholes (HF10 and HF6) vertical fractures propagate a meter or
     more above the bottoms of the casing before they roll over and become sub-
     horizontal. None of the vertical fractures extended from  the bottom of a borehole
     to the ground surface.
    
          A sub-horizontal fracture propagated from the notch at borehole HF6, but
     none of the other notches contained a hydraulic fracture. The notch at HF6 was two
     cm below the basket, whereas it was four to eight cm below the basket at the other
     boreholes.
    
          The features of Zone One are a result of the nucleation of hydraulic
     fractures in the vicinity of an open borehole. It appears that nucleation occurs as a
    vertical fracture in the wall of the open boring. The vertical fracture propagates
    outward and upward and rolls over to a shallowly-dipping fracture either at or above
    the top of the open interval.
    
          The notches cut in the walls of the borehole were ineffective at nucleating
    hydraulic fractures. Presumably this occurred because the notches were too shallow
    to affect the concentration of stress in the vicinity of the borehole. A notch of larger
    diameter is expected to be more effective at nucleating a  horizontal hydraulic
    fracture at the borehole. We were unable to create a larger notch using the simple
    mechanical device designed for the project.
    
    Zone Two
    
          Zone Two is characterized by a subhorizontal fracture occurring within
    several m of the parent borehole. The subhorizontal fracture surrounds the
    borehole in some cases (e.g. HF6 and HF12).  In other cases (HF5 and HF9) the
    subhorizontal fracture  is limited to one side of the borehole, typically the side
    opposite the main body of the fracture.
                                          164
    

    -------
          The subhorizontal fracture of Zone Two is absent from some of the fractures
    (e.g. HF7, HF10 and HF11), and it is unclear whether this Zone is typical. A
    hydraulic fracture lacking Zone Two would resemble 5.14 except the vertical
    fracture of Zone One would roll over to the shallowly-dipping fracture of Zone
    Three.
    
    Zone Three
    
          Zone Three composes most of the idealized hydraulic fracture.  It is slightly
    elongate in plan, with aspect ratios between 1.5:1 and 2:1, and the major axis lies
    along a line between the parent borehole and the vent  The length of the fracture in
    Zone Three is roughly three times the depth where the  fracture rolls over in Zone
    One.
    
          The fracture in Zone Three is, in general, planar to slightly trough-like and it
         „	*_._.J .*.!_._ 1_ _ __ _. l_ ._. 1 ._. *-.+. ._._--_L1_. ^-/W\ fJm^*m _.«*_««. IB. KI^BVJB. K^ *1 ^-ft «««J ^Cft\  T«« *1»«
    dips toward the borehole at roughly 20° (dips range between 12° and 25°). In the
    field, this Zone is commonly composed of several distinct lobes, which taken
    together result in the idealized form shown in Figure 5.14.
    
          Traces of the fracture in Zone Three are remarkably straight over the length
    of the zone.  In detail, however, the traces are slightly stepped (Fig. 5.14b); they
    consist of straight segments of approximately one meter connected by risers of
    roughly 0.1 m (e.g. Figs. 3.7,3.8,3.11).
    
          In plan, Zone Three is asymmetric with respect to the borehole. At some of
    fractures (HF5, HF7 and HF10, HF11), the borehole lies at the edge of the fracture,
    whereas at others the borehole is contained within the fracture but it lies much
    closer to one end than to the other. The borehole lies between the load applied by
    the backhoe and the fracture of Zone Three.
    
          The major axis of the idealized fracture is perpendicular to the strike of the
    vertical fracture in Zone One.
    
    Zone Four
    
          In Zone Four, the idealized fracture curves upward and is subvertical where
    it intersects the ground surface. The vertical fracture in Zone Four typically extends
    to a depth of ten cm, although depths range from a maximum of 30 cm at HF9 to a
    minimum of 2 cm at the northern lobe of HF6.  The vertical fracture of Zone Four
    is 0.5 m to 1.5 m  along strike. It is perpendicular to the major axis of Zone Three,
    and it is roughly parallel to the vertical fracture in Zone One.
    
    
    DISCUSSION: DEVELOPMENT OF THE IDEALIZED FRACTURE
    
          A general history of the development and growth of the idealized fracture
     Fig. 5.14J will improve our understanding of the processes that occurred during the
     leld test.  In the  following section,  we will use the results of experiments of
    hydraulic fracturing of rock or soil and the results of theoretical analyses based on
    linear elastic fracture mechanics to infer a conceptual model of the development of
    the idealized fracture.
                                        165
    

    -------
    Load- due to
        backhoe
      I
    beokhoe
                                                                      vent
                        1  11*"^ If— .
                         :;  noToh
    
               Zone  2.1.     2
     Figure 5.14a. Idealized hydraulic fracture created at the ELDA test site.  Inferred
                 from exposures of fractures created beneath level ground, a.) Oblique
                 view, bo Section along major axis of the fracture.
                                       166
    

    -------
           When pumping begins at the start of the fracturing procedure, the pressure
    within the borehole increases, affecting the state of stress until failure occurs in the
    till enveloping the open interval of the borehole.  The initial fracture is vertical,
    containing the axis of the borehole and apparently is unrelated to the notch.
    
           Several investigators have reported results from the laboratory indicating
    that hydraulic fractures nucleated from cylindrical holes in rock will contain the axis
    of the hole, even though the applied far-field state of stress favored an orientation
    normal to the axis f Medlin and  Masse, 1979; Daneshy, 1973; Haimson and
    Fairhurst, 1969).  Similar results occurred when hydraulic fractures were created
    from cylindrical holes in soil using the apparatus described in Part I.
           This result can be explained by an analysis of stresses in the vicinit
    pressurized cylindrical hole in an elastic medium where the principal confining
    stresses act parallel or normal to the axis of the hole (Poulos and Davis, 1974). The
    analysis shows that pressure within the hole can induce tensile stresses in the
    enveloping medium that act parallel to the circumference of the hole. Stresses
    acting parallel to the axis of the hole, however, are unaffected by pressure within the
    hole. Accordingly, tensile stress adjacent to a vertical borehole develop normal to a
    vertical plane, regardless of the relative magnitudes of the principal confining
    stresses. The vertical fractures observed adjacent to the boreholes apparently result
    from those tensile stresses.
    
           A disk-shaped notch should favor the nucleation of a fracture in the plane of
    the notch, assuming the minimum applied compression  is normal to the notch.
    Indeed, this was the reason the notches were created. It is clear from the results of
    the field test that the notches had little effect on the nucleation of hydraulic
    fractures. Failure to nucleate a horizontal fracture at the borehole is a significant
    shortcoming of the design of the borehole because it reduces the overall size of the
    hydraulic fractures. This occurs because the vertical fractures at the borehole grow
    upward before rolling over, essentially reducing the depth of initiation and limiting
    the length of the fracture prior to venting.
    
           The development of a vertical fracture could be inhibited by decreasing the
    length of the open interval in the boring, and by increasing the depth of penetration
    of the notch.  Recently, we have used a high-velocity water jet to cut notches that
    extend IS to 25 cm into till, many times larger than the ones cut mechanically during
    the field test.
    
           Once formed, the vertical fracture grows outward and upward climbing
    above the open interval of the borehole. At some point, either adjacent to or
    several dm to one meter above the basket, the vertical fracture changes orientation,
    or rolls over, flattening abruptly to a shallow dip. Typically, the fractures roll over
    to a nearly  flat-lying orientation, but in some cases they dip 15° to 25° toward the
    borehole. This change of orientation appears to occur as the fracture grows out of a
    zone where stresses are influenced by the pressurized borehole and into a zone
    influenced  by the far-field stresses.
    
           The idealized fracture then grows roughly horizontally as much as several m
    away from  the borehole and changes orientation again,  curving upward to dip
    roughly 20° toward the borehole. This change in orientation is inferred to result
                                         167
    

    -------
     from a mechanical interaction between the fracture and the overlying ground
     surface. The interaction has several effects on the propagating fracture, all of which
     are related to the removal of resistance of the material above the ground surface.
     Perkins and Kern (1961), for example, analyze one effect of mechanical interaction
     by comparing the aperture of a shallow circular horizontal fracture, which dilates by
     lifting the overburden, to the aperture of a deep circular fracture, which dilates by
     compressing the overburden. They show that fracture apertures depend on the size
     of the fracture relative to its depth, expressed as the ratio of radial length a to depth
     of initiation d. Apertures of the two fractures are equal when a/d = 1.33, but the
     aperture of the fracture that lifts the overburden greatly exceeds that of the other
     fracture when a/d > 1.33.
    
           Pollard and Holzhausen (1979) show how mechanical interaction could cause
     a fracture to turn upward and propagate toward the ground surface.  They
     calculated the stress intensity factors, K\ and K\\, for a two-dimensional, static, planar
     fracture of half-length a in an elastic medium bounded by a free surface (the ground
     surface).  The direction of propagation of the fracture can be predicted using the
     relative magnitudes of K\ and An, according to theories proposed by Cottrell and
     Rice (1979) or other investigators cited by Pollard and Holzhausen (1979). Those
     theories state that non-zero values ofKu lead to propagation out of the plane of the
     original fracture. The results of Pollard and Holzhausen indicate that K\\ of a
     horizontal fracture increases as a/d increases; K\\ is negligible when a/d <  0.3, it
     increases gradually when 0.3 <  a/d < 1.0, and it increases rapidly when a/d > 1.0
     (Pollard and Holzhausen, 1979; fig 5).  This is because the upper surface of a
     horizontal fracture is displaced  farther from the axial plane than the lower surface,
     resulting in shear at the tip and non-zero values of AH.
    
           Those results suggest that the depth of initiation is critical to propagation
     path. A horizontal fracture  considerably shorter than the depth of initiation will
     propagate horizontally, but as the fracture length approaches that depth the fracture
     will tend to curve out of plane and propagate upward.
    
           The analyses described in the previous paragraph are for static fractures, and
     as such they are limited to the favored direction of a miniscule extension of a
     stationary planar fracture. Dynamic analyses, which track the  movement of a
     fracture as it propagates, show how the interactions described  above affect fracture
     form. Narandren and Cleary (1983) present the results of a dynamic analysis
     predicting the path of propagation or a horizontal hydraulic fracture beneath a free
     surface. The results (Narandren and Cleary 1983; fig. 9) show that the fracture is
     nearly flat-lying when it is short relative to its depth, but the dip increases noticeably
     when a/d > 0.7. The analysis was terminated when a/d = 1.55 and the fracture
     curves upward yielding a shape suggestive of the cross-section  of the idealized
     fracture (Fig. 5.14b and Narandren and Cleary 1983; fig. 9).
    
           Results of theoretical analyses outlined above indicate that we should expect
     a hydraulic fracture to begin to propagate toward the ground surface after it has
    grown away from the parent borehole a distance approximately equal to the depth
     of initiation.  Thus, the analyses indicate that the maximum length a fracture
     reaches before venting is scaled to its depth of initiation. This implies that we could
    have increased the sizes of the fractures created during the test by using deeper
    boreholes.
                                        168
    

    -------
           The idealized fracture grows at nearly a constant dip once it begins to
    propagate toward the ground surface (neglecting vertical growth at the borehole).
    Processes causing the dip to remain constant are unclear because available analyses
    (i.e. Pollard and Holzhausen, 1979; Narandren and Geary, 1983) suggest that the
    dip will increase as the fracture climbs and the effect of the ground surface
    increases.  One possibility is that during upward growth the vertical stress, resulting
    from the weight of the overburden at the fracture tip decreases, whereas the lateral
    stress remains constant. Thus the tendency to steepen, due to effects of the ground
    surface, is balanced by a tendency to flatten, due to an increase in the ratio of lateral
    to vertical stress at  the tip.  This possibility must still be checked
           The dip of the idealized fracture is uniform in general, but it varies slightly in
    detail, resulting in gentle step-like features. The steps suggest that slight deviations
    from the average dip are rapidly corrected by the fracture curving back to the initial
    orientation. Cotterell and Rice, who analyze the stability of slightly-curved Mode I
    cracks, show how a small deviation from the optimal path will result in a step-like
    form in a fracture (Cotterell and Rice, 1980; fig. 10).
    
           Many analyses of hydraulic fractures assume that they are symmetric with
    respect to the axis of the borehole, but most of the fractures created in the field test
    were asymmetric.  The hydraulic fracture created beneath sloping ground (HF13) is
    elongate in the downslope direction. Where the ground is level, however, applied
    loads seem to influence the direction of propagation because the dominant
    directions of propagation are typically away from the backhoe parked near each
    borehole. Apparently gradients in vertical stress, caused by either topography or
    applied loads, affect the dominant direction of propagation  of horizontal or
    shallowly-dipping hydraulic fractures. This conclusion raises the intriguing
    possibility of artificially loading the ground surface to cause fractures to propagate
    in a particular direction.
    
           It is reasonable to expect that hydraulic fractures of similar size and shape as
    those described here could be created under conditions similar to the ELD A site. It
    would be misleading, however, to suggest that similar hydraulic fractures can be
    expected at any site. The large lateral stresses in the till at the ELD A site certainly
    affected the dips of the fractures, and thus the sizes that they could achieve before
    venting. Measuring in situ stresses will be a vital first step in predicting the
    orientation of hydraulic fractures.
    
           The principles of linear elastic fracture mechanics, specifically the
    applications to  hydraulic fracturing of rock, seem to offer general explanations of
    the development of the idealized fracture created in till at the ELD A site.
    Predicting the forms of hydraulic fractures at other sites, however, will require a
    theoretical model specifically tailored to analyze conditions of the near-surface.
                                       169
    

    -------
                                  SECTION SIX
    
                SETTING AND DESIGN OF THE FIELD TEST - 1989
          Hydraulic fracturing field tests of 1989 were performed at two sites. Oni
          me as the field test of 1988, located 10 km north of downtown Cincinnati:
                                                                      One was
    the same as the field test of 1988, located 10 km north of downtown Cincinnati near
    the southeastern corner of the ELD A Company landfill. The other is located IS
    miles north of the ELD A landfill, on the grounds of the Goettle Construction
    Company.
          Hydraulic fracturing at the ELD A landfill site took place in June and July,
    1989. Field analysis of the fractures, including trench mapping, took place in July
    and August, 1989.  Preliminary testing of the hydraulic fracturing apparatus was
    performed at the Goettle Construction Company site in March and April, 1989.
    Hydraulic fractures for fracture monitoring purposes were created at the Goettle
    site in July, 1989. Fractures created at the ELDA site were exposed by excavating a
    network of trenches, whereas fractures at the Goettle site were not exposed. As a
    result, more is known about the fractures and subsurface at the ELDA site,  and the
    following chapters will focus on that site.
    
    
    QUALITY ASSURANCE AND CONTROL
    
          The success of the hydraulic fracturing effort is one measure of quality
    assurance steps taken during the execution of this project. The fact that fractures of
    similar shape and size were created repeatedly, at different times and locations, is a
    measure or the success of quality control.
    
          All instruments used during the fracturing process were calibrated according
    to industry standards. Calibration of custom designed instruments, such as the
    gauge for measuring the flow rate of the injection pump, was conducted in a series
    of identical tests, performed by different staff researchers. Dosages of chemical
    ingredients of fracturing gels were measured with margins of error no greater than
    0.1%, hence exceeding petroleum industry standards. Accuracy achieved in
    measuring borehole depths and notch lengths was comparable to accuracy
    demanded by field analysis of the fractures (see Section 7).
    
    
    SITE CHARACTERISTICS
    
          The vicinity of the ELDA test site is an area of gently sloping ground
    bounded on the southwest by a steep slope, three to four meters tall, and on the
    northeast by a moderate to steep descent into  a deep (> 10 m) trough. The  ground
    surface at the time of the test was the product of sou excavation for fill purposes,
    and it was three to four meters below the natural ground surface.  The ELDA
    fracturing site is a gently sloping, elongate strip, trending N55W, roughly 50 m long
    and 10 m wide.
    
          The area used during 1989 is in the vicinity of the area used during 1988. We
    estimate that the southeastern end of the 1989 site is 40 to 60 m south of the
    southwestern corner of the area shown in Figure 4.2. Accurate correlation of the
                                       170
    

    -------
     two sites was not attempted because all the features shown in Figure 4.2 were
     removed during excavation activities between autumn 1988 and spring 1989.
     Geology
           The ELD A site is underlain by glacial till (Illinoian?), composed mostly of
     clayey silt to silty clay. The stratigraphic level is slightly higher (5 to 10 m) than that
     of the site of the 1988 test. A two meter section of the stratigraphy, exposed in
     trenches used to examine the fractures, is described in the following section.
     Description of Soil Profiles in Trenches
           The soil profile is generally uniform in the SE-NW direction, according to
     observations in trenches.  Gravelly till lies at the top of the soil sectionjust beneath
     the ground surface. The thickness of the gravelly till increases to the NW, from 20
     cm at the SE end of trench A, to 30 cm at the midpoint of trench A, 20 m away from
     the SE end, to 50 cm at the NW end of the trench.
    
           Organic-rich silty clay to clayey silt  underlies the gravelly till. The thickness
     of the organic silty clay is 40 cm at both ends of trench A, but it pinches to half that
     thickness at the midpoint of the trench.
    
           Massive brown silt, 25 to 30 cm thick, underlies the organic silt. At the NW
     end of the trench, the massive silt is directly underlain by a 90 cm thick unit of
     laminated silty clay, with interbedded clayey silt.  At the SE end and at the midpoint
     of the trench, the massive silt and the laminated silty clay are separated by other
     units. At the midpoint of the trench, a 50 cm thick sequence of interbedded, 10 cm
     thick, laminated clay and 10 to 20 cm thick massive brown silt underlies the upper
     massive silt. At the SE end, 50 cm of clayey silt with interbedded silty clay lie below
     the massive silt. At both the SE end and the midpoint of the trench, the 50 cm thick
     sequence of silt and clay is underlain by 105 to 110 cm of laminated silty clay with
     interbedded clayey silt.
    
           At the NW end of the trench, the laminated silty clay is underlain by 20 cm of
     loam, followed by 20 cm of gravelly till, at the bottom of the trench. At the
     midpoint of the trench, the laminated silty  clay is underlain by 10 cm of gravel at the
     bottom of the trench, and at the SE end of the trench, the laminated silty clay
     reaches to the bottom of the trench.
    Moisture Contents
           Moisture contents of three samples extracted from the western intersection
    of Trenches A and F, were measured in accordance with the ASTM Standard
    Designation D 2216-80. The three samples represent three different soil types,
    described above, under Site Characteristics.
    
           Sample 1, extracted 77 cm above the horizontal datum, represents the
    organic-rich silty clay to clayey silt, which occurred just below the gravelly till
    throughout the trench network.  Its moisture content is 23.7%.
                                         171
    

    -------
           Sample 2, extracted 44 cm above the datum line, represents the massive,
    brown silt, lenses of which also extend across the entire trench network.  Its
    moisture content is 14.6%.
    
           Sample 3, extracted 15 cm above the datum, consists of laminated siltv clay.
    Although the layer from which Sample 3 was removed lies above the lowest lens of
    massive silt, the sample represents the same type of soil that constitutes the one-
    meter-thick, laminated silty clay unit pervasive in the lower half of the entire trench
    network. Its moisture content is 25.2%.
    METHOD OF FRACTURING
    
           Figure 6.1 illustrates the equipment and sequence of operations involved in
    hydraulic fracturing. Figure 6.2 shows details of the drive head used to prepare a
    borehole for hydraulic fracturing. Figure 63 outlines five steps of hydraulic
    fracturing.
    
    Equipment
    
           Equipment used to prepare and pump gel and sand mixtures for hydraulic
    fracturing included two water tanks, two gel mixing tanks (labeled a in Figure 6.1),
    two blenders for crosslinking and adding sand (d in Figure 6.1), a water pump, a gel
    pump (b in Figure 6.1), hoses for gel transfer (c in Figure 6.1), and a grout pump to
    inject proppant into the subsurface (e in Figure 6.1). Equipment used to create
    boreholes for fracturing included rod and casing (fin Figure 6.1; also see Figure
    6.2), a drive head with pointed tip (Figure 6.2), a water jet rod to create starter
    notches, a high-pressure pump for the water jet, and an electrically powered
    jackhammer to drive the fracture lance into the ground. Monitoring equipment
    included an IBM-PC compatible, portable (laptop) computer (h in Figure 6.1), a
    pressure transducer (g in Figure 6.1), and a motorized power generator.
    
           The two water tanks were made of plastic, and had capacities of 1.1 nr1 (300
    gal.) each. The two^longate eel tanks were made of galvanized steel, and had
    capacities of 0.95 nr (250 gal.) each. The two blenders for crosslinking and adding
    sand were part of a grouting pump system by Chem-Grout Company of Chicago,
    Illinois, and they had capacities of 0.2 nr (50 gal.) each. Rotary mixing blades
    within the blenders were powered by a diesel engine, which also powered the
    moyno-type grouting pump. Crosslinked gel and sand were fed from the blenders
    directly into the grouting pump, which pumped the blend downhole and into the
    fractures.
    
           At the Goettle site, a five horsepower centrifugal pump moved water from a
    nearby lake into the two water tanks, and from the water tanks into the gel tanks.
    At the ELDA site, the water pump was not needed, because the water tanks were
    filled directly from a water  truck, and water flowed downhill from the water tanks
    into the gel tanks.
    
           An eight horsepower, 7.6 cm (3 in) centrifugal pump rapidly circulated gel in
    the gel tank to hydrate the guar gum, and to prevent the formation of lumps, or
    "fisheyes," in the gel. The same pump also moved gel from the gel tanks into the
                                     172
    

    -------
        a.   mixing tank
        b.   circulation pump
        o.   valve
        d    blender
        e    Injection pump
        f.   borehole
        g.   transducer
        h.   data acquisition
    Figure 6.1. Scheme for hydraulic fracturing operation performed during the 1989
              field test
                                173
    

    -------
     crosslinking blenders. Gel circulated from the gel tank, into the pump, and back
     into the geltank through 5 cm (2 in) tubing (Figure 6.1). A "T" valve (labeled c in
     Figure 6.1) in the tubing diverted the hydrated gel into the crosslinking blenders,
     also through plastic tubing.
    
           A Zenith Supersport 286 laptop personal computer, with 40 megabyte hard-
     drive was linked to the pressure transducer, which was attached to the borehole
     casing, in order  to monitor pressure changes in the proppant pumping hose (the
     pipeline) at or above ground surface.
    
    
     The Injection Fluid
    
           Guar gum-based chemical mixtures were added to the injected water to
     create a gel capable of carrying sand into the fractures. A complex derivative of
     guar gum, manufactured by Halliburton Services company - a major service
     company to oilfields in the United States - was used in the first four fractures
     created at the ELDA site. A simple guar gum, manufactured by Hi-Tek Polymers,
     Inc., was used in the remaining fractures, including twelve at the ELDA site, and
     five at the Goettle site.
    
           Comparison of the two gel products led to the conclusion that the simple,
     nonderivitized guar gum system manufactured by Hi-Tek Polymers is better suited
     for shallow, environmental applications than the highly derivitized system
     manufactured by Halliburton Services. The chemical kinetics of the nonderivitized
     gel corresponds  better to the requirements of fracturing at shallow depths.
    
           Guar gum-based  gels consist of four main ingredients, which are added to
     water in a particular, timed order, causing a three-step chemical process. The
     ingredients include guar gum powder, buffering powder or solution, crosslinking
     powder or solution, and  breaking powder or solution. The three steps are, in order,
     gelling, crosslinking, and breaking. Gelling and crosslinking increase the viscosity of
     water by one and two orders of magnitude, respectively.  Breaking reduces the
     viscosity back down to almost that of water (1 cP), after a period of time sufficiently
     long to allow completion of the fractures.
    
           Gelling increases the viscosity of water from 1 cP to about 15 to 20 cP,
     caused by the addition of about 3.6 to 4.8 kgs of guar gum powder per cubic meter
     (0.03 to 0.04 Ibs/gal) of buffered, pH-neutral water. Halliburton's WG-18 gum
     powder, known as a CMHPG (Carboxyl-methyl-hydroxyl-propyl-guar)  gum, was
     required in doses of 4.8 to 5.03 kgs/nr (0.04 to 0.042 Ibs/gal) to obtain the
     necessary viscosity. Doses of 3.6 kg/mj (0.03 Ibs/gal) were insufficient. Hi-Tek
     Polymer's Jaguar 408-D gum powder, known as a "standard guar," was required in
     doses of 036 to 3.95 kg/nr3 (0.03 to 0.033 Ibs/gal).
          Halliburton's WG-18 guar gum required 0.3 kg/m3 (0.0025 Ibs/gal) of an
    acidic buffering agent, containing 60% sulfamic acid, supplied by Halliburton
    Service Company under the name of BA-2.  In conjunction with Hi-Tek Polymer's
    Jaguar 408-D guar gum, a buffer is required only if the water is pH-neutral or
    slightly caustic, because the guar gum increases the pH of neutral (pH=7.0) water to
    about 7.5 or 8.0, whereas a pH value of 7.0 is needed for the successful hydration of
                                        174
    

    -------
     the guar gum powder to form a gel.  Cincinnati City Water used at the ELD A site
     was slightly acidic, so that a buffer was not needed.  Lake water used at the Goettle
     site was slightly caustic, with a pH of 7.5 to 8.0, so that a buffering agent was needed.
     Calcium chloride, added in doses of 4.8 to 5.4 kg/nr (0.04 to 0.045 Ibs/gal), served
     both as a buffer, and as an inhibitor of swelling in clays. Calcium chloride crystals
     were added to Goettle Lake Water before the guar gum powder, to allow sufficient
     time for the crystals to dissolve.
    
           Once guar gum powder is added to water, sufficient time must pass for the
     guar to hydrate before the required viscosity is reached and a crosslinker may be
     added. The hydration time is only about 5 minutes for the Halliburton gel, whereas
     it is 20 to 30 minutes for the Hi-Tek Polymers gel. This is the first of two critical
     differences between the chemical kinetics of the two gel systems applied during
     fracturing at the ELDA site (only Hi-Tek Polymers gel was applied at the Goettle
     site for the fractures described in this report). Both gel systems require several
     minutes of agitated mixing of the guar gum, to avoia formation of small lumps of
     gum, called "fish-eyes" in the oilfield hydraulic fracturing industry.
    
           The viscosity of the hydrated gel is increased to 100 to 200 cP by the addition
     of a crosslinker.  The crosslinker supplied by Halliburton, called CL-19, contains 11
     to 30% ethyJene glycol, and it was added to the hydrated gel in doses of 1.9 to 2.8
     cm* per cnT (0.0005 to 0.00075 gal per gal).  The Hi-Tek Polymers gel required
     borate as a crosslinker, available under the brand name "Borax," and it was added to
     the hydrated gel in doses of 0.24 to 0.68 kg/nr3 (0.002 to 0.0057 Ibs/gal).  The
     kinetics of the crosslinking reaction are different for the two systems compared in
     this study. The Halliburton gel typically requires five to six minutes of vigorous
     agitation for crosslinking to occur. The High-Tek Polymers gel, on the other hand,
     is crosslinked almost instantly, within seconds after the crosslinker is introduced
     under vigorous agitation. This is the second of the two critical differences between
     the chemical kinetics of the Halliburton and Hi-Tek Polymer gel systems.
    
           Concomitantly to the addition of crosslinker, breaker was added to ensure
     the eventual breakdown of the gel to a liquid  solution with viscosity almost equal to
     that of water. Laboratory tests showed the final viscosity of broken gels to be about
     2 cP, for both the Halliburton and Hi-Tek Polymers gels.  The breaker supplied by
     Halliburton, called GBW-3, contains at least 60% of an unspecified carbohydrate
     material, and it was added in doses of 0.095 to 0.099 kg/nr (0.0008 to 0.00083
     Ibs/gal).  The breaker supplied by Hi-Tek Polymers, called Breaker-F, contains a
     cellulase enzyme, and it was added in doses of 0.12 to 0.30 kg/nr (0.001 to 0.0025
     Ibs/gal).  The time required to break down the gel to the final viscosity was 18 to 24
     hours for the Halliburton gel, and 24 to 48 hours for the Hi-Tek Polymers gel.
    Although this contrast between the Halliburton and Hi-Tek Polymers systems may
    influence scheduling of the application of newly created fractures for remediation
    purposes, the contrast is nonessential because it does not affect the fracturing
    process itself.
    
          The two critical differences between the chemical kinetics of the Halliburton
    and Hi-Tek Polymers gel systems noted above - the different times required for
    hydration and crosslinking of the gels - render the Hi-Tek Polymers gel system
    more suitable for application to hydraulic fracturing for remediation purposes. The
    relatively small volumes required for our application makes the long hydration times
                                      175
    

    -------
    necessary for Hi-Tek Polymers gel tolerable, since tanks of hydrated but
    uncrosshnked gel sufficiently large to fill several fractures can be prepared in
    advance. Moreover, the pumping rate of the system we used is limited by the time
    required to completely mix a batch of gel, crosslink and sand, so the rapid
    crosslinking of the High Tek Polymers gel decreased the time required to mix each
    batch and thus improved the efficiency of our procedure. The gel and crosslink
    system from Halliburton is designed for procedures used to create hydraulic
    fractures in the oil industry, which benefit from short hydration times and delayed
    crosslinking.
    
    The Fracturing Procedure
    
          The borehole used to create fractures during the 1989 tests was designed to
    inhibit the formation of vertical fractures adjacent to the bore (as in the 1988 tests),
    and to facilitate the creation of stacks of multiple fractures. A device (Fig. 6.2),
    which we term a "fracturing lance", is the central element of the design. The lance,
    illustrated in Figure 6.2, consists of casing (EX casing) and an inner rod (EW drill
    rod), both of which are tipped at one end with hardened cutting surfaces that are
    formed into a conical point. A drive head at the other end of the lance secures both
    the casing and the rod. Individual segments of rod and casing were 1.52 (5 ft) long,
    and they were threaded together as required by borehole depth, which ranged from
    9.6 to 3.9m (2 ft to 12 ft).
    
          To prepare for fracturing, the lance was driven 0.1 to 0.2 m below the bottom
    of a borehole; either an open hole or through a hollow-stem auger (Fig. 63).  The
    rod was removed leaving till exposed at the bottom of the casing. Lateral pressure
    of the soil on the wall of the casing effectively sealed the casing in the ground,  much
    as cement sealed steel pipe in the borehole during the 1988 tests. Another
    apparatus, composed of steel tubing with a narrow (diam. 0.025 cm; 0.01 in) orifice
    at one end, was inserted into the casing.  Water injected into that apparatus with a
    pressure washer pump (rated at 17.2 kPa; 2500 psi,0.02 nr/rnin; 5 gpm) formed a
    jet that cut laterally into the soil. The water jet was rotated, producing a disk-
    shaped notch extending up to 40 cm away from the borehole (Fig. 6.3). A simple
    device, built from a steel tape extending the length of a tube and making a right
    angle bend at the end of the tube, was inserted into the casing and used to verify and
    measure the radius of the slot. A hydraulic fracture was created, using a procedure
    that will be described in detail below, and then the rod  and point were reinserted
    and the lance advanced from 7 to 50 cm. The rod and point were retracted and the
    procedure repeated, creating a second fracture 7 to 50 cm below the first one.
    Several fractures were created, stacked one on top of another, from a common
    borehole.
    
          The actual procedure of hydraulic fracturing was commenced by engaging
    the injection pump-a Robbins and Meyers Moyno 2J6  CDQjmmp was on the
    grouting rig used in the test  (e in Figure 6.1). At least 0.02 m~ (5 gall of water were
    pumped downhole first, followed by a. pad of at least 0.02 nr (5 gal) of
    uncrosslinked gel (without sand). The gel was crosslinked and sand added in
    concentrations ranging from 240 to 1438 kg of sand per cubic meter of gel (2 to 12
    Ibs/gal). Two sizes of sand were used in the  tests; #5 sand ranges from 0.8 to  1.8
    mm with a median of 1.5 mm, whereas #7 sand ranges  from 0.4 to 0.9 mm with a
                                      176
    

    -------
                                      Drive  head
                                       Cutting sleeve
                                  Point
    Figure 6.2. Fracturing lance, used to prepare boreholes for hydraulic fracturing
               during the 1989 field test
                                 177
    

    -------
         1
      Casing^
    Extension
       rod
                Lance tip
                                                      Pressure from coll
                                                        seals casing
                                        Hydraulic fracture
    Figure 63. Five steps of hydraulic fracturing: (1) create borehole; (2) extend
                borehole beyond casing; (3) remove rod and tip of fracturing lance;
                (4) create starter notch with water jet; (5) create fracture by injecting
                proppant into notch.
                                     178
    

    -------
     median of 0.75 mm.  The sand-laden gel was pumped downhole at a rate of 0.02 to
     0.04 mr/min (5 to 10 gpm).
    
           Injection pressure measured at the top of the borehole increased rapidly to at
     least 138 kPa (20 psi). As soon as fracturing commences - typically about 30
     seconds after initial engagement of the injection pump - the pressure dropped
     abruptly to about 68.9 kPa (10 psi). As pumping progressed, the pressure gradually
     decreased, although changes in gel viscosity, sand concentration, or pumping rate
     cause temporary pressure fluctuations.
    
           Assuming unlimited supply of gel and sand, pumping of proppant into the
     growing fracture would continue until either the fracture vents,  or the pump jams.
     The latter occurs if there is "screen-out", either at the tip of the  fracture or at the
     borehole, causing immobilized sand to back up into the injection hose, or even into
     the injection pump. At that point, pressure measured by the transducer increases
     dramatically, and the operation must cease.
    
           Most fracturing episodes performed as part of this study were complete when
     the planned volume of proppant had been injected.
    
    
     Discussion of the Fracturing Method
    
           The 1989 tests showed that hydraulic fractures can be created and propped
     with sand using a simple apparatus intended for injection grouting. An apparatus
     similar to the one we used should be readily available for rent throughout the
     United States, and several manufacturers offer similar units for sale.
    
           Use of the lance is more rapid than drilling a borehole and cementing casing,
     as we did in  1988.  Moreover, the technique is versatile; a hole fractured using the
     lance method can be completed with a gravel pack and screen, or the hole can be
     left unscreened if a well is not required.
    
           The design of the fracturing lance facilitates the creation of multiple
     fractures.  Of course, methods of creating multiple fractures are well-known in the
     oil industry, but they commonly require sophisticated packers and explosive charges
     to perforate  casing. The technique used in this research is intended to take
     advantage of the shallow depths and penetrability of soil, where those sophisticated
     methods are unnecessary.
    
           There are several drawbacks to the method used during the 1989 tests. The
     blenders were underpowered, so that mixing sand with crosslinked gel was time
     consuming. As a result, the rate of blending a batch of sand-laden gel was
     increased, limiting the injection rate. A more powerful blender, and perhaps
     another blender design would improve the design.
    
           The progressive cavity pump used to inject slurry suffered excessive wear,
     and the rotor and stator had to be replaced after pumping almost llm3 (2900 gal)
    of gel-sand mixtures. The progressive cavity grouting pump, while effective for high
    concentrations of #7 (0.75 mm grain size) size sand, proved ineffective - it jammed
    -  for even moderate to low concentrations of the coarser (1.5 mm grain size), #5
                                       179
    

    -------
    sand. We are currently evaluating several other styles of pump for future
    applications.
    
          Use of the fracturing lance requires creating hydraulic fractures during
    drilling of a borehole, rather than after completion or the well. This means that
    drilling is interupted for fracturing to take place, a practice that would increase the
    time required to drill each well. Techniques of isolating a zone to be fractured using
    straddle packers, a common practice in oil wells and water wells, are currently being
    considered as an alternative to the use of the fracturing lance. In some cases,
    creating a fracture below the bottom of a borehole and then drilling down through it
    could result in the deposition of mud, as a skin between the bore and the fracture.
    Completion and development techniques, which eliminate well-bore skin from the
    vicinity of hydraulic fractures, will be crucial to obtaining the maximum
    performance from fractured wells.
                                      180
    

    -------
                                  SECTION SEVEN
    
        HYDRAULIC FRACTURES CREATED DURING THE FIELD TEST - 1989
    
           In contrast to fractures created during the field test of 1988, fractures created
     in 1989 are filled with 5 to 25 mm of sand, making them clearly visible. As in the
     1988 field test, fractures were excavated with a backhoe, creating a network of
     trenches two m deep.  Detailed mapping of the trench walls revealed the geometry
     of the fractures.
    
           A 40 m long, longitudinal trench (trench A, Fig. 7.1) intersects all seven
     boreholes, and roughly coincides with the minor  axes of the elliptical fractures. A 5
     to IS m long, transverse trench also intersects each of the seven boreholes (trenches
     B - H), roughly coinciding with the major axes of the fractures. A 6 m long trench
     (trench I), parallel to the longitudinal trench, intersects one of the transverse
     trenches (trench G).
    
           A horizontal datum line, extending along  the entire trench network, was used
     during mapping. Smooth trench surfaces were cut by knives to allow precise
     measurement of fracture thicknesses. Dyed proppant sand was barely
     distinguishable from undyed sand, but the thickness of proppant made dye
     unnecessary, in contrast to fractures created during the 1988 field test.
    
           Excavation of the trenches was completed one week after creation of the last
     fracture. Mapping and analysis of the fractures commenced immediately, and
     continued for three weeks. This section presents the results of mapping and
     measuring fracture traces in the nine trenches.
    
    
     QUALITY ASSURANCE AND CONTROL
    
          To assure consistent quality of data, measurements were periodically
     repeated by different field investigators. Datum lines were checked daily to correct
     for sagging or other potential sources of inaccuracy.
    
          Accuracy achieved in field analysis of fractures was comparable to accuracy
     standards for similar analyses of structural field data related to geomechanical
     studies, or to petroleum exploration and production.
    
    
     FORMS OF THE HYDRAULIC FRACTURES
    
          At the ELDA site, we created 20 fractures from 7 boreholes, spaced 3.5 to 10
     m apart horizontally, and roughly defining a straight line, as shown in Figure 7.1.
    The boreholes are located on a 10 to 15 m wide bench, bounded to the southwest by
    a 5 m tall, steep ascent, and to the northeast by a  10 m deep descent. The boreholes
    are located 2 to 5 m away from the wall. Individual boreholes contain 2 to 5
    fractures, spaced 0.1 to 03 m apart vertically at the borehole, as indicated in Table
                                      181
    

    -------
                         a-     c
    oo
    to
                    0             10
                       meters
                                         Contours  In  meters
                                       Figure 7.1. Map of Elda Landfill site.  Solid lines are
                                                 topographic contours; dotted lines are trench
                                                 outlines.  Locations of boreholes and cross
                                                 sections are also shown.
    

    -------
    TABLE 7.1: FRACTURE SIZE
    Fracture
    EL1F1
    EL1F2
    EL2F1
    EL2F2
    EL3F1
    EL3F2
    EUF1
    EUF2
    EL4F3
    EL/4F4
    EL5F1
    EL5F2
    EL5F3
    EL5F4
    EL5F5
    EL6F1
    EL6F2
    EL6F3
    EL7F1
    -EL7F2
    Depth
    (m)
    1.5
    1.7
    1.5
    1.7
    1.5
    1.7
    0.9
    1.2
    1.5
    1.8
    1.1
    1.2
    1.5
    1.6
    1.6
    1.4
    1.6
    1.9
    1.3
    —
    Major
    Axis
    (m)
    8.0
    4.5
    5.0
    3.0
    8.0
    5.0
    5.0
    7.0
    5.0
    4.5
    8.0
    7.0
    3.0
    1.0
    ~_
    6.0
    8.5
    5.5
    7.0
    7.0
    Minor
    Axis
    (m)
    5.0
    2.0
    3.5
    2.0
    3.5
    2.5
    4.0
    4.0
    4.5
    3.5
    5.0
    6.0
    2.5
    1.0
    ....
    5.0
    5.5
    4.5
    5.5
    5.0
    Max Dist
    From BH
    (m)
    7.2
    3.8
    4.0
    1.2
    6.0
    4.6
    3.8
    4.6
    3.0
    2.5
    3.2
    3.8
    2.0
    0.8
    ....
    5.0
    7.0
    4.0
    4.4
    5.8
    Aspect
    Ratio
    (m)
    1.6
    2.3
    1.4
    1.5
    2.3
    2.0
    13
    1.8
    1.1
    1.3
    1.6
    12
    12
    1.0
    .._
    1.2
    15
    1.2
    1.3
    1.4
    Area
    (m2)
    31.4
    7.1
    13.7
    4.7
    22.0
    9.8
    15.7
    22.0
    17.7
    12.4
    31.4
    33.0
    5.9
    0.8
    ..._
    23.6
    36.7
    19.4
    30.2
    27.5
                             183
    

    -------
           \ c \
    Figure 7.2. Fracture map of trenches B and C. Thickest line shows outline of
               topmost fracture.
                                      184
    

    -------
                                                    meter
                                                       D
    Figure 7.3. Fracture map of trenches D, E and F. Thickest line shows outline of
               topmost fractures.
                                    185
    

    -------
          A\
          \H ..'
             meter
    Figure 7.4. Fracture map of trenches G, H and I. Thickest line shows outline of
               topmost fractures.
                                     186
    

    -------
     Outlines and Dips of Fractures
    
           A network of nine interconnected trenches, shown in Figure 7.1, exposes the
     fractures and reveals their shapes. Most of the fractures are elliptical in outline,
     with aspect ratios (length of major axis divided by length of minor axis) ranging from
     1.0 (fracture EL5F4 in Figure 7.4) to 2.3 (fracture EL6F2 in Figure 7.2; Table 7.1).
     The major axes of the fractures are roughly parallel to the dip direction of the
     ground surface, which is generally parallel to the transverse trenches (trenches B -
     H) and perpendicular to the wall along the southwest edge of the bench (Figures
     7.1-7.4).
    
           Many of the fractures are flat and subhorizontal in cross section, with dips
     ranging from 0 degrees (fractures E16F2 in Figure 7.12, E13F1 in figure 7.14, E14F1
     in Figure 7.17, and EL5F1 and EL5F3 in Figure 7.18); to 5 degrees (fractures
     EL3F2 in Figure 7.14, and EL1F1 in Figure 7.15). Some fractures are subhorizontal
     near the borehole, and dip toward the borehole by up to IS degrees near the
     fracture margin,'thus resembling a shallow bowl or plate (fractures EL6F1 in Figure
     7.12, EL7F2 In Figure 7.13, and EL4F3 and EL4F4 in Figure 7.17).
    
           The deepest point of each fracture, which coincides with the borehole in
     almost all cases, is located near the end of the major axis, close to the wall along the
     southwest edge of the bench (Figures  7.1-7.4). The only notable exception is
     fracture EL4F5 (Figure 7.17), in which the deepest point is located approximately
     1.3 m north of the borehole, and the borehole is located near the center of the
     fracture, but closer to the downslope (northeast)  edge of the fracture than to the
     wall along the southwest side of the bench (Figures 7.1 and 7.17).
    
    
     Venting of Fractures
    
           Five of the fractures created at the ELD A site vented onto the slope
     northeast of the bench. Fractures EL1F1 and EL2F2 both vented along the same
     horizon. Fractures EL2F1 and EL3F1 each vented along horizons above the EL1F1
    venting horizon. Finally, fracture EL3F2 vented along both the EL3F1 and EL1F1
    venting horizons.  The vertical distance between the three venting horizons is about
    30cm.
    
           Fluid that vented from all of those fractures was virtually devoid of sand,
    even though the fluid was laden with sand when injected.  Sand proppant filled all
    the fractures, but it terminated several m before the point of venting, according to
    exposures on transverse trenches cut along lines between boreholes and vents.  Two
    scenarios seem possible: either the aperture of the fractures in the vicinity of the
    vents was too small to permit the passage of sand, or sand was deposited because
    the gel that was carrying it leaked off-this is termed "tip-screenout" by Smith and
    others (1987).  In either case, the gel was separated from the sand by straining
    through immobilized sand at the fracture tip.
                                       187
    

    -------
                  5 ra
                                      EL6
        B
    B
    Figure 7.5. Cross section B-B'. See Figure 7.1 for location.
                 5  m
                                       EL7
    Figure 7.6. Cross section C-C. See Figure 7.1 for location.
                                      188
    

    -------
    Figure 7.7. Cross section D-D'. See Figure 7.1 for location.
    Figure 7.8. Cross section E-E'. See Figure 7.1 for location.
    Figure 7.9. Cross section F-F. See Figure 7.1 for location.
                                189
    

    -------
    Figure 7.10. Cross section G-G'. See Figure 7.1 for location.
                                 EL5
         H
    Figure 7.11. Cross section H-H1. See Figure 7.1 for location.
                                       190
    

    -------
    I  meter
                                                    EL6
                    TRENCH B
    EAST  WALL
             Figure 7.12. Map of east wall of trench B, showing fracture
                      traces, borehole and ground surface (dashed).
    

    -------
    1  meter
                                                EL7
                  TRENCH  C
    EAST  WALL
              Figure 7.13. Map of east wall of trench C, showing fracture
                        traces, borehole and ground surface (dashed).
                        Thick fracture is filled with perlite.
    

    -------
                              1 METER
                                                                     EL3
    vo
                                       TRENCH  D
    EAST  WALL
                                  Figure 7.14. Map of east wall of trench D, showing fracture
                                           traces, borehole and ground surface (dashed).
    

    -------
    1  meter                                          ELI
                   TRENCH  E    -   EAST  WALL
      Figure 7.15. Map of east wall of trench E, showing fracture
               traces, borehole and ground surface (dashed).
    

    -------
                     1  meter
    VO
                                          TRENCH  F
    EAST  WALL
                                  Figure 7.16. Map of east wall of trench F, showing fracture
                                           traces, borehole and ground surface (dashed).
    

    -------
                             1 meter
    vo
                                                                EL4
                                          TRENCH G   -    EAST WALL
                               Figure 7.17. Map of east wall of trench G, showing fracture
                                        traces, borehole and ground surface (dashed).
    

    -------
    10
                          1  meter
                                                   EL5
                                        TRENCH  H   -    EAST  WALL
                                    Figure 7.18. Map of east wall of trench H, showing fracture
                                            traces, borehole and ground surface (dashed).
    

    -------
                                  1 meter
    00
                                         TRENCH  I    -    NORTH   WALL
                                Figure 7.19. Map of north wall of trench I, showing fracture
                                        traces and ground surface (dashed).
    

    -------
     Continuity of Fractures
    
           Traces of fractures along trench walls, roughly coincidental with the major
     axes, are commonly continuous. Continuous fractures represent over half (61%) of
     the 18 fractures exposed along major-axis cross sections, including fractures EL3F1-
     F2 (Figure 7.14), EUF1-F4 (Figure 7.17), EL6F1-F2 (Figure 7.12), and EL7F1
     (Figure 7.13). The remaining 39% of fractures exhibit at least one discontinuity
     along the major-axis cross section.
    
           Discontinuities along fractures either step up or step down, in the direction
     pointing away from the borehole (toward the fracture tip). Most discontinuous
     fractures step up by 2 to 12  cm, at lateral intervals of 1 to 2 m. Up-stepping
     fractures include EL2F1 (Figure 7.16), EL5F1-F2 (Figure 7.18), and EL7F2 (Figure
     7.13), representing 22% of all fractures. Some fractures step downward by 4 to 8
     cm, about 0.8 of the distance from the borehole to the fracture tip. Down-stepping
     fractures include EL1F1 (Figure 7.15) and EL6F3 (Figure 7.12), representing 11%
     of the fractures. Finally, one fracture, EL1F2 (Figure 7.15), steps up by 2 cm at a
     distance of 1.5 m from the borehole, and steps back down again by 2 cm at a
     distance of 1.7 m from the borehole. This up-down-stepping fracture represents 6%
     of the total fracture count.
    
           Fractures that are discontinuous on a trench wall are expected to be joined
     somewhere out of the plane of the wall.  Exposures were insufficient to verify this
     expectation.
    
    
     Fracture Margins
    
           The cross sections exposed in the trenches do not afford a clear view of
     fracture margins, since the terminations of fractures occur only as single points in
     the cross sections.  We did not excavate a fracture surface at its margin, due to time
     constraints.  Our assumption, hence, is that the fracture margins are generally
     smooth, as shown in Figures 12-1 A.
    
           Some evidence suggests that fracture margins may be lobate. Trench I,
     which reveals the only cross section perpendicular to the direction of proppant flow
     and at some distance away from the borehole, exposes the trace of merged fractures
     EL4F2 and EL5F2 (Figure 7.19). Fracture EL4F2 contains weight #5 sand,
    whereas fracture EL5F2 contains weight #7 sand (Table 73). The merged fractures
     exhibit lenses of weight #5 sand, 10 to 20 cm wide in the horizontal direction,
    separated by 10 to 40 cm long fracture segments dominated by weight #7 sand.  In
    some cases the lenses of #5 sand are separated by fracture segments occupied
    exclusively by #7 sand, particularly in the north wall of Trench I (Figure 7.19).
    These observations suggest that the proppant in fracture EL4F2 propagated along a
    lobate front.
    Coalescence of Fractures
    
          Fractures created at equal depths in neighboring boreholes are prone to
    merging. The topmost (first) fractures created at boreholes EL6 and EL7 - EL6F1
                                        199
    

    -------
    TABLE 7.2: FRACTURE THICKNESS
    Fracture
    EL1F1
    EL1F2
    EL2F1
    EL2F2
    EL3F1
    EL3F2
    EWF1
    EWF2
    EMF3
    EL4F4
    EL5F1
    EL5F2
    EL5F3
    EL5F4
    EL5F5
    EL6F1
    EL6F2
    EL6F3
    EL7F1
    EL7F2
    GE1F1
    GE1F2
    GE2F1
    GE2F2
    GE3F1
    Fracture
    Thickness
    (mm)
    11.0
    4.0
    5.0
    2.0
    8.0
    11.0
    8.0
    17.0
    17.0
    16.0
    11.0
    12,0
    10.0
    7.0
    -~— .
    14.0
    13.0
    13.0
    20.0
    10.0
    	
    — — .
    — ..
    — —
    	
    Max Thick.
    Distance
    (m)
    5.0
    1.5
    3.5
    0.5
    3.5
    2.5
    1.0
    2.5
    2.5
    1.5
    2.0
    2.0
    1.5
    0.5
    ......
    2.0
    2.0
    2.5
    2.5
    U
    	
    	
    	
    	
    	
    Borehole
    Uplift
    (mm)
    9.0
    	
    3.0
    3.0
    5.0
    4.0
    13.0
    15.5
    11.5
    25.0
    20.5
    17.0
    7.5
    9:0
    	
    	
    ......
    
    23.1
    _. —
    	
    11.3
    	
    	
    99.5
    Ave
    Notch
    Radius
    (cm)
    15.2
    14.0
    10.2
    10.2
    11.4
    	
    17.8
    17.8
    17.8
    11.4
    15.2
    15.2
    19.1
    25.4
    20.3
    19.1
    12.7
    19.1
    15.2
    17.8
    	
    	
    8.9
    	
    	
    Max
    Notch
    Radius
    (cm)
    17.8
    14.0
    12.7
    11.4
    12.7
    
    22.9
    203
    27.9
    12.7
    19.1
    15.2
    20.3
    38.1
    26.7
    22.9
    15.2
    38.1
    17.8
    40.6
    	
    	
    10.2
    	
    — ..
                                200
    

    -------
    and EL7F1 (Figure 7.2) - may merge laterally to form a large, single fracture. The
    two boreholes are spaced 4.5 m apart.  Similarly, the topmost fractures at boreholes
    ELI, EL2 and ELS - EL1F1, EL2F1 and EL3F1 - merge to form a 12 m long,
    continuous fracture (Figure 73). There is a continuous fracture at the topmost (Fl)
    horizon between boreholes ELI and EL2, which are only 3.25 m apart, but there is a
    3 m gap between the tips of fractures EL1F1 and EL3F1 along the cross section
    connecting the two boreholes, which are 6.25 m apart (Figure 7.3). The two
    fractures, EL1F1 and EL3F1, merge at an unknown distance north of boreholes
    ELI and EL3, indicated by the fact that they vent along the same horizon. The
    topmost fractures at boreholes EL4 and ELS, spaced 4.25  m apart, also merge,
    forming a single fracture (Figure 7.4). The fracture is 10 m long, with major axis
    parallel to the line connecting the boreholes, and with an approximate area of 47
    m2. (Figure 7.4).
    
          The topmost fractures (Fl) were created at approximately the same depth (1
    m) at all seven boreholes. The greatest distance between  two boreholes with a
    continuous fracture at the Fl horizon, along the cross section connecting the
    boreholes, is 4.5 m (EL4F1 and EL5F1 in Figure 7.4). The smallest distance
    between two boreholes (ELI and ELS) with a unfractured gap between the two Fl
    fractures, along the cross section connecting the boreholes, is 6.25 m. Those two
    fractures apparently merged north of the boreholes (EL1F1 and EL3F1 in Figure
    7.3) because fluid vented during the formationof EL3F1 at the same location as
    EL1F1. The smallest distance between two boreholes (ELS and EL7) with no
    apparent merging of the Fl fractures is 6.5 m. This suggests that, for injection of the
    volumes of gel and sand injected into till at the ELDA site, about 6 meters is  the
    maximum distance allowable between boreholes if connectivity of fractures at the
    same horizon is desirable.
    
          Only two cases of a fracture coalescing with an overlying fracture are
    documented by the trench sections. The first is  in the east wall of Trench C (Figure
    7.13), where perlite-filled  fracture EL7F2, created 30 cm below sand-filled fracture
    EL7F1, merges with the overlying fracture at a distance of 2 m north of the
    borehole.  The relatively rapid rise of the perlite-filled fracture probably occurred
    because the density of the perlite slurry was less than that of the sand slurries. The
    only example of a rapidly  rising, sand-filled fracture merging with an overlying
    fracture occurs in the east wall of Trench G (Figure 7.17), where fracture EL4F3,
    created 30 cm below fracture EL4F2, merges with the overlying fracture at a
    distance of 2.5 m north of the borehole. The  concentration of sand in EL4F3 is
    similar to other, more flat-lying fractures (Table 7.3).
    
          Coalescence of several other pairs of fractures, stacked closely together at
    the borehole, is inferred from indirect evidence. The fracture EL2F2 was observed
    to vent from the same horizon as the vent of EL2F1, which was 15 cm above EL2F2
    at the borehole. It seems likely that EL2F2 coalesced at some point with EL2F1,
    even though they appear to be separated in all trench exposures. The same
    evidence indicates that EL3F2 merged with EL3F1.
    
          In all those cases, the fractures coalesced several m away from the borehole.
    Two fractures, EL5F3 and EL5F4, merged very close (<20 cm) to the borehole and
    the two were indistinguishable except in the vicinity of the borehole. The fracture
    EL5F4 was created 7 cm below EL5F3, the closest spacing we attempted. The
                                         201
    

    -------
    TABLE 7.3: PROPPANT CONCENTRATION
    Fracture
    EL1F1
    EL1F2
    EL2F1
    EL2F2
    EL3F1
    EL3F2
    EL4F1
    EL4F2
    EIAF3
    EL4F4
    EL5F1
    EL5F2
    EL5F3
    EL5F4
    EL5F5
    EL6F1
    EL6F2
    EL6F3
    EL7F1
    EL7F2
    GE1F1
    GE1F2
    GE2F1
    GE2F2
    GE3F1
    Sand
    Size
    #7
    #7
    #7
    #7
    #7
    #7
    #7
    #5
    #7
    #7
    #7
    #7
    #7
    #5
    #7
    #7
    #7
    #7
    #7
    PERLTTE
    #7
    #7
    #7
    #7
    #7
    Total
    Sand
    (#)
    300
    200
    200
    500
    200
    200
    200
    100
    300
    250
    300
    300
    400
    300
    300
    600
    600
    600
    1200
    50
    1200
    1200
    200
    1100
    3500
    Total
    Gel
    (gal)
    150
    50
    130
    100
    100
    40
    50
    50
    50
    50
    45
    45
    95
    45
    45
    80
    65
    50
    130
    25
    175
    190
    40
    155
    500
    Injected
    Cone
    (#/gal)
    2.0
    4.0
    1.5
    5.0
    2.0
    5.0
    4.0
    2.0
    6.0
    5.0
    6.7
    6.7
    4.2
    6.7
    6.7
    7.5
    9.2
    12.0
    9.2
    2.0
    6.9
    6.3
    5.0
    7.1
    7.0
                            202
    

    -------
    TABLE 7.4: FLOW PARAMETERS
    Fracture
    EL1F1
    EL1F2
    EL2F1
    EL2F2
    EL3F1
    EL3F2
    EL4F1
    EL4F2
    EL4F3
    EL4F4
    EL5F1
    EL5F2
    EL5F3
    EL5F4
    EL5F5
    EL6F1
    EL6F2
    EL6F3
    EL7F1
    EL7F2
    GE1F1
    GE1F2
    GE2F1
    GE2F2
    GE3F1
    Pad
    Size
    (gal) (
    5 H20
    5 H20
    5 H20
    5 H20
    5 H20
    5 H20
    15 H20
    5 H20
    5 H20
    5 H20
    15 H20
    5 H20
    5 H20
    5 K20
    5 H20
    5 H20 10 GEL
    5 H20 10 GEL
    5 H20 10 GEL
    5 H20 10 GEL
    5 H20
    5 H20 10 GEL
    5 H20 10 GEL
    5 H20 5 GEL
    5 H20 10 GEL
    5 H20 10 GEL
    Flow
    Rate
    gal/min]
    5
    5
    10
    10
    15
    15
    5
    5
    5
    5
    5
    5
    5
    5
    5
    5
    5
    5
    5
    5
    5
    5
    4
    4
    5
    Peak
    Pressure
    I (psi)
    28
    21
    49
    29
    40
    13
    36
    31
    11
    26
    32
    27
    18
    7
    
    24
    35
    46
    44
    26
    38
    85
    29
    115
    85
    Stress
    Intensity
    kPaCcnr5)
    698
    450
    1090
    582
    920
    «•_«•
    1082
    885
    197
    526
    865
    698
    441
    72
    
    662
    824
    1355
    1209
    852
    _____
    _____
    568
    _____
    _____
                                203
    

    -------
    trench cross section (Figure 7.18) does not show this coalescence, because of the
    scale of the cross section.
    
          We conclude that, for the conditions of the ELD A site, vertical fracture
    spacings of 30 cm will typically result in distinct fractures. Vertical spacings of IS
    cm will result in fractures that merge locally but remain distinct elsewhere, whereas
    vertical spacings of 7 cm will result in fractures that rapidly merge. Spacings
    between 15 and 30 cm probably would be a reasonable lower limit.
    Surface Texture of Fractures
    
          Cross sections of the fractures (Figures 7.12-7.19) exhibit few undulations or
    other signs of roughness along the fractures. Borehole EL6 (Figure 7.12) contains
    good examples of both a typical, smooth fracture - EL6F2 - and an exceptionally
    rough, undulose fracture - EL6F3.  There is no significant difference between the
    injected sand concentration of the two fractures, so the difference in surface texture,
    as expressed in the fracture traces, is assumed to be due to inhomogeneities of the
    soil profile.
    
    
    DIMENSIONS OF THE HYDRAULIC FRACTURES
    
          Surface areas of the fractures created at the ELDA site, estimated from
    fracture outlines shown in Figures 72-7A, range from less than one square meter to
    over sixty square meters (Table 7.1). Data shown in Tables 7.1 and 73 reveal no
    statistically significant correlations between surface areas of fractures on one hand,
    and injected proppant volumes or injected sand concentrations on the other hand.
    
          Comparison of the relationship between total injected volume of gel and
          PVI m m mm « *       J»          f        >rw^ IV  W m\. f     1   * ..1   1 * •	1
                            cture surface an
                            o 750 kg of sand
    ,, 	     i by 0.14 square i
    proppant. For lower sand concentrations 400 to 500 kg/nr (4.0^5.0 Ibs/gal), tie
    data indicate a decrease of area by 0.07 square meters per additional 0.004 rrr (gal).
    These data suggest that factors other than proppant volume influence the surface
    area of hydraulic fractures.  A fracture tip screen-out could inhibit propagation,
    thereby limiting their surface area.
    
          Comparison of the relationship between the concentration of sand in the
    injected proppant, measured as pounds of sand per gallon of gel (Table 73), versus
    fracture surface area (Table 7.1), for fractures within narrow ranges of total volume
    of injected proppant, indicates an increase of area by two to seven square meters
    per additional pound of sand per gallon of gel. Assuming that sanding-out at
    fracture tips influences fracture surface area, as suggested above, these data indicate
    that the likelihood of early "sand-out," and, consequently, of small surface area, is
    not directly related to the sand concentration, as intuition might suggest. Clearly,
    more work is needed toward an understanding of hydraulic fracture propagation.
                                      204
    

    -------
    THICKNESS PROFILES OF THE HYDRAULIC FRACTURES
    
          Trenches B through H provide an excellent opportunity to measure the
    thickness of proppant sand along the major axes of fractures (Figures 7.20-7.26).
    Although thickness commonly varies by two to five mm over distances of 20 cm or
    less, third order regression yields asymmetric curves with maximum thickness close
    to the center of the fracture, and hence, northeast of the borehole. Thirteen
    fractures (72%) are skewed away from the borehole, so that the midpoint along the
    fracture trace is located between the point of maximum thickness and the borehole.
    Four fractures (22%) are skewed toward the borehole, so that the point of
    maximum thickness is located between the fracture  midpoint and the borehole
    (EL2F2, EL3F2, EL4F4 and EL6F2). Only one fracture (EL4F1) exhibited a
    symmetrical thickness profile, in which the distance between the fracture midpoint
    and the point of maximum thickness is less than 10% of the distance between the
    fracture midpoint and the edge of the fracture.
    
    
    DIRECTION OF PROPAGATION
    
           Most of the fractures propagated preferentially in the downslope direction,
    as shown in Figures 7.5-7.11. Maps of fracture outlines (Figures 7.2-7.4), as well as
    measurements of major axis lengths and maximum distances from boreholes (Table
    7.1), indicate  that the distance from the borehole to the downslope edge of a
    fracture may be as much as seven times longer than the distance from the borehole
    to the upslope end of the fracture. Clearly, the additional overburden pressure on
    the upslope side prevented fracture propagation under the steep bank bounding the
    southwest edge of the bench (Figures 7.1,7.5-7.11).
    
           Contours of the uplift of ground surface during fracturing, discussed in the
    next section, indicate slightly different orientations of major and minor semi-axes of
    the elliptical fracture outlines defined by uplift contours than the orientations
    indicated by trench cross sections. In both cases, though, the direction of
    propagation is clearly downhill.
    
    
    DISCUSSION
    
           In some respects, hydraulic fractures created during the 1989 tests are
    remarkably similar to the ones created during 1988; in plan, both are elongate with
    aspect ratios of 2:3, and they are highly asymmetric with respect to their parent
    boreholes.  In other respects, fractures created during the two years are much
    different. Those created during 1989 are nearly flat-lying, and none climbed to the
    ground surface to vent. Where they did vent, the fractures grew horizontally until
    they intersected a downwardly sloping ground surface, much like HF13 from the
    1988 tests.
    
           The fractures created in 1989 were slightly smaller in area than those of the
    previous year; however, the sizes of the all the 1988 fractures were limited by
    venting, whereas most of the 1989 fractures did not vent and they could have been
    larger if more fluid was pumped into them. Sand filled all the fractures created in
    1989, whereas it only filled a few during the 1988 tests. The average thickness of
                                        205
    

    -------
                          TRENCH   B
          -3-2-101234567
    ^ ^
     6
     6
     OT
     VI
     QJ
         15
         10
     O
    
    
    H   5
                                                 F2
                      EL 6
     O    -3-2-101234667
     
    -------
                               TRENCH  C
    r  -»
    
    6
    
    6
     CO
     w
     0)
    
     C
    ^}
     o
      -3   -2   -
     0)
     cd
    30
    
    
    25
    
    
    20
    
    
    15
    
    
    10
    
    
     5
    
    
     0
                          EL7
    
                      Trench t
            -3   -2   -1
                      Fracture   Length  (m)
      Figure 751. Graph of fracture thickness (mm) as a function of the distance away
               from the borehole (m), for fractures EL7F1 (top), and EL7F2
               (bottom). Thicknesses were measured along east and west walls of
    
               trench C
                                 207
    

    -------
    r- ->
     6
     g
     W
     OT
     CD
     fi
                              TRENCH D
                                           456
     O
     (D
     O
    12
    10
     8
     6
     4
     2
    F2
                       EL3i
                      Fracture Length  (m)
     Figure 7.22. Graph of fracture thickness (mm) as a function of the distance away
              from the borehole (m), for fractures EL3F1 (top), and EL3F2
              (bottom). Thicknesses were measured along east and west walls of
              trench D.
                               208
    

    -------
                               TRENCH E
     01
     If}
     CD
     PI
    X
     O
    -2-1012345678
     O
     (0
                                                            8
                       Fracture Length (m)
     Figure 723. Graph of fracture thickness (mm) as a function of the distance away
              from the borehole (m), for fractures EL1F1 (top), and EL1F2
              (bottom). Thicknesses were measured along east and west walls of
              trench E.
                                 209
    

    -------
    r  -^
     6
     a
     W
     w
     Q)
     CJ
    44
     O
    £-•
     0)
     JH
     3
    -*J
     o
     CO
     JH
          6
    6
                              TRENCH F
                                                    Fl
                   EL 2
                    Trench
                      A
     -2     -1
                      Fracture Length  (m)
     Figure 7.24. Graph of fracture thickness (mm) as a function of the distance away
              from the borehole (m), for fractures EL2F1 (top), and EL2F2
              (bottom). Thicknesses were measured along east and west walls of
              trench F.
                                 210
    

    -------
                    Fracture Length (m)
    
    Figure 7.25. Graph of fracture thickness (mm) as a function of the distance away
              from the borehole (m), for fractures EMF1, EL4F2, EL4F3 and
              EL4F4. Thicknesses were measured along east and west walls of
              trench G.
                                  211
    

    -------
                         TRENCH H
        10
    
         8
    
         6
    
         4
    
         2
                       F3
    ELS
      Trench
        A
         -3      -2-10       1
    
                     Fracture Length (m)
    Figure 726. Graph of fracture thickness (mm) as a function of the distance away
              from the borehole (m), for fractures EL5F1 (top), EL5F2 (middle)
              and EL5F3 (bottom). Thicknesses were measured along east and
              west walls of trench H.
                                  212
    

    -------
    sand in the 1989 fractures (11 mm) exceeded the maximum thickness (9 mm) from
    the previous year.
    
          Hydraulic fractures created during the 1989 test increased the steady-state
    rate of inflow into boreholes in unsaturated ground (data are in Section 1) by factors
    between 3.1 and 9.0. That amount of increase in flow rate could significantly
    improve the rate of remediation of a contaminated site. There are certainly aspects
    of the technology that can be improved - most of which we are currently pursuing -
    but the technology used during the 1989 tests appears to be capable of providing
    important improvements in remediation. We conclude that the technique of
    hydraulic fracturing is ready to be tested during remediation.
                                      213
    

    -------
                                 SECTION EIGHT
    
                     MONITORING HYDRAULIC FRACTURES
    
          Monitoring the geometry of hydraulic fractures created to improve
    remediation should be done for two purposes. On one hand, the performance of a
    fracture in delivery or recovery is sensitive to its geometry, so monitoring will yield
    information that will improve the design of delivery or recovery systems.  On the
    other hand, some sites will require that hydraulic fractures do not interfere with
    features, such as sewers, foundations, or electric cables, so monitoring will yield
    information that could suppress the growth of errant fractures.
    
          A wide range of methods is available to predict and monitor the geometry of
    hydraulic fractures. The majority of them were developed within the petroleum
    industry and they are typically used where hydraulic fractures cut rock at depths of
    hundreds or thousands of meters; the conditions typical of oil wells. To our
    knowledge, none of those methods has been evaluated where hydraulic fractures cut
    soils at depths of a few to several tens of meters; the conditions typical of many
    remedial applications.
    
          During field tests described in previous chapters we examined four
    techniques of monitoring processes related to hydraulic fracturing.  Those
    techniques include measuring 1) the pressure of fracturing fluid; 2) tilting of the
    ground surface over a fracture; 3) uplift of the ground surface; 4) electrical
    resistivity of ground containing a hydraulic fracture. The objectives of our study
    were to examine the level of effort required to implement and reduce data from
    each technique, and to evaluate the potential that it could have in future
    applications.
    
    
    INJECTION PRESSURE
    
          The pressure of fracturing fluid varies with time during creation of a
    hydraulic fracture, and the pressure record is a basic tool of monitoring. The form
    of the pressure record can be used to infer when propagation starts, as well as
    various aspects of fracture geometry.  Problems such as plugging of the fracture with
    sand result in large pressure surges that are readily identified on the pressure
    record.
    
    Method
    
          Injection pressure was recorded during a few of the tests in 1988 and all of
    the tests in 1989.  Pressure was measured at the ground surface using a Druck
    transducer interfaced with a data acquisition system and lap-top computer.
    Measurements were taken every second and both written to a disk file and displayed
    as a plot on the video screen. The pressure records for the tests are shown in Figure
    8.1.
    
    Results
    
           Most of the records obtained during the 1989 tests are similar in general
    form, and consist of the following four basic periods:
                                      214
    

    -------
          EL1F1
       30.0
       24.0
    n  MA
    
    
    
    OJ
    en
    in
    
    -------
          EL2F1
    60.0
    to.o
    »
    t-l
    n
    n
     ZS.O
    OH
    10.0
    HA
    ~ ( 1 I
    •
    
    
    
    
    -1 1 1
    
    -
    -
    •fk
    •3
    I ^
    -Ar-^_ M -
    aj-1*!" 1 i i i 1 i i A'H i i 1 i t i 1 i r i 1 i L
         0.0     2.0     4JO     tJ>     B.O    10.0    12.0     l*.0
    
    
                            time (min)
          EL2F2
       M-O
    n
    in
    v
    
    
    P4 ""t
                             |  i—i  i j T—I r-j--!—r  i 7
    I  "I
    
    I  q
                                     i i  i  [ r  i i  I  i i  i
         0-0     U     4J)     IJ    BJ)     10J    lifl
                            time  (min)
     Figure 8.1, continued.  Pressure records from fracturing tests during 1989.
                                      216
    

    -------
          EL3F1
                                I
                                       V
                           e.o     (.0     10.0
                           time (min)
                                              12.0     14.0
         EL3F2
                          time  (min)
    Figure 8.1, continued. Pressure records from fracturing tests during 1989.
                                  217
    

    -------
          EL4F1
     0)
     «-  200
     w
     0)
           i  i i  t i  1*1 iiiiiiiiiiitiiiiiiiiiiii
               z.o    «.o
                                 tio    WLO
                            time (mln)
                                                  14.0    it.c
          EL4F2
    ra
    O.
    tQ
    u
       1E.D
            tt i  I i  i  i 1
         to     u>
                           fill L
                                  I ,  1 ,  I ,  1 ,  t
                            •J     U     1OO
    
                            time  (min)
    Figure 8.1, continued. Pressure records from fracturing tests during 1989.
                                    218
    

    -------
           EL4F3
        1ZJ>
                            10.0        16.0
    
    
    
                            time (min)
          EL4F4
        25.0 -
         0.0     to     4.0     e.o
                                          10.0     12.0    14.C
                           time  (min)
    Figure 8.1, continued.  Pressure records from fracturing tests during 1989.
                                    219
    

    -------
          EL5F1
       88.0
     n
     a,
     v
     u
     n.
                     I . I  I	I	1
         0.0
                    B.O
                               10.0
                           time (min)
         EL5F2
                          time (min)
    Figure 8.1, continued.  Pressure records from fracturing tests during 1989.
                                   220
    

    -------
          EL5F3
         tO   2.0
                           time (min)
          EL5F4
                          10    U    100    liO
                          time (min)
    Figure 8.1, continued. Pressure records from fracturing tests during 1989.
                                  221
    

    -------
          EL6F1
       110.0
    ra   TO.O
    en
    n
    0)
                    1(IiII
                                     '  • '  i
    m.lt i r i  I i  t i
                                                 u
                                                 wl r  i i  i
                                         4.0
                                                 fl.0
                            time (min)
          EL6F1A
        28.0
                           ICLO      U.O
    
    
                           time (min)
                                            taj)
                                                     2flJ
    Figure 8.1, continued. Pressure records from fracturing tests during 1989.
                                    222
    

    -------
           EL6F2
      n
      a
      3 no -
      w
      0)
                  S.D
                          10.0      18.0      200
    
    
                            time (tnin)
                                                  29.0
          EL6F3
                                   IB.O
    
    
                            time  (min)
                                            80.C
                                                     23.0
    Figure 8.1, continued. Pressure records from fracturing tests during 1989.
                                    223
    

    -------
         EL7F1
         . i  i i  i  ; i  i  i i
    to
    CO  30.0
    J-!
    P,
        0.0
         0.0
                                                       f  -I
                                  1  •  ••-•'-  I •  •  • •  I  ' '
                         20.0      30.0
    
                            time  (min)
         EL7F2
       ZBJJ
    en
    CO
    a
       «•
                                         i  i  i  1  i  i
    fit
    
     o-
    
     l
                 I  i  i  i  t  i i  i  I  .1  i  i  I i  i  i  t  t  i  i.
                2J      10      14      U      1IU      124
                            time (min)
    Figure 8.1, continued. Pressure records from fracturing tests during 1989.
                                    224
    

    -------
            GE1F1
          90.0
          20.0
    3
    CO
    g   10.0
    
    
    O.
                -
                          a-
                          «
                         : rt-
        -10.D
          0.0
                                                      ?;
                                                       ir
                                                          \
                                             J?    *{,'
                                             ,5 s   (*2  «'
                  l°-°     ZO.O     30.0      «0.0     SO.O
    
                             time (min)
           GE1F2A
     3
     CO
     (0
     V
     L,
                            time (rain)
    Figure 8.1, continued. Pressure records from fracturing tests during 1989.
                                   225
    

    -------
           GE1F2B
        ISOJ
     g  ""
     
    -------
            GE2F1
         300 -
          0.0    2.0    4.0    6.0    8.0   10.0    12.0   110   16.0    18.0
           GE2F2
        120.0
    OJ
    t-l
    3
    m
    m
    0)
          OLD     10J)     Z04>     SOU)    4041    60.0
    Figure 8.1, continued. Pressure records from fracturing tests during 1989.
                                        227
    

    -------
        Period I: Pressure increases rapidly for several seconds and reaches a peak
            between 70 and 275 kPa (10 and 40 psi)1.
    
        Period II: Pressure decreases over one to several minutes and reaches a
            minimum between 20 and 70 kPa (3 and 10 psi).
    
        Period HI: Pressure increases to between 35 and 70 kPa (5 and 10 psi), and then
            decreases gradually.  This is the longest of the periods and lasts from
            several minutes to more than an hour, depending on the volume of fluid
            pumped into the ground.
    
        Period IV: Pressure decreases abruptly and is followed by a gradual flattening
            of the slope.
    
           The fracturing periods are associated with various events during the
    fracturing procedure. Typically, a change in flow rate causes an instantaneous
    change in pressure, whereas a change in fluid properties results in a change in
    pressure roughly 30 seconds later. This lag is the same length of time it took fluid to
    traverse the injection hose and reach the transducer.
    
           Inflation of the pumping system prior to fracturing occurs during Period I.
    We infer that the abrupt decrease in pressure marking the beginning of Period n
    marks the onset of fracture propagation. The increase in pressure marking the
    beginning of Period III typically occurs roughly 30 seconds after we began to pump
    gefand sand, and it is due to the increase in viscosity of that fluid compared with
    fluid that was injected initially. As the fracture increases in size, we expect a gentle
    decrease in pressure and this  behavior is seen during Period III of most tests.
    Pressure fluctuations within Period III are typically associated with changing from
    one batch of gel to another. Increasing the sand concentration, for example, is
    commonly followed by an increase in pressure within Period in (e.g. EL6F3). Other
    fluctuations in pressure during Period HI are associated with minor adjustments in
    pumping rate. The decrease in pressure during the beginning of Period IV is
    associated with a reduction in fluid viscosity when we'changed from pumping sand-
    laden gel to water at the end of a test.  A sharp drop in pressure, marked by a nearly
    vertical slope on the record, corresponds to the time when the pump stopped.
    Pressures were not recorded after the pump was turned off, so we are unable to
    evaluate the methods of Nolte and Smith (1981), who use the record of pressure
    after shut-in to determine leakoff parameters.
    
          The pressure record can be used to diagnose problems associated with the
    fracturing equipment  When  the fracture is growing during Period HI, a major
    increase in pressure (EL6F1,  GE1F2A, GE1F2B) typically indicated a blockage
    downstream of the transducer, whereas a major decrease (e.g. EL54F4) indicates an
    upstream blockage.  Downstream blockages were caused when plugs of sand formed
    at the entrance to the fracture. This problem is caused either by excessive early
    leakoff, or by an early concentration of sand that was too great. Therefore, when
    the pressure record indicated a downstream blockage, we recut the starter notch,
    increased the volume and viscosity of the pad, decreased the initial sand
    1 Cited pressures are for tests at the ELDA site.  Pressures at the Goettle site were
        typically greater, as seen on the records.
                                        228
    

    -------
    concentration and tried again to create the fracture. This strategy was successful.
    Unexpected pressure drops indicating upstream blockages were caused, at least
    during our tests, by clogging of the pump usually because of inadequate crosslinking.
    
           Venting of the fracture to the ground surface occurred during the creation of
    five fractures (EL1F1, EL2F1, EL2F2, EL3F1, EL3F2), and in most cases venting
    has negligible effect on the form of the pressure record. One explanation is that the
    pressure of the injection fluid at the tip of the fracture is essentially zero, so that
    pressure at the tip would be unaffected by venting.  All of those fractures, except
    EL1F1, merged with another earlier fracture, but we were unable to identify a
    feature of the records that indicates merging.
    
           The form of the pressure record, after removing fluctuations due to changes
    in fluid properties or pumping rate, is related to the geometry of the fracture and
    rate of leakoff.  Methods are currently available for inferring the geometry of
    vertical fractures from the forms of records during Period III (Nolle and Smith, 1981
    and 1987). Some  of those methods are applicable to shallow, flat-lying, asymmetric
    fractures such as the ones we created, but most of the methods of interpretation
    have been developed for oil-field applications where fractures are vertical.
    
           Leakoff characteristics can be determined from pressure records made
    during pumping and after the pump has been shut off (Nolle,  1979; Nolte and
    Smith, 1987). Records made during this work were terminated too early to make
    reliable measurements of leakoff characteristics, however. Future work should
    focus on obtaining the leakoff characteristics, and using that information to
    anticipate tip screen-out (Smith and others,  1987; Nolte, 1984).
    
    
    SURFACE TILT
    
          The ground surface overlying a hydraulic fracture is deformed as the fracture
    inflates. Characteristics of the fracture can be inferred by measuring the
    deformation and then using various theoretical analyses.  In this study, deformation
    was determined either by measuring changes in slope of the ground surface with
    tiltmeters, or by measuring changes in altitude with a leveling device. The work
    with tiltmeters was conducted during the 1988 tests  and is described in the following
    section.
    Method
    
          Surface tilt was measured as a function of time using Model 722 tiltmeters
    rented from Applied Geomechanics Inc., Santa Cruz, California. That model is
    cylindrical in shape and designed to be placed below the ground surface in a boring.
    It is accurate to within 0.1 microradian.
                                       229
    

    -------
          The tiltmeters were installed slightly below the ground surface in vertical
    boreholes.  Sand was placed around the tiltmeters and the orientation of their major
    axis was adjusted to vertical.  Signals from the tiltmeters were conditioned using
    electronic equipment rented from Applied Geomechanics Inc. and then recorded on
    our data acquisition system.
    
          A total of eight tiltmeters were installed around the cluster of boreholes 4,5,
    6, and 7 (Fig. 8.2).  Three of those boreholes are at the vertices of an equilateral
    triangle, and the other is at the center of the triangle. Six of the tiltmeters were
    arranged in a hexagonal pattern around the three boreholes, and the other two
    tiltmeters were near the central borehole (Fig. 8.2).  The arrangement placed three
    tiltmeters at roughly equal distances from each borehole, and placed the others at
    greater distances. A single layout of tiltmeters was necessary because installation
    required several hours and we were unable to move  them on the day of the 1988
    test. One of the tiltmeters, tiltmeter F, yielded erratic data and was omitted from
    the interpretations.
    Results
           Creating hydraulic fractures caused tilts of as much as 2000 microradians
    (Appendix A), a strong signal for the instruments we used. Tilt signals varied widely
    as functions of time and location (Fig. 8.3), and a meaningful interpretation of this
    information is only possible after data reduction. Qualitative information about the
    location and orientation of the fracture can be obtained by plotting tilts as vectors in
    plan view. Suchplots can be created in real time and interpreted as the fracture is
    being created. This procedure was not attempted during field tests.
    
           Tilt data can also be used to obtain quantitative estimates of fracture
    geometry. This procedure involves inverting the tilt measurements using a
    mathematical solution for surface displacement over a fracture at depth.  The
    solution used in this work assumes that the fracture is planar and shaped  like a
    rectangle. A total of eight parameters, which specify tne size, location, aspect ratio,
    and aperture of the fracture, are obtained from a nonlinear inversion of that
    solution (Davis, 1983).  Engineers at Applied Geomechanics Inc. developed the
    inversion technique, and analyzed the data from the tiltmeters. They estimated the
    geometry of each fracture at several times during its creation, and their results are
    compared with maps of the fractures in Figures 8.4 through 8.6.
    
           Each of Figures 8.4 through 8.6 consists of a series of two or three contour
    maps illustrating the change of fracture geometry during the fracturing process.
    Figure 8.4 shows three  stages in the development of fracture HF5 (created from
    borehole #5); Figure 8.5 shows three stages in the development of fracture HF6
    (borehole #6); and Figure 8.6 shows two stages in the development of fracture HF7
    (borehole #7).
    
           Table 8.1 compares orientations (strikes and dips) of fractures determined
    from tiltmeter data to those determined from excavation of the fractures. Following
    the convention used by Applied Geomechanics Inc., the subcontractor responsible
    for tiltmeter data, the direction of fracture dip represents a clockwise, 90° rotation
    from the azimuth of the fracture strike. Coordinates of the center of the fracture
    relative to the borehole, calculated from tiltmeter data, along with fracture aperture
                                         230
    

    -------
           CD
                   B
                   ©
                    #6
                   l.69m
         G
         CD
    
     #4©
    3.84m
        ©
        H
                          Borehole
                                   1. 83m
          0      1
           meter
                            F   Tlltmeter
                            ©« -- '
                   C
                   ©
    #5
    1.04m
    
       ©
                                                E
                                               ©
       Boreholes and Tlltmeters,   ELDA  Landfill
    Figure 8.2. TDtmeter array from 1988 tests.
                               231
    

    -------
       788. .-r
      "79B-.B
      -1988
          .8BB
                              (*to>
                                                   xE a
    Figure 8.3. Tilt as a function of time.
                                        232
    

    -------
      2298 -T"•XC--B
      1698 -••
      189B
      40B.
     -2BB.
     -888.
     -14BB
                                                  xE 0
      693. -r
     -BBB.
          .aaa
                  SB.B188.13B.20B.
                         Tin* <«to)
    Figure 8.3, continued.  Tilt as a function of time.
                                        233
    

    -------
     (thickness), are also given in Table 8.1. Coordinates are measured in ft, with the
     positive* direction pointing eastward, and the positivey direction pointing
     northward.  Fracture aperture is measured in inches.
     TABLE: 8.1. ACTUAL AND PREDICTED ORIENTATIONS OF FRACTURES
     Era    Actual Strike        Tilt-meter       Actual Dip        Tilt-meter
                                   Strike                             Dip
     HF5a         80                98             22                 8
     HF5b         84               143             26                37
     HF5c         40               359             17                37
     HF6a       359               332              9                11
     HF6b       359               334              9                23
     HF6c       349                 4             14                29
     HF7a       228               242             27                 1
     HFTb       228               213             27                27
    Borehole 5
    
           Inversion of data from borehole 5 indicates that a very thin (0.03 cm; 0.01 in),
    nearly horizontal fracture, striking 98° and dipping 8° SW, formed 1 minute after
    fracturing commenced (Figure 8.4a).  The "tiltmeter strike" is fairly close to the
    strike based on excavation (80°), although the "tiltmeter dip" is considerably
    shallower than the "excavation dip" (23°). One minute later (Figure 8.4b), tiltmeter
    data indicate a somewhat thicker fracture (0.03 cm; 0.01 in), with strike rotated
    clockwise to 143°, and dip increased to 37° SW. Although the tiltmeter strike at this
    stage is less consistent with the excavated strike than it was a minute earlier (Figure
    8.4a), the tiltmeter dip is closer to the excavated dip. Tiltmeter data shown in
    Figures 8.4a and 8.4b suggest that the smaller, eastern lobe of the fracture formed
    during the first two minutes of fracturing. Tiltmeter data for the final phase of
    fracturing (Figure 8.4c), 3 minutes and 20 seconds after fracturing commenced,
    indicate Formation of a fracture twice as thick (0.61 cm; 0.24 in) as a minute earlier,
    striking 359° and dipping 37° NE. Fracture propagation during this phase occurred
    several ft west of the borehole. The location and orientation of the fracture based
    on tiltmeter data strongly suggests that the larger, western lobe of the fracture (as
    mapped from excavation) formed late during the fracturing process (after 3
    minutes).
                                       234
    

    -------
              HF5
                                     a
                                               0   1   Z
                                                meters
              HF5
    Z.O'O.E
                                              0   1   Z
                                               met era
              HF5
                                              0   1   Z
                                               Betera
    Figure 8.4. Hydraulic fracture HF5 and interpretation of surface tilts at three
               times during the test.
                                    235
    

    -------
           The fracture inferred from early tiltmeter data at borehole 5 is inconsistent
    with excavation data, but the later two fractures are consistent with field
    observations. The second frame in the interpretation indicates that the eastern lobe
    of HF5 formed first and was followed by growth of the western lobe. Apparently,
    propagation of the eastern lobe was arrested before it could vent. Continued
    injection resulted in growth of the western lobe, which vented in two locations west
    and northwest of the borehole. The tiltmeter inversion underestimates the area of
    the fracture, but overestimates its dip, based on excavation of the fractures.
    
    
    Borehole 6
    
           Comparison of tiltmeter inversion data with the fracture map based on
    excavation around borehole 6, shown in Figure 8.5, indicates that comparatively
    steeper-dipping (23°) but still NW-striking (334°) portions of the fracture formed up
    to 2 m away from the borehole, as late as 7 minutes and 40 seconds after fracturing
    commenced (Figure 8.5b). Finally, distal portions of the fracture, dipping as much
    as 29° SE and striking NNE (4°), formed over 8 minutes after fracturing
    commenced.
    
          Tiltmeter and excavation data for fracture HF6 clearly illustrate the ability to
    trace the dynamic development of a fracture using tiltmeter inversion data.
    
    
    Borehole 7
    
          Development of a relatively small fracture is documented in Figure 8.6,
    showing tiltmeter inversion data and excavation data for fracture HF7,2 minutes
    (Figure 8.6a) and 3 minutes (Figure 8.6b) after commencement of fracturing.
    Comparison of the first frame (Figure 8.6a) with other frames representing 2
    minutes after commencement of fracturing (Figures 8.4b and 8.5a) reveals
    comparatively low fracture dip (1° NW). Tiltmeter-generated strike and dip of the
    fracture 3 minutes after commencement, shown in Figure 8.6b, is in fairly good
    agreement with excavation data (Table 8.1).
    
    
    Discussion
    
          Propagation direction and strike are predicted within roughly 45° and dip is
    generally within 15°-mostly dip is overestimated. Sizes of fractures are generally
    underestimated, except in the case of HF7 where it is markedly overestimated. The
    general locations of the fractures is predicted with reasonable accuracy,  however.
    
          The results of the inversions of tiltmeter data are remarkably consistent with
    field mapping, considering that the inversion scheme must reconcile the tilt field
    resulting from a fracture whose geometry is much more complex than the simple
    rectangular model it uses. The inversions of tilt data taken during the 1989 tests,
    where the fracture forms were quite simple, certainly would have yielded more
    accurate results.  Moreover, some of the error in the tilt data results from
                                         236
    

    -------
    HF6
    a.
    HF6
    b.
                                                        0  1   t
                                                         •at era
    HF6
                                                                      •0.6
                                                                   0  1   £
                                                                    metera
              Figure 8.5.  Hydraulic fracture HF6 and interpretation of surface tilts at three
                         times during the test.
    

    -------
    oo
                     HF7
    a.
                                                  0   t    Z
                                                    meters
                                                           X
    HF7
                                        0    1    Z
                                          meters
                            Figure 8.6. Hydraulic fracture HF7 and interpretation of surface tilts at three
                                       times during the test.
    

    -------
    placement of the tiltmeters. For example, the inversion of data from HF7
    overestimates the extent of the southern boundary of the fracture, but there are no
    tilt data from the southern side so the inversion must resort to extrapolation.
    
          Inversion of tiltmeter data should provide estimates of location during
    propagation. Application of this technique to real-time monitoring — inversion in
    the field during fracturing - should provide a safeguard against the growth of errant
    fractures, and it may allow for small adjustments in the fracturing process to
    optimize fracture performance.
    
    
    SURFACE UPLIFT
    
    
    Design
    
          Standard surveying techniques were employed to measure surface uplift
    incurred during fracturing. Elevations of points distributed radially around each
    borehole were measured before and after individual fracturing episodes. Values of
    the difference between elevations before and after fracturing, measured in mm at
    each surveying point, are contoured for several fractures in Figures 8.7 • 8.10.
    
          Superimposed on the uplift contours are contours representing the thickness
    of the fracture associated with the measured  uplift. Thicknesses of fractures were
    measured directly on trench walls exposing traces of the fractures.
    
          Uplift contours shown in Figures 8.7 and 8.10 represent total uplift incurred
    during the creation of a chronological series of fractures.  Accordingly, the
    superimposed thickness contours indicate the combined thickness of all fractures
    involved in the uplift.
    
    
    Results
    
          Surface uplift and fracture thickness contours combined for fractures EL6F2
    and EL6F3 are shown in Figure 8.7. Contours for fractures EL7F1  and EL7F2 are
    shown in Figures 8.8 and 8.9, respectively.  Combined contours for fractures EL5F1
    through EL5F4 are shown in Figure 8.10. Fracture thickness contours are bold, and
    surface uplift contours are medium weight in the figures.
    
          Several characteristic features were used to compare the uplift and thickness
    contours for each fracture or group of fractures.  These include the strike of the
    major axis of the ellipse roughly defined by the contours; the ratio of major to minor
    axis (aspect ratio); the maximum values of uplift and thickness; the distances from
    the borehole to the points of maximum value (the tops of the domes defined by the
    contour sets); and the distances from the points of maximum value to the points of
    zero uplift or thickness, measured along the major axes.
                                      239
    

    -------
    EL6FE-F3
        •- C \
                                     B  -.
    Uplift Contour
    
    Thickness Contour
    
    Trench Out 11ne
                                                   —  —  Uplift  Grid
        Figure 8.7. Thickness and uplift contours, combined for fractures EL6F2 and
                  EL6F3.
    

    -------
    EL7F1
           C \
    Uplift Contour
    Thickness Contour
    Trench Outline
    Uplift Grid
                                                          0	j
                                                          meter
                                                      EL6
              Figure 8.8. Thickness and uplift contours for fracture EL7F1.
    

    -------
    Borehole 6
    
          Total surface uplift and total fracture thickness associated with fracturing
    episodes EL6F2 and EL6F3 are shown in Figure 8.7.  Both contour sets are roughly
    elliptical in shape, with aspect ratios of about 1.5 to 1.6. The major axis of uplift
    contours generally trends to the northwest, whereas that of thickness contours
    trends roughly to the north. The distance between the maximum and zero values
    along the major axes is about 5 to 5.5 m in both contour sets. The distances between
    the borehole and the points of maximum value are 3.5 m for the uplift contours, and
    2.5 m for the thickness contours. Maximum uplift is 20 mm and maximum thickness
    is 14 mm, indicating 30% deflation.
    
          It appears that the general shape and size of the contour domes defined by
    uplift and thickness contours for fractures EL6F2 and EL6F3 are roughly
    equivalent, whereas the major axis trends differ by about 45°. The major axis trend
    for the thickness contours coincides more closely with the downslope direction than
    that for the uplift contours. The downslope direction is approximately parallel to
    the trends of trenches B and C.
    
          A noteworthy feature of the surface deformation is the zone of subsidence
    beyond the zone of uplift. Subsidence appears to have occurred along or beyond the
    outer margins of the fractures, a result predicted by theoretical analyses (Pollard
    and Holzhausen, 1979).
    Borehole 7
    
          Uplift and thickness contours are plotted separately for two fractures created
    at borehole 7 - EL7F1 and EL7F2. Shapes and degrees of consistency between
    uplift and thickness data are quite different for the two fractures.
    
          Contours of fracture EL7F1, shown in Figure 8.8, are fairly circular, with
    radial distances of approximately 3.5 meters between the maximum value (center of
    the circle) to values of zero, in both the uplift and thickness contour sets. The
    maximum uplift is 30 mm, and maximum thickness is about half of maximum uplift,
    indicating 50% deflation. Distances from the borehole to points of maximum uplift
    and thickness are 1.5 m and 2 m, respectively. Subsidence was observed on two of
    the lines of data, but the other lines recorded only uplift (Fig. 8.8).
    
          The two contour sets reflect good general agreement over the shape and
    areal extent  of fracture EL7F1.  Doubling of surface uplift, measured immediately
    after creation of the fracture, compared to fracture thickness reflects significant
    settling, or deflation, of the fracture and its overburden following the creation of the
    fracture.
    
          The main difference between the contour sets lies in the orientations of the
    elliptical, inner contours of both sets. Inner contours of uplift trend northeastward,
    whereas those of thickness trend northward. These trends reflect bias due to the
    distribution  of data points, which is limited to two orthogonal axes for uplift, and the
    trench walls for thickness.
                                         242
    

    -------
           The surface deformation over EL7F2 (Figure 8.9) is unusual; there is a zone
     of subsidence containing the borehole and the maximum uplift (5 mm) is located
     about 5 m north of the borehole. The thickest part of EL7F2 is 5 m north of the
     borehole, and that location is coincident with the point of maximum uplift.  The
     basin-like zone of subsidence, however, is difficult to explain. The fracture EL7F2
     dips slightly 5° to the south, and Pollard and Holzhausen (1979) show that a dipping
     fracture will produce an asymmetric deformation field. The sense of asymmetry
     predicted theoretically, however, is opposite that which is observed; it predicts a
     zone of subsidence centered north of the well. The zone of subsidence is difficult to
     explain.
    
           Differences in the trends of major axes of the two contour sets are that same
     as those exhibited by contours of the overlying fracture (EL7F1).  The differences
     are due to the distribution of data points, which was dictated by trench location for
     thickness contours, and by the surveying grid configuration for uplift contours. Time
     constraints prevented the use of more surveying data points.
    
    
     Borehole 5
    
           The clearest correlation between uplift and thickness contours is exhibited by
     the combined contours for fractures EL5F1, EL5F2, EL5F3, and EL5F4, shown in
     Figure 8.10. Both contour sets are roughly circular, with maxima (center of circle)
     at or near the borehole. Maximum uplift, located approximately 1 m east of the
     borehole, is almost 30 mm, and  maximum thickness, located at the borehole, is 20
     mm, so thickness is 33% less than uplift.  The average distance from the maximum
     value to the zero contour is 4 m for uplift, and 3 m for thickness, so the area
     enclosed by the zero uplift contour is 33% larger than the area enclosed by the zero
     thickness contour.
    
           The shapes of the uplift and thickness contour sets are similar.  Both contour
     sets exhibit a conspicuous, localized disturbance in the northeastern hemisphere,
     about midway between the borehole and the periphery (zero contour). The
     irregularity seems to reflect a discontinuity in one of the fractures. This is not
     surprising, since the trace of fracture EL5F1, shown in Figure 7.18, is markedly
     discontinuous.
    
           The areal distributions of the two contour sets, however, are different, as
     manifested by the distances from the borehole to the maxima, and by the areas
     enclosed within the zero contours. Compared to the area underlain by fractures, the
     area of surface uplift is 33% larger, and shifted approximately 2 m eastward, or
     downslope  (see Figure 7.1 for topography).
    
    
     ELECTRICAL RESISTIVITY
    
           Electrical resistivity is another method of sensing hydraulic fractures  without
     disturbing the subsurface. Adding an electrolyte to the fracturing fluid will result in
     a highly conductive fracture that is embedded in poorly conductive till. Three
    geophysical techniques of monitoring fracture location by sensing the conductivity
    contrast were employed at the ELDA site during the 1988 fracturing test. These
                                     243
    

    -------
    EL7FE
           .  C
    .  B
    Uplift Contour
    Thickness Contour
    Trench Out 11ne
    Uplift Grid
                                                            0	1
                                                            meter
                                                        EL6
               Figure 8.9. Thickness and uplift contours for fracture EL7F2.
    

    -------
    EL5F1-F4
                                                            Uplift  Contour
    
    
                                                            Thickness  Contour
    
    
                                                        	  Trenoh  Out line  .
    
    
                                                        —  Uplift  Grid
                                                        0	1
                                                        I	1
                                                        meter
            Figure 8.10. Thickness and uplift contours, combined for fractures EL5F1,
                     EL5F2, EL5F3 and EL5F4.
    

    -------
    were, the mise-a-la-masse, Wenner, and dipole-dipole methods. The mise-a-la-
    masse method was the most successful, yielding encouraging results for static and
    dynamic tests during the 1988 field test. It revealed fields of well-developed changes
    in apparent resistivity following fracturing. In addition, mise-a-la-masse showed
    abrupt changes in apparent resistivity as a function of time during a test, probably
    indicating the fracture had reached a critical location with respect to the sensing
    electrodes. The resistivity measurements were made in collaboration with Dr.
    Donald Stierman, who is a specialist in applying geophysical techniques to waste
    sites (Stierman, 1988). The following two sections are based on Stierman's findings.
    
    
    Design
    
           All resistivity measurements were made using a SoilTest R-60 resistivity
    system and a digital ammeter. The R-60 transmitter consists of 6 batteries wired
    such that output voltages can be varied between 45 and 840 volts in 6 dial-selected
    steps. Porous pots containing saturated copper sulfate solution as electrolyte served
    as electrodes. Distances were measured with a fiberglass surveyor's tape.
    
           Slurry injected into the hydraulic fractures was spiked with about 3% KC1 to
    ensure strong contrast between the electrical resistivities of the fluid and the
    formation. Samples collected were tested in the laboratory and found to exhibit
    electrical resistivities of about 0.3 ohm-meters at 25° C, consistent with a
    concentration of between 20 and 30 ppb of total dissolved solids. The fluid was
    about 120 to 400 times more conductive (less resistive) than the till in which the new
    fractures were developed.
    
           Electromagnetic ground conductivity measurements were also made using a
    Geonics EM34-3XL in the horizontal dipole mode. Results were convened to
    apparent resistivity so that they could be compared directly with DC resistivity
    measurements.
    
           The mise-a-la-masse method uses a conductive rock or soil body as one
    current electrode, with a second current electrode set at some distance away.  A
    roving potential electrode maps the electrical field resulting from injecting current
    into the rock or soil body. If the ground is homogeneous, the current electrode acts
    as an electrical monopole at the center of a series of concentric circular
    equipotential contours. This field is distorted in the presence of subsurface
    conductors and the equipotentials tend to outline any buried conductive structures.
    This technique has been applied to mapping leaks from waste disposal lagoons
    (Stierman, 1988).
    
           During the 1988 ELDA fracturing test, one current electrode was planted
    over 200 m to the east, effectively at infinity over the 20-m diameter array.
    Boreholes H4, H5, H6, or H7 served as the second current electrode. Potential was
    measured with respect to a porous pot planted over 200 m to the west.  Colinear
    electrodes were planted along lines connecting boreholes. The total number of
    electrodes is 18.
                                     246
    

    -------
     Results
    
           Following the creation of fracture H4, mise-a-la-masse measurements were
     made for boreholes 4 and 6. Results were dramatic at borehole 4, shown in Figure
     8.11, which is a contour map of apparent resistivity at borehole  4 after fracture H4.
     Results were unimpressive at borehole 6 during creation of fracture H4. Post-
     fracture measurements show resistivity decreased 10 to 15% for electrodes nearest
     borehole 4, but increased slightly for borehole 6 (Figure 8.12).  There is a strong
     correlation between the resistivity decrease and distance from borehole 4, with an
     elongation of this trend southeastward, toward borehole 5. This elongation suggests
     that fracture H4 propagated preferentially toward the southeast. The generally
     elliptical shape of the resistivity change contours suggests a largely horizontal
     fracture.
    
           Boreholes 5,6 and 7 were shallower than borehole 4, and the shallower
     fractures created at these boreholes were expected to obscure deeper fractures.
     Due to time constraints, results shown in Figure 8.13 represent the composite effect
     of fractures H4 and H5. Again, there is a clear correlation between a decrease in
     resistivity and the distance from borehole 5 for mise-a-la-masse using borehole 5 as
     a current electrode. Resistivity on borehole 4 following fracture 5 (Figure 8.14)
     shows a resistivity increase at most electrodes. The "bulls-eye" patterns of opposite
     polarities suggests a horizontal fracture extending some 3 to 5 m in each direction
     from the borehole.
    
           Following the creation of fracture H7, mise-a-la-masse measurements on
     borehple 4, contoured in Figure 8.15, show complex distortion of the electrical field.
     Resistivity decreases at most electrodes with a maximum decrease of 13.5% on a
     corner of the array opposite from the fractured borehole.  Resistivity increases at
     one of the electrodes nearest boreholes 4 and 7. This pattern suggests a single
     vertical fracture extending from borehole 7 nearly to borehole 6.
    
           Changes for  mise-a-la-masse  measurements on borehole 7 (Figure 8.16)
     represent a composite of the effects  of fractures H4, H5 and H7. The composite
     result is a decrease  in resistivity of about 10%, extending north and south
     approximately 6 m (20 ft), with less distortion of the electrical field east or west.
     This suggests the fracture from borehole 7 trends north-south rather than east-west.
     The fracture vented into the shallow hole dug for one of the electrodes (E4).  It is
     interesting that the  change in resistivity measured by this electrode, directly in the
     new fracture, is less than that observed at other electrodes.
    
           Following fracturing of borehole 6, mise-a-la-masse measurements of
     fracture H4 showed an increase in apparent resistivity of 6 to  10% in the vicinity of
     the borehole  (Figure 8.17). A region of low apparent resistivity  extends west  from
     borehole 4 toward boreholes 7 and 6, suggesting that fractures have intersected.
     Mise-a-la-masse measurements of composite of the effects of fractures H5, H7 and
     H6 on borehole 6, contoured in  Figure 8.18, yield a more complex pattern. It is
    interesting that the largest decrease in apparent resistivity for this final mise-a-la-
    masse measurement on borehole 6 is nearer borehole 4 than 6.  Because these
    calculations reflect only changes after measurements made subsequent to fracture
    H4 (Figure 8.12), the interpretation that fractures H6, H4, and possibly H7 are
    interconnected must be checked against direct observations.
                                       247
    

    -------
                                                                    10ft
    
    Figure 8.11. Apparent resistivity contour map (mise-a-la-masse) on borehole 4,
               following fracture H4. Contour interval is 1 ohm-meter.
                                    248
    

    -------
                                                                   10ft
    Figure 8.12. Apparent resistivity contour map (mise-a-la-masse) on borehole 6,
               following fracture H4. Contour interval is 1 ohm-meter.
                                    249
    

    -------
                                                                    10ft
    Figure 8.13. Apparent resistivity contour map (mise-a-la-masse) on borehole 5,
               following fractures H4 and H5. Contour interval is 1 ohm-meter.
                                   250
    

    -------
                                                                    10ft
    Figure 8.14. Apparent resistivity contour map (mise-a-la-masse) on borehole 4,
               following fracture H5. Contour interval is 1 ohm-meter.
                                    251
    

    -------
                                                                    10ft
    Figure 8.15. Apparent resistivity contour map (mise-a-la-masse) on borehole 4,
               following fractures H4, H5 and H7.  Contour interval is 1 ohm-
               meter.
                                    252
    

    -------
                                                                  10ft
    Figure 8.16. Apparent resistivity contour map (mise-a-la-masse) on borehole 7,
               following fractures H4, H5 and H7.  Contour interval is 1 ohm-
               meter.
                                    253
    

    -------
                                                                    10ft
    Figure 8.17. Apparent resistivity contour map (mise-a-la-masse) on borehole 4,
               following fracture H6. Contour interval is 1 ohm-meter.
                                   254
    

    -------
                                                                    10ft
    Figure 8.18. Apparent resistivity contour map (mise-a-la-masse) on borehole 6,
               following fracture H6. Contour interval is 1 ohm-meter.
                                  255
    

    -------
              X DECREASE. FRAC 4
                                                   X DECREASE. FRAC 5
             » DECREASE. FRAC 7
    * DECREASE U4 FRAC 6
    Figure 8.19. Summary of mise-a-la-masse apparent resistivity changes relative to
                borehole 4 resulting from fractures H4, H5, H6 and H7. Cross-
                hatching indicates locations of fractures.
                                     256
    

    -------
     borehole 7 toward borehole 6, representing a unilateral rupture, which is probably
     vertical. Figure 8.19d shows a decrease in apparent resistivity toward the midpoint
     between boreholes 6 and 7, with a decrease in apparent resistivity over much of the
     remaining array. This means electrical current density has decreased over most of
     the array, so current must have been diverted through a new conductor developed
     during the creation of fracture H6.
    
           In conclusion, it is clear that hydraulic fracturing with a conductive fluid in a
     relatively resistive  formation causes observable, but complex changes in die
     electrical properties of the formation.  The mise-a-la-masse method yields valuable
     results, but further research is needed before its use could become commonplace.
     Specifically, measurements made as fractures are being propagated may help
     understand and interpret the sometimes complex results.
    
    
     DISCUSSION
    
           Pressure records, surface uplift, and surface tilt all are available methods of
     monitoring hydraulic fractures at shallow depths.  Pressure records require simple
     equipment (a transducer and a data recorder), and indicate the onset of fracturing,
     show various fluctuations in pumping schedule and provide a rapid indicator of
     various problems.  Methods of estimating orientation, asymmetry, and tip screenout
     (e.g. Nolle and Smith, 1981; Smith and others, 1987) from the pressure records of
     shallow fractures seems feasible.
    
           Surface uplifts of a cm or more over hydraulic fractures were readily
     measured with a standard engineer's leveling telescope. Contour plots of uplift
     correlated with shallow hydraulic fractures, providing estimates of the leading edge
     and point of maximum thickness. Numerical inversion of uplift data should improve
     the qualitative correlations attempted in this work, providing estimates of the size,
     thickness, orientation and location of hydraulic fractures at depth.
    
           During the  1988 tests, surface tilts were measured in real time  with
     sophisticated tiltmeters and a datalogger.  Inversion of those data after the tests
     provided estimates of how the fractures developed as a function of time.
     Comparison with maps indicates that the inversion provides crude simplifications of
     the actual fracture  forms. In future applications, real time inversion or the tiltmeter
     data could yield estimates of fracture form as it is being created.
    
           Electrical resistivity measurements, particularly those using the mise-a-la-
     masse technique, are able to sense the shallow fractures created during these tests.
     Methods of inverting electrical signals to obtain estimates of fracture geometry,
    which are currently being developed, will be required before the mise-a-la-masse
    method is practical for monitoring hydraulic fractures.
                                     257
    

    -------
                                   SECTION NINE
    
                          SUMMARY AND CONCLUSIONS
    
           The objective of this study was to evaluate the feasibility of using hydraulic
    fracturing under conditions of contaminated regions. Most current applications of
    hydraulic fracturing are in oil reservoirs, and available techniques are designed for
    that application. Oil reservoirs are typically deeper and composed of different
    materials than contaminated regions, so the suitability of reservoir fracturing
    methods for contaminated soil use is unknown. Contaminants commonly occur in
    soils, or unlithified sediments, that are weaker and more compliant than the
    limestone or sandstone typical of reservoirs. Previous studies of hydraulic fracturing
    of soil have focused on methods of preventing fracture initiation during various
    geotechnical practices, so little is known about the details of the propagation of a
    hydraulic fracture in soil. Moreover, most contaminants occur at relatively shallow
    depths, so that hydraulic fractures could vent to the ground surface before they grow
    any appreciable distance from a well. The technique will be of limited value if
    venting limits fracture lengths to, say, a few well diameters.
    
           The project consisted of studies in the laboratory and in the field, as well as
    an investigation of possible applications and a review of previous work. Laboratory
    studies were intended to reveal details of the process of hydraulic fracturing of soil,
    and to assess the ability of linear elastic fracture mechanics to predict the essential
    details. Field studies were intended to show whether fractures of useful size could
    be created and propped with sand at shallow depths in soil.
    
    
    POTENTIAL APPLICATIONS
    
           Most remedial systems requiring fluid flow either into or out of the
    subsurface could benefit from hydraulic fracturing. Pump and treat systems are
    obvious candidates because they employ procedures resembling those used in
    petroleum recovery, where the benefits  of hydraulic fractures are without question.
    A review of data from oil wells, gas wells, and water wells that have been
    hydraulically fractured indicate a consistent increase in yield. Preliminary data from
    this research indicate that steady-state rates of inflow into unsaturated ground are
    greater, by factors of 3.1 to 9.0, into wells intersecting hydraulic fractures as
    compared with those in unfractured ground. Magnitudes of increase in observed
    yields, in general, are similar to magnitudes calculated using simple theoretical
    analyses. Based on a review of published data and simple analyses, we expect that
    hydraulic fracturing could increase yields of contaminant recovery wells consistently
    by as much as five times, commonly by as much as 10 times, and in some cases it
    could increase yields by factors much more than 10 times.  The magnitude of the
    effect will depend on sue conditions, the methods used to create hydraulic fractures,
    and the methods used to complete and develop associated wells.
    
           Pump and treat systems are by no means the only remedial methods that
    might benefit from hydraulic fracturing. The yields of vapor-producing wells, such
    as those recovering natural gas or steam, are unproved  by hydraulic fracturing, so by
    analogy we expect that vapor extraction systems could benefit from this technology.
    Similarly, hydraulic fracturing could be used in conjunction with steam stripping - a
    process developed  to improve yields of oil wells and currently being tested under
                                       258
    

    -------
     remedial conditions. Horizontal, sheet-like hydraulic fractures placed below a
     contaminated region could be used as gravity drains to intercept the leachate from
     soil flushing systems (Murdoch and others, 1987), improving the effectiveness of that
     remedial action.
    
           The novel application of delivering solid material, as granules pumped into
     hydraulic fractures, is noteworthy. Bio-remediation systems stand to benefit in
     particular because either nutrients for microorganisms, or the microorganisms
     themselves, could potentially be delivered as fine-grained solids in hydraulic
     fractures. Air sparging-pumping air into the ground to encourage the growth of
     aerobic bacteria-could be more effective in ground containing hydraulic fractures.
    
    
     LABORATORY STUDIES: SUMMARY OF RESULTS
    
           1. Hydraulic fracturing in the Center Hill clay can be predicted and analyzed
     using methods of linear elastic fracture mechanics, similar to those used to predict
     hydraulic fracturing in rock. The large body of published papers analyzing hydraulic
     fracturing in rock can be used to predict field applications  in soil, with appropriate
     modifications for boundary conditions (e.g. ground surface), material properties
     (e.g. elastic constants, leakoff parameters), and state of stress encountered at
     shallow depths.
    
           2. Hydraulic fractures created in the laboratory consist of the following
     zones, which are arranged in order from the point of injection to the leading edge:
    
    
     1) Starter slot This feature was created during sample preparation;
    
     2) Parent fracture A continuous fracture surface marked by slight steps, ridges, or
                 grooves.
    
     3) Lobes The parent fracture twists or curves slightly, breaking into a family of
                 discontinuous fracture lobes.
    
     4) Unwetted tip Leading edge of the fracture is unstained  by dye in the injection
                 fluid.
    
           Those zones were identified on almost all fracture surfaces. The only
     exception is that the unwetted tip zone was absent from fractures created in samples
     of less than 21% moisture; the entire fracture surface was stained. Published
     descriptions of the appearance of hydraulic fractures created in rock are similar to
     those produced in soil during this study.
    
           3. Records of injection pressure as a function of time show three
    characteristic periods, which are interpreted as follows:
    
    
    Period I: Injection pressure increases linearly with time, slope constant. Indicates
                 inflation of the starter slot.
                                    259
    

    -------
    Period II: Slope of pressure record decreases but remains positive. Change in slope
                 indicates onset of fracturing, positive slope indicates stable
                 propagation.
    
    Period III: Pressure reaches a maximum and slope becomes negative. Indicates
                 unstable propagation.
    
           In some samples, particularly the wetter ones, propagation begins slightly
    before the change in slope of the record, and the onset of propagation may have
    been overestimated by methods used in the study.
    
           4.  Moisture content and length of starter slot strongly affect the pressure
    record. As moisture content decreases, the duration of stable propagation (Period
    n) diminishes and the slope steepens during unstable propagation (Period III).
    Moreover, the pressure required  to initiate fracturing increases as moisture content
    decreases. Similar trends occur when the length of the starter slot decrease; the
    duration of stable propagation diminishes and the pressure required to initiate
    fracturing increases.
    
           5. The critical stress intensity factor can adequately predict the pressure
    required to initiate hydraulic fracturing of soil. It is independent of the length of the
    starter slot, for the slot lengths used in this study, and it predicts the driving pressure
    at the onset of fracturing to within 10 percent, on average. It was more accurate for
    samples wetter than the plastic limit (20% moisture by weight) and less accurate for
    samples drier than that value.
    
           6. Under the confining stress used in the laboratory experiments, the size of
    the crack-tip process zone is estimated to increase from 0.13 to 0.47 cm as moisture
    decreases from 27 to 20%, according to preliminary measurements. The accuracy of
    the critical stress intensity method diminishes as the size of the process zone
    increases, which is consistent with the findings that accuracy of the laboratory tests
    decreased as samples became drier. Ouchterlony (1982) suggests that the starter
    slot should be at least 93 times longer than the process zone to ensure small scale
    yielding and maintain sufficient accuracy. Slot lengths ranged from 1.22 cm to 5.08
    cm in the experiments, suggesting that the criteria of Ouchterlony are satisfied when
    using samples moister than the plastic limit. Those criteria are not met for the
    shorter slot lengths used in the drier samples, suggesting that slots at least 5 cm  long
    should be used in samples drier than the plastic limit.
    
           7. The value of the critical stress intensity is highly sensitive to moisture
    content, decreasing sharply from  roughly 200 kPa cm1* to 30 kPa cm1/2 as moisture
    content increased by a few percent (from 20 to 22%).  The sharp change in critical
    stress intensity corresponds to the plastic limit of the Center Hill clay.
    
           8. Critical stress intensity  goes to zero for samples of Center Hill clay greater
    than 32% moisture, although hydraulic fractures are readily created under those
    moisture conditions. Critical stress intensity fails as apredictor of moisture
    conditions that preclude the formation of a hydraulic fracture. Presumably,
    hydraulic fracturing will be impossible at extremely high moisture contents when the
    soil approaches a slurry in composition, but the moisture content at which this
                                       260
    

    -------
     occurs exceeds those used in this study and cannot be predicted using critical stress
     intensity.
    
           9. Theoretical analyses based on linear elastic fracture mechanics and linear
     viscous fluid mechanics explain many of the details of the laboratory experiments,
     including the development of an unwetted tip, the average propagation velocity,
     changes in the form of the pressure record with changes in slot length or moisture
     content. Analyses used to predict field applications should include the effects of
     leakoff, which were omitted from the analyses presented here.
    
    
     FIELD STUDIES:  SUMMARY OF CONCLUSIONS
    
           1. Hydraulic fractures of useful size can be created and propped with sand at
     shallow depths in glacial till.
    
           2. Pumps and blenders used by the oil industry will successfully create
     hydraulic fractures  at shallow depths. Blenders and pumps used for the 1988 test
     were designed for pumping rates, pressures and volumes that greatly exceed the
     specifications required for shallow depths. Several shortcomings, including sparse
     proppant density in the fractures created, during the 1988 test, appear to be related
     to lack of control of the oversized equipment.
    
           A blender and pump unit which is readily available for injection grouting and
     was used in the 1989 tests consistently resulted  in hydraulic fractures propped with
     sand. That equipment is inexpensive and simple to operate relative to the oil-field
     equipment, but it too has drawbacks. The blenders were underpowered, resulting in
     long mixing times that limited the rate of injection. A progressive cavity pump used
     to inject slurry suffered excessive wear, and as a result the rotor and stator had to be
     replaced at the end of the project.
    
           3.  Vertical fractures nucleated at the walls of open cylindrical boreholes,
     even when the far-field state of stress favored horizontal fractures.  Shallow
     horizontal notches,  cut 3.75 cm into the walls of the open boreholes, failed to
     nucleate a horizontal fracture in most cases.  The vertical fractures extend several
     cm to several dm away from the open borehole, and then abruptly (typically within a
     few cm) roll over and become flat-lying.  Vertical fractures at the borehole reduced
     the maximum lengths fractures achieved before venting, and they may have
     restricted the transport of proppant.
    
           Vertical fractures created during the 1988 test were a consequence of the
     borehole design because they were eliminated by decreasing the length of the
     cylindrical hole and increasing the size of the notch. Hydraulic fractures adjacent to
     boreholes in till were always horizontal when they were nucleated from notches cut
     10 to 20 cm in till with a water jet, as in the 1989 tests.
    
           4. The form of the fractures created during the 1988 test is characterized by
     the following four zones arranged in increasing  distance from the parent borehole:
    
    
    Zone 1: a sub-vertical orientation adjacent to the borehole.
                                      261
    

    -------
    Zone 2: a flat-lying orientation in the vicinity of the borehole.
    
    Zone 3: a planar to trough-like fracture dipping gently (15° to 25°) toward the
                 borehole. This zone extends in one direction away from the borehole
                 and composes most of the fracture.
    
    Zone 4: a steeply-dipping orientation that vents to the ground surface
    
           In plan the fractures are crudely elliptical, with aspect ratios of roughly 2:3.
    The parent borehole is typically near one end of the major axis and the vent is at the
    other end.
    
           The form of the fractures created during the 1989 test was simpler than that
    of the previous year; the fractures were nearly flat-lying from their parent borehole
    to then- leading edge.  Some fractures dipped gently toward the borehole, but even
    the steepest dip was only 5°. In plan, the fractures created during 1989 were slightly
    elongate (aspect of 2:3) and they were highly assymetric with respect to the
    borehole; in plan they are similar to the results of the previous year.
    
           5. Fractures created during the 1988 dipped more steeply than those of the
    1989 test. The dip is important because it defines how far the fracture can extend
    before venting.  Three differences between the 1988 and 1989 tests are recognized
    that could have affected the average dip.
    
    
           a) Rate of injection during the 1989 test was less than that of the 1988 test.
    Pollard and Holzhausen (1979) show that decreasing  the pressure gradient within
    the fracture (say, by decreasing pumping rate) will inhibit growth toward the ground
    surface.
           b) Sand concentration, and thus slurry density, was greater during the 1989
    test Increasing fluid density will decrease the static drivr	—-——•"«—*'- -
    Secor and Pollard, 1975; Abou-Sayed and others, 1984),
           c) The topographic surface overlying fractures during the 1989 test was
    benched, whereas it was nearly flat over the dipping fractures created in 1988. It
    appears that topography may contribute to the orientation of shallow hydraulic
    fractures (perhaps by inducing horizontal variations in vertical stress), but
    mechanisms controlling this suggestion must be investigated in more detail.
    
           6. The major axes of hydraulic fractures created during both tests extended
    away from parent boreholes in directions that minimize the vertical stress. During
    the 1988 tests, fractures either propagated away from a backhoe parked next to the
    borehole, or they propagated in the same direction as the overlying slope. During
    the 1989 tests, all fractures grew away from a steep embankment.
           Those observations suggest that preferred propagation directions of
     horizontal fractures can be crudely predicted based on topography; they will
    grow in
                                      262
    

    -------
     the downslope direction. This raises the possibility that the direction of propagation
     could be influenced by artificially loading the ground surface, which could be
     important at sites underlain by regions that should not be fractured.  It also indicates
     that a horizontal hydraulic fracture created next to a building will tend to grow away
     from the building.
    
           7.  Fractures were created and filled with sand at depths ranging from 1 to 4
     m. Dimensions of fractures created between 1 and 2 m are known in detail from
     excavations.  They are as follows: during the 1988 tests, the maximum length, from
     borehole to the leading edge, ranged from 1.8 to 13.5 m and it averaged 578 m.
     Areas covered by those fractures ranged from 2.2 to 90 m2, and averaged 25 m2.
     The thickest part of a fracture created in 1988 was 9mm, and many of the fractures
     created that year were nearly devoid of sand.
                                                         i
           During the  1989 tests, the maximum lengths ranged from 0.8 to 7.2 m, and
     averaged 4.0 m.  Areas covered by the fractures ranged from 0.8 to 36.7 m2, and they
     averaged 19.2 m2.  All the fractures contained sand, ranging in maximum thickness
     from 2 to 20 mm, and averaging 11 mm.
    
           The fractures created in 1989 were slightly smaller in area than those of the
     previous year; however, the sizes of all the 1988 fractures were limited by venting,
     whereas most of the 1989 fractures did not vent and they could have been larger if
     more fluid had been pumped into them. The average thickness of sand in the 1989
     fractures exceeded the maximum thickness from the previous year.
    
           The increase in thickness of sand, from the 1988 to the 1989 tests, apparently
     results from changes in the above-ground mixers and pumps. The above-ground
     equipment used in 1988 was designed for flow rates and pressures required to create
     hydraulic fractures in oil wells, which are as much as several orders or magnitude
     more than the flow rates and pressures required for our applications. The oil-field
     equipment lacked the control required to mix and pump the relatively small
     applications that we required. This simply indicates that filling hydraulic fractures
     with sand will be facilitated by using above-ground equipment that is designed for
     the conditions of the application.
    
           8. A method of creating fractures from a device driven into the ground was
     developed to facilitate the creation of multiple fractures in soil. As many as four,
     fiat-lying hydraulic fractures were stacked  at spacings of 30 cm without intersecting
     their neighbors. When one fracture was created 15 cm below another, the lower
     fracture would commonly climb upward and intersect the overlying fracture several
     meters from the borehole. One spacing of 7 cm was attempted; the lower fracture
     intersected the upper one several dm from the borehole.
    
           It would be possible to stack flat-lying fractures at spacings between 15 and
    30 cm throughout a contaminated region at the ELDA site, and similar spacings
    should be possible under similar site conditions.
    
           9. Pressure records, surface uplift, and surface tilt all are available methods
    of monitoring hydraulic fractures at shallow depths. Pressure records require simple
    equipment (a transducer and a data recorder), and indicate the onset of fracturing,
    show various floatations in pumping schedule and provide a rapid indicator of
                                     263
    

    -------
    various problems.  Methods of estimating orientation, assymetry, and tip screenout
    (e.g. Nolte and Smith, 1981; Smith and others, 1987) from the pressure records of
    shallow fractures seems feasible.
    
           Electric resistivity measurements, particularly those using the mise-a-la-
    masse technique, were able to detect the hydraulic fractures created during the 1988
    tests. Methods of inverting electrical signals to obtain estimates of fractures
    geometry will be required, however, before this method yield quantitative estimates
    of fracture geometry.
    
           Surface uplifts of a cm or more over hydraulic fractures were readily
    measured with a standard engineer's leveling telescope. Contour plots of uplift
    correlated with shallow hydraulic fractures, providing estimates of the leading edge
    and point of maximum thickness.  Numerical inversion of uplift data should provide
    valuable estimates of the size, thickness and location of hydraulic fractures at depth.
    
           During the 1988 tests, surface tilts were measured in real time with tiltmeters
    and a datalogger. Inversion of those data after the tests provided estimates of how
    the fractures developed as a function of time.  Comparison with maps indicates that
    the inversion provides crude simplifications of the actual fracture forms. In future
    applications, real time inversion of the tiltmeter data could yield estimates of
    fracture form as it is being created.
    
           Electrical resistivity measurements, particularly those using the mise-a-la-
    masse technique, are able to sense the shallow fractures created during these tests.
    Methods of inverting electrical signals to obtain estimates of fracture geometry,
    which are currently being developed, will be required before the mise-a-la-masse
    method is practical for monitoring hydraulic fractures.
                                   264
    

    -------
                               REFERENCES CITED
     Aamodt, R.L and M. Kuriyagawa. 1982. Measurement of instantaneous shut-in
           pressure in crystalline rock. Proceedings of Workshop on Hydraulic
           Fracturing Stress Measurements. Zoback, M.D. and C. Haimson.(eds.) U.S.
           National Comm. for Rock Mechanics. Menlo Park, CA. 394-402.
    
     Abe, H., T. Mura and LM. Keen 1976. Growth rate of a penny-shaped crack hi
           hydraulic fracturing of rocks. J. Geop. Res. 81:5335-52
    Abou-Sayed, A.S. 1977. Fracture toughness, K, of triarially loaded Indiana
           milestone. In: Proc. 18th U.S. Sym. Rock Mech. 1-8.
    
    Abou-Sayed, A.S., C.E. Brechtel and RJ. Clifton. 1978. In-situ stress determination
           by hydrofracturing: A fracture mechanics approach. J. Geop. Res. 83:2851-
           2862.
    
    Abou-Sayed, A.S., K.P. Sinha and RJ. Clifton. 1984. Evaluation of the influence of
           in-situ reservoir conditions in the geometry of hydraulic fractures using a 3-D
           simulator: Part l~Technical Approach, SPE Paper 12877. In: Proc. Uncon.
           Gas Rec. Sym., Pittsburg, PA.(May) 433^38.
    
    Abramowitz, M. and LA. Stegun. 1964. eds. Handbook of Mathematical Functions
           with Formulas, Graphs, and Mathematical Tables.  U.S. Govt. Printing
           Office. Washington, DC. 1046.
    
    Acharya, A. 1988. Hydraulic-fracture-treatment design simulation. SPE Production
           Engineering. 40:139-142.
    
    Advani, S.H., H. Khattab and J.K. Lee. 1985. Hydraulic fracturing geometry,
           modeling, prediction and comparisons, SPE Paper 13863. In: Proc. Low
           Perm. Gas Res. Sym., Denver, CO.(May) 135-139.
    
    Ahmed, U., J. Strawn, M. Wilson and J. Schatz. 1983. Effect of stress distribution on
           hydraulic fracture geometry: A laboratory simulation study in one meter
           cubic blocks, SPE Paper 11637. In: Proc. SPE Sym. Low Permeability,
           Denver, CO. 1-5.
    
    Anderson, G.D. 1979. Laboratory experiments on hydraulic fracture growth near an
           interface. In: Proc. U.S. Sym. Rock Mech., Golden, CO. 19:333-339.
    
    Anderson, FLA. 1974. Slurry injection of vein materials in the Coeur d'Alene
           district. Economic Geology. 69:414-415.
    
    Anderson, R.W. and A.M. Phillips. 1988. Practical application of economic well-
           performance criteria to the optimization of fracturing treatment design. SPE
           Production Engineering. 40:223-228.
    
    Babcock, R.E., C.L. Prokop and R.O. Kehle.  1967. Distribution of propping agents
           in vertical fractures. Prod. Monthly. 11-18.
                                     265
    

    -------
    Barenblatt, G.1.1962. The mathematical theory of equilibrium cracks in brittle
          fracture. Adv. Appl. Mech. 7:55-129.
    
    Baski, H. 1987. Hydrofracturing of water wells. Water Well J. 41:6:34-35.
    
    Blot, M.A., L. Masse and W.L. Medlin. 1982. A two-dimensional theory of fracture
          propogation, SPE Paper 11067. In: Proc. Ann. SPE Fall Tech. Conf., New
          Orleans, LA. 1-15.
    
    Biot, M.A., W.L. Medlin and L. Masse. 1983. Fracture penetration through an
          interface. Soc. Petr. Eng. J. 857-869.
    
    Bjemim, L and K.H. Andersen. 1972. In-situ measurement of lateral pressures in
          clay. Geotechnique. 15:11-20.
    
    Bjemim, L, J.K.L. Nash, R.M. Kennard and R.E. Gibson. 1972. Hydraulic
          fracturing in field permeability testing. Geotechnique. 22:319-332.
    
    Boone, TJ., A.R. Ingraffea and J.C. Roegiers. 1989. Visualization of hydraulically
          driven fracture propagation in poroelastic media using a superworkstation. J.
          Petr. Tech., Trans. AIME. 41:  ' ™
    Bouteca, MJ. 1988. Hydraulic fracturing model based on a three-dimensional
                               analysis off
          Production Engineering. 3:445-454.
    closed form: Tests and analysis of fracture geometry and containment. SPE
    Bouwer, H. 1978.  Groundwater Hydrology. McGraw-Hill Book Co. New York.
          480.
    
    Briones, A.A. and G. Uehara. 1977a. Soil Elastic Constants: I. Calculations from
          sound velocities. Soil Sci. Soc. Am. J. 41:22-25.
    
    Briones, A.A. and G. Uehara. 1977b. Soil Elastic Constants: II. Application to
          analysis of soil cracking. Soil Sci. Soc. Am. J. 41:26-29.
    
    Brooker, E.W. and H.O. Ireland. 1965. Earth pressures at rest related to stress
          history. Canadian Geotech. J. 2:1-15.
    
    Brunsing, T.P. 1987. The block displacement method field demonstration and
          specifications. U.S. EPA Project Summary. 600-52-87-023:1-6.
    
    Brunsing, T.P. and R.B. Henderson. 1984. A laboratory technique for assessing the
          in-situ constructability of a bottom barrier for waste isolation. In: Proc. Nat.
          Conf. Manag. Uncont Haz. Wast Sites, Washington, DC 135-140.
    
    Bryant, D.G. 1968. Intrusive breccias associated with ore, Warren (Bisbee) Mining
          District, Arizona. Economic Geology. 63:1-12.
    
    
    Cinco-Ley, H., V.F. Samaniego and N.A. Dominguez. 1978. Transient pressure
          behavior for a well with a finite-conductivity vertical fracture. Soc. Petr. Eng.
          J. 253-264.
    
    
                                      266
    

    -------
     Cinco-Ley, H., V.F. Samaniego and A. Dominguez. 1981. Transient pressure
           analysis for fractured wells. J. Petr. Tech., Trans. AIME. 1749-1764.
    
     Clark, J.B. 1949. A hydraulic process for increasing the productivity of wells. Petr.
           Trans. AIME. 2510:1-8.
    
     Cleary, M.P. 1978a. Primary factors governing hydraulic fractures in heterogeneous
           stratified porous formations. In: Proc. ASME Energy Tech. Conf. & Exh.,
           Houston, TX.(Nov.) 1-12.
    
     Cleary, M.P. 1978b. Some deformation and fracture characteristics of oil shale. In:
           Proc. U.S. Sym. Rock Mech., Rapid City, SD. 72-82.
    
     Cleary, M.P. 1980a. Analysis of mechanisms and procedures for producing
           favourable shapes of hydraulic fractures, SPE Paper 9260. In: Proc. 55th Ann.
           SPE Fall Tech. Conf. Exh., Dallas, TX. 1-12.
    
     Cleary, M.P. 1980b. Comprehensive design formulae for hydraulic fracturing, SPE
           Paper 9259. In: Proc. 55th Ann. SPE Fall Tech. Conf. Exh., Dallas, TX. 1-16.
    
     Cleary, M.P. 1988. The engineering of hydraulic fractures: State of the art and
           technology of the future. J. Petr. Tech., Trans. AIME. 13-21.
    Cleary, M.P. and S.K. Wong. 1985. Numerical simulation of unsteady fluid flow and
           propagation of a circular hydraulic fracture. Int. J. Num. Anal. Meth.
           Geomech. 9:1-14.
    
    Cleary, M.P., A.R. Crockett, J.I. Martinez and V.M. Narendran. 1983a. Surface
           integral schemes for fluid flow and induced stresses around fractures in
           underground reservoirs, SPE Paper 11632. In: Proc. SPE Sym. Low
           Permeability, Denver, CO. 283-292.
    
    Cleary, M.P., M. Kavyadas and K.Y. Lam. 1983b. Development of a fully three-
           dimensional simulator for analysis and design of hydraulic fracturing, SPE
           Paper 11631. In: Proc. SPE Sym. Low Permeability, Denver, CO. 271-278.
    
    Clifton, R J., E.R. Simonson, A.H. Jones and S J. Green. 1976. Determination of the
           critical-stress-intensity factor K from internally pressurized thick-walled
           vessels. Experimental Mechanics. 233-238.
    
    Cornet, F.H., J.M. Hosanski, F. Bernaudat and E. Ledoux. 1982. Shallow depth
    Cotterell, B. and J.R. Rice. 1980. Slightly curved or kinked cracks. Int. J. Fracture.
          16:155-169.
    
    Crockett, A.R., R.M. Willis and M.P. Cleary. 1989. Improvement of hydraulic
          fracture predictions by real-time history matching on observed pressures
          SPE Production Engineering. 4:408-416.
                                      267
    

    -------
     Daneshy, A-A. 1973a. Experimental investigation of hydraulic fracturing through
           perforations, SPE Paper 4333. In: Proc. 2nd Ann. SPE European Meeting,
           London, England. 1-12.
     Daneshy, AA 1973b. On the design of verticla hydraulic fractures. J. Petr. Tech.,
           Trans. AIME. 83-97.
    
     Daneshy, AA 1973c. A study of inclined hydraulic fractures. Soc. Petr. Eng. J.
           13:61-68.                                                   e
     Daneshy, AA. 1973d. Three-dimensional propagation of hydraulic fractures
           extending from open holes. Rock Mechanics. 157-179.
    
     Daneshy, A.A. 1976a. Hydraulic fracture propagation in layered formations. Soc.
           Petr. Eng. J. 33-41.
    
     Daneshy, A.A. 1976b. Rock properties controlling hydraulic fracture propagation,
           SPE Paper 5752. In: Proc SPE European Spring Meeting, Amsterdam, The
           Netherlands.(April) 1-8.
    
     Daneshy, A.A. 1978a. Hydraulic fracture propagation in layered formations. Soc.
           Petr. Eng. J. 33-41.
    
     Daneshy, A.A. 1978b. Numerical solution of sand transport in hydraulic fracturing.
           J. Petr. Tech., Trans. AIME. 132-140.
    
     Davis, P.M. 1983. Surface deformation associated with a dipping hydrofracture. J.
           Geop. Res. 88:5826-5834.
    
     Dean, R.H. and S.H. Advani. 1984. An exact solution for pistonlike leak-off of
           compressible fluids. J. Energy Res. Tech., Trans. ASME. 106:539-542.
    
     De Laguna, W. 1966a. Disposal of radioactive wastes by hydraulic fracturing, Part I:
           General concept and first field experiments. Nuclear Eng. & Design. 3:338-
          352.
    De Laguna, W. 19665. Disposal of radioactive wastes by hydraulic fracturing, Part
          II: Mechanics of fracture formation and design of observation and
          monitoring wells. Nuclear Eng. & Design. 3:432-438.
    
    Detournay, E. 1979. The interaction of deformation and hydraulic conductivity in
          rock fracture. Improved stress determination proceedures by hydraulic
          fracturing.  Univ. of Minn., Civil & Mineral Eng. Dept. Minneapolis. 1-42.
    
    Dickinson, R.M. 1988. Review of consolidation grouting of rock masses and
          methods for evaluation. U.S. Army Corps of Eng. Final Report REMR-GT-
          8.86.
    
    Dugdale, D.S. 1960. Yielding of steel sheets containing slits. J. Mech. Phys. Solids.
          8:100-104.
                                      268
    

    -------
     Dyes, A.B., C.E. Kemp and B.H. Claude. 1958. Effect of fractures on sweep-out
           pattern. Petr. Trans. AIME. 213:245-248.
     Elbel, J.L. and P.A. Sookprasong. 1987. Use of cumulative-production type curves in
           fracture design. SPE Production Engineering. 2:191-198.
     Ernst, P. 1980. Linking of boreholes. Meded. Rijks. Geol. Dienst. 33:35-40.
     Fairhurst, C. 1964. Measurement of in-situ rock stresses, with particular reference to
           hydraulic fracturing. Felsmechanik. 2:129-147.
     Farmin, R. 1934. Pebble dikes and associated mineralization at Tintic, Utah.
           Economic Geology. 29:356-370.
     Farrell, DA, E.L. Greacen and W.E. Larson. 1967. The effect of water content on
           axial strain in a loam soil under tension and compression. Soil Sci. Soc. Am.
           Proc. 31:445-450.
     Geertsma, J. and F. de Klerk. 1969. A rapid method of predicting width and extent
           of hydraulically induced fractures. J. Petr. Tech., Trans. AIME. 1571-1581.
     Geertsma, J. and R. Haafkens. 1979. A comparison of theories for predicting width
           and extent of vertical hydraulically induced fractures. J. Energy Res. Tech.,
           Trans. ASME.  101:8-19.
     Griffith, A.A. 1920. The phenomena of rupture and flow in solids. Phil. Trans. Roy.
           Soc. London. 221:163-198.
     Griffith, A.A. 1924. The theory of rupture. In: Proc. 1st Int. Cong. Appl. Mech.,
           Delft. 55-63.
     Griffiths, S.K., R.H. Nilson and F.A.  Morrison. 1983. Hybrid analytical/numerical
          computation of heat transfer in a gas driven fracture. Paper No. 83-WA/HT-
          35, Winter Meeting, Boston, MA. 1-9.
     Gringarten, A.C. 1982. Flow-test evaluation of fractured reservoirs. Geol. Soc.
          Amer. Spec. Paper. 189:237-263.
     Gringarten, A.C. and H J. Ramey. 1973. The use of source and Green's functions in
          solving unsteady flow problems in reservoirs. Soc. Petr. Eng. J. 285-296.
     Gringarten, A.C. and H J. Ramey. 1974. Unsteady-state pressure distributions
          created by a well with a single horizontal fracture, partial penetration, or
          resticted entry.  Soc. Petr. Eng. J. 413-426.
    Gringarten, A.G, HJ. Ramey and R. Raghaven. 1974. Unsteady-state pressure
          distributions created by a well with a single infinite-conductivity vertical
          fracture. Soc. Petr. Eng. J. 347-360.
    Gronseth, J.M. 1979. Stress determination by hydraulic fracturing. Improved stress
          determination procedures by hydraulic fracturing. Univ. of Minn., Civil &
          Mineral Eng Dept. Minneapolis. 1-114.
                                       269
    

    -------
    Haimson, B. and C. Fairhurst. 1967. Initiation and extension of hydraulic fractures
          in rocks. Soc. Petr. Eng. J. 7:310-318.
                                                              -permeable
    Haimson, B. and C. Fairhurst. 1969. Hydraulic fracturing in porous
          materials. J. Petr. Tech., Trans. AIME. 21:811-817.
    
    Haimson, B. and C. Fairhurst. 1970. In-situ stress determination at great depth by
          means of hydraulic fracturing. (28) Rock Mechanics: Theory and Practice.
          Somerton, w.H.(ed.) The American Institute of Mining, Metallurgical, and
          Petroleum Engineers, Inc. New York. 559-584.
    
    Haimson, B. and E J. Stahl. 1970. Hydraulic fracturing and the extraction of
          minerals through wells. In: Proc. 3rd Sym. on Salt, Northern Ohio Geol. Soc.,
          Cleveland, OHL2:421-432.
    
    Hanson, M.E., G.D. Anderson, R J. Shaffer, D.N. Montan, B. Haimson and M.P.
          Cleary. 1978a. I .T.L gas stimulation program quarterly progress report
          January through March 1978. Lawrence Livermore Laboratories Report. No.
          UCRL 536-78-1-2:1-25.
    
    Hanson, M.E., G.D. Anderson, R J. Shaffer, D.N. Montan, L.D. Thorson, W. Lin, C.
          Haimson and M.P. Cleary. 1978b. IJJ, gas stimulation program quarterly
          progress report October through December 1977. Lawrence Livermore
          Laboratories Report. UCRL 50036-784:1-25.
    
    Hanson, M.E., G.D. Anderson, R J. Shaffer, D.F. Towse, A.G. Duba, N.R.
          Burkhard, W. Lin, M.P. Cleary and B.C. Haimson. 1979. LLL gas stimulation
          program quarterly progress report January through March 1979. Lawrence
          Livermore Laboratories Report. UCRL 50036-79-1:1-24.
    
    Harrison, E., W.F. Kieschnick and W J. McGuire. 1954. The mechanics of fracture
          induction and extension. Petr. Trans. AIME. 20:252-263.
    
    Hartley, R. and M.G.R. Bosnia. 1985. Fracturing in chalk completions. J. Petr.
          Tech., Trans. AIME. 73-79.
    
    Hartsock, J.H. and J.E. Warren. 1961. The effect of horizontal hydraulic fracturing
          on well performance. J. Petr. Tech., Trans. AIME. 1050-1056.
    
    Helmer, 0.1957. Mechanical analysis of the dike pattern of the Spanish Peak area,
          Colorado. Geol. Soc. Am. Bull. 68:567-576.
    
    Holzer, T.L., T.L. Youd and T.C Hanks. 1989. Dynamics of liquefaction during the
          1987 Superstition Hills, California Earthquake. Science. 244:56-59.
    
    Holzhausen, G.R., C.C Card, J.M. Raisbeck and T.L Dobecki. 1985a. Hydraulic
          fracture growth during steam stimulation in a single-well test, SPE Paper
          13619. In: Proc. SPE Regional Meeting, Ventura, CA.
                                      270
    

    -------
    Holzhausen, G.R., C.S. Haase, S.H. Stow and G. Gazonas. 1985b. Hydraulic-
           fracture growth in dipping anisotropic strata as viewed through the surface
           deformation field. In: Proc. 26th U.S. Sym. Rock Mech., Rapid City, SD.
    
    Horsrud, P., R. Risnes and R.K. Bratli. 1982. Fracture initiation pressures in
           permeable poorly consolidated sands. Int. J. Rock Mech. Min. Sci. Geomech.
           Abs. 19:255^266.
    
    Howard, G.C. and C.R. Fast. 1957. Optimum fluid characteristics for fracture
           extension. Drilling & Prod. Prac., API. 261-270.
    
    Howard, G.C. and C.R. Fast. 1970. Hydraulic Fracturing.  Society of Petroleum
           Engineers AIME. New York. 198.
    
    Hubbert, M.KL and D.G. Willis. 1957. Mechanics of hydraulic fracturing. Petr.
           Trans. AIME. 210:153-168.
    
    Huck, PJ., MJ. Waller and S.L. Shimondle. 1980. Innovative geotechnical
           approaches to the remedial in-situ treatment of hazardous materials disposal
           sites. In: Proc. Nat. Conf. Control Haz. Mater. Spills, Louisville, KY. 421-426.
    
    Hurlburt, S. 1989. Moving ahead with hydrofracing. Water Well J.(Feb.) 37-40.
    
    Inglis, C.E. 1957. Stresses in a plate due to the presence of cracks and sharp corners.
           Inst. Naval. Architects, Trans., London. 55:1.
    
    Irwin, G.R. 1957. Analysis of stresses and strains near the end of a crack traversing a
           plate. J. Appl. Mech., Trans. ASME. 361-364.
    Jaworski, G.W., J.M. Duncan and H.B. Seed. 1979. An experimental study of
          hydraulic fracturing, Report UCB/GT/79-02. U.S. Bur. Reclam.(Oct.)
    234.
    Jaworski, G.W., J.M. Duncan and H.B. Seed. 1981. Laboratory study of hydraulic
          fracturing. J. Geotech. Eng. GT6:713-732.
    
    Johnson, A.M. 1970. Physical Processes in Geology.  Freeman Cooper and
          Company. San Fransisco. 577.
    
    Kaufman, J.G. and F.G. Nelson. 1974. More on specimen size effects in fracture
          toughness testing. Fracture Toughness and Slow-Stable Cracking, ASTM
          STP 559. Irwin, G.R.(ed.) Amer. Soc. for Testing & Materials. Baltimore.
          74-85.
    
    Keck, R.G., M.P. Geary and A. Crockett. 1984. A lumped numerical model for the
          design of hydraulic fractures, SPE Paper 12884. In: Proc. SPE
          Unconventional Gas Recovery Sym., Pittsburgh, PA. 507-514.
    
    Kehle, R.0.1964. The determination of tectonic stresses through analysis of
          hydraulic well fracturing. J. Geop. Res. 69:259-273.
    
    Kerr, R.A. 1987. Hot dry rocks: Problems, Promise. Science. 238:1226-1228.
    
    
                                      271
    

    -------
    Kesler, S.E., N. Russell, M. Seaward, M. Rivera, K. McCurdy, G.L Gumming and
          J.F. Sutler. 1981. Geology and geochemistry of sulfide mineralization
          underlying the Pueblo vTejo gold-silver oxide deposit, Dominican Republic.
          Economic Geology. 76:1096-1117.
    
    Khristianoyic, S.A. and Y.P. Zheltov. 1955. Formation of vertical fractures by means
          of highly viscous liquid. In: Proc. 4th World Petroleum Congress, Rome.
    Kirkham, D., M.F. De Boodt and L. De Leenheer. 1959. Modulus of rupture
          determination on undisturbed soil core samples. Soil Science. 87:141-144.
    
    Klepaczko, J.R., M.N. Bassim and T.R. Hsu. 1984. Fracture toughness of coal under
          quasi-static and impact loading. Eng. Frac Mech. 19:305-316.
    
    Knauss, W.G. 1976. Fracture of solids possessing deformation rate sensitive material
          properties. In: Proc. ASME Winter Ann. Meeting, New York. AMD-19:70-
          103.
    
    Koenig, L. 1960. Survey and analysis of well stimulation performance. J. Am. Water
          Works Assoc. 52:333-350.
    
    Lam, K.Y. and M.P. Geary. 1984. Slippage and re-initiation of hydraulic fractures at
          frictional interfaces. Int. J. Num. Anal. Meth. Geomech. 8:589-604.
    
    Lambe, T.W. and R.V. Whitman. 1969. Soil Mechanics.  John Wiley and Sons.
          New York. 553.
    
    Landrum, B.L. and P.B. Crawford. 1961. Horizontal fractures do affect ultimate
          recovery. The Petroleum Engineer. 80-98.
    
    Lawn, B.R. and T.R. Wilshaw. 1975. Fracture of Brittle Solids.  Cambridge
          University Press. Cambridge.
    
    Leach, R.E. 1977. Hydraulic fracturing of soils: A literature review. Misc. Paper S-
          77-6 U.S. Army Waterway Experiment Station.(Mar.) 1-25.
    
    Luiskutty, C.T., L. Tomusta and I.D. Palmer. 1989. A three-dimensional
          semianalytical model of hydraulic fracture growth through weak barriers.
          SPE Production Engineering. 4:226-232.
    
    Macaulay, D. 1987. Hydrofracturing the hard rock well. Ground Water Age. 21:22-
    
    
    Massarch, K.R., R.D. Holtz, B.C. Holm and A. Fredriksson. 1975. Measurement of
          horizontal in-situ stresses. In: Proc. ASCE-GED Spec. Conf. In-situ Meas.
          Soil Prop., Raleigh, NC 1:266-286.
    
    Massarsch, K.R.  1978. New aspects of soil fracturing in clay. J. Geotech. Eng. Div.
          GT8:1109-1123.
    
    McGuire, W J. and V J. Sikora. 1960. The effect of vertical fractures on well
          productivity. Petr. Trans. AIME. 219:401-403.
    
                                       272
    

    -------
     McLennan, J.D. and J.C. Roegiers. 1982. Do instantaneous shut-in pressures
           accurately represent the minimum principal stress?  Proceedings of
           Workshop on Hydraulic Fracturing Stress Measurements. Zoback, M.D. and
           B.C. Haimson.(eds.) U.S. National Comm. for Rock Mechanics. Menlo Park,
           CA. 181-207.
    
     Medlin, W.L and L. Masse. 1979. Laboratory investigation of fracture initiation
           pressure and orientation. Soc. Petr. Eng. J. 129-144.
    
     Medlin, W.L and L Masse. 1982. Plasticity effects in hydraulic fracturing, SPE
           Paper 11068. In: Proc. Ann. SPE Fall Tech. Conf., New Orleans, LA. 1-14.
    
     Medlin, W.L and L. Masse. 1984. Laboratory experiements in fracture propagation.
           Soc. Petr. Eng. J. 256-268.
    
     Mendelsohn, DA. 1984. A review of hydraulic fracture modeling, Part I: General
           concepts, 2D models, motivation for 3D modeling. J. Energy Res. Tech.,
           Trans. ASME. 106:369-376.
    
     Meyer, B.R. 1989. Heat transfer in hydraulic fracturing. SPE Production
           Engineering. 4:423-429.
    
     Mony, P.M. 1989. Hydro-frac: One Man's Opinion. Water Well J.(Feb.) 48-50.
    
     Morgenstern, N.R. and P.R. Vaughan. 1963. Some observations on allowable
           grouting pressures. In: Proc. Conf. Grouts and Drill. Muds, Instit. Civ. Eng.,
           London. 36-42.
    
     Mori, A. and M. Tamura.  1987. Hydrofracruring pressure of cohesive soils. Soils and
           Foundations. 27:14-22.
    
     Morita, N., D.L. Whitfill and H.A. Wahl. 1988. Stress-intensity factor and fracture
           cross-sectional shape predictions from a three-dimensional model for
           hydraulically induced fractures. SPE Production Engineering. 40:1329-1342.
    
    Morrisson, T.E. and J.H. Henderson. 1960. Gravity drainage of oil into large
           horizontal fractures. Petr. Trans. AIME. 219:7-15.
    
    Murdoch, L.C., G. Holzhausen, J.C. Roegiers and B.K. Naceur. 1987. Innovative
           delivery and recovery systems: Hydraulic fracturing. EPA Center Hill
           Research Facility Interim Report. 46.
           , H. 1982. Hot dry rock reservoir development and testing in the USA.
          Hydraulic Fracturing and Geothermal Energy. Nemat-Nasser, S., H. Abe
          and S. Hirakawa.(eds.) Martinus Nijhoff Pub. The Hague. 33-58.
    
    Muskat, M. 1937. Use of data on build-up of bottom hole pressures. Trans. AIME.
          123:44-48.
                                      273
    

    -------
     Nakatsuka, K., H. Takahashi and M. Takanohashi. 1982. Hydraulic fracturing
           experiment al Nigorikawa and fracture mechanics evaluation.  Hydraulic
           Fracturing and Geothermal Energy. Nemat-Nasser, S., H. Abe and S.
           Hirakawa.(eds.) Martinus NijhoffPub. The Hague. 95-111.
    
     Narendran, V.M. and M.P. Cleary. 1983. Analysis of growth and interaction of
           multiple hydraulic fractures, SPE Paper 12272. In: Proc. SPE Sym. Res. Sim.,
           San Francisco, CA. 389-397.
    
     Narendran, V.M. and M.P. Cleary. 1984. Elastoslatic interaction of multiple
           arbitrarily shaped cracks in plane inhomogeneous regions. Eng. Frac. Mech.
           1 U*dV l _ClMh
           19:481-506.
    
     NAVFAC. 1982. Soil Mechanics Design Manual 7,1. Dept. of Navy, Navy
           Facilities Eng. Command. Alexandria. 348.'
    
     Nilson, R.H. 1981. Gas-driven fraclure propagation. J. Appl. Mech., Trans. ASME.
           48:757-762.
    
     Nilson, R.H. 1986. An integral method for predicting hydraulic fracture propagation
           driven by gases or liquids. Int. J. Num. Anal. Meth. Geomech. 10:191-211.
    
     Nilson, R.H. and S.K. Griffiths. 1986. Similarity analysis of energy transport in gas-
           driven fractures. Int. J. Fraclure. 30:115-134.
    
     Nilson, R.H., WJ. Proffer and R.E. Duff. 1985. Modelling of gas-driven fractures
           induced by propellant combustion wilhin a borehole. Int J. Rock Mech. Min.
           Sci. Geomech. Abs. 22:3-19.
    
     Nolle, K.G. 1979. Determination of fraclure parameters from fracluring pressure
           decline, SPE Paper 8341. In: Proc. 54th Ann. SPE Fall Tech. Conf Exh., Las
           Vegas, NV. 1-16.
    
     Nolle, K.G. 1982. Fraclure design considerations based on pressure analysis, SPE
           Paper 10911. In: Proc. SPE Cotton Valley Sym., Tyler, TX.
    
     Nolle, K.G. 1984. Determination of proppant and fluid schedules from fracturing
           pressure decline, SPE Paper 13278. In: Proc. 59th SPE Ann. Tech. Conf. &
           Exh., Houston, TX.
    
     Nolle, K.G. 1988a. Application of fracture design based on pressure analysis. SPE
           Production Engineering. 3:31-42.
    
    Nolle, K.G. 1988b. Principles for fraclure design based on pressure analysis. SPE
           Production Engineering. 3:22-30.
    
    Nolle, K.G. and M.B. Smilh. 1981. Interpretation of fracturing pressures. J. Petr.
           Tech., Trans. AIME. 1767-1775.
    
    Nolle, K.G. and M.B. Smilh. 1987. Hydraulic Fracluring Pressure Analysis.  Isl NSI
          Technologies, Inc. Tulsa. 96.
    
    
                                      274
    

    -------
     Nordgren, R.P. 1972. Propagation of a vertical hydraulic fracture. Soc. Petr. Eng. J.
           306-314.
     Northrop, DA, N.R. Warpinski, R.A. Schmidt and C.W. Smith. 1978. Stimulation
           Improved Drilling Meth., Tulsa,
     Novotny, EJ. 1977. Proppant transport, SPE Paper 6813. In: Proc. SPE Ann. Tech.
           Meeting & Exh., Denver, <~
    Ode, H. 1957. Mechanical analysis of the dike pattern of the Spanish Peak area,
           Colorado. Geol. Soc. Am. Bull. 68:567-576.
    
    Ouchterlony, F. 1982. Review of frature toughness testing of rock. SM Archives.
           7:131-211.
    
    Paillet, F.L. 1985. Problems in fractured-reservoir evaluation and possible routes to
           their solution. The Log Analyst. 26:26-41.
    
    Palmer, I.D. and C.T. Luiskutty. 1986. Comparison of hydraulic fracture models for
           highly elongated fractures of variable height. J. Energy Res. Tech., Trans.
           ASME. 108:107-115.
    
    Papadoloulos, J.M., V.M. Narendran and M.P. Cleary. 1983. Laboratory simulations
           of hydraulic fracturing, SPE Paper 11618. In: Proc. SPE Sym. Low
           Permeability, Denver, CO. 161-168.
    
    Pascal, H. 1986. Rheological behavior effects of non-newtonian fracturing fluids on
           propagation of vertical hydraulic fracture. Int. J. Num. Anal. Meth.
           Geomech. 10:443-448.
    
    Perkins, T.K. and LJR. Kern.  1961. Widths of hydraulic fractures. J. Petr. Tech.,
           Trans. AIME. 937-949.
    
    Phillips, WJ. 1972. Hydraulic fracturing and mineralization. J. Geol. Soc. London.
           128:337-359.
    
    Phillips, WJ. 1973. Mechanical effects of retrograde boiling and its probable
           importance in the formation of some porphyry ore deposits. Instit. Mining
           Metallurgy Trans. 82:B90-B98.
    
    Pollard, D.D. 1973. Derivation and evaluation of a mechanical model for sheet
           intrusions. Tectonophysics. 19:233-269.
    
    Pollard, D.D. 1978. Forms of hydraulic fractures as deduced from field studies of
           sheet intrusions. United States Symposium of Rock Mechanics. Kim,
           Y.S.(ed.) University of Nevada. Reno. 1-9.
                                       275
    

    -------
    Pollard, D.D. and G. Holzhausen. 1979. On the mechanical interaction between a
           fluid-filled fracture and the earth's surface. Tectonophysics. 53:27-57.
    Pollard, D.D., O.H. Muller and D.R. Dockstader. 1975. The form and growth of
           fingered sheet intrusions. Geol. Soc. Am. Bull. 86:351-363.
    Pollard, D.D., P. Segall and P.T. Delaney. 1982. Formation and interpretation of
           dilatant echelon cracks. Geol. Soc. Am. Bull. 93:1291-1303.
    Poulos, H.G. and E.H. Davis. 1974.  Elastic Solutions for Soil and Rock Mechanics.
           John Wiley & Sons, Inc. New York. 411.
    Prats, M. 1961. Effect of vertical fractures on reservoir behavior-incompressible
           fluid case. Soc. Petr. Eng. J. 105-118.
    Prats, M. 1981. Effect of burial history on the subsurface horizontal stresses of
           formations having different material properties. Soc. Petr. Eng. J. 21:658-662.
    Ramey, H J. 1967. Application of the line source solution to flow in porous media:
           A review. Producers Monthly. 4-27.
    Rice, J.R. 1968. A path independent integral and the approximate analysis of strain
           concentration by notches and cracks. J. Appl. Mech., Trans. ASME. 379-386.
    Rice, J.R. 1979. The mechanics of quasi-static crack growth. In: Proc. 8th U.S. Nat.
           Cong. Applied Mechanics. 191-216.
    Robinson, B.M., S.A. Holditch and W.S. Whitehead. 1988. Minimizing damage to a
           propped fracture by controlled flowback procedures. SPE Production
           Engineering. 40:753-759.
    Roegiers, J.C., J.D. McLennan and D.L. Murphy. 1982. Influence of preexisting
           discontinuities on the hydraulic fracturing propagation process. Hydraulic
           Fracturing and Geothermal Energy.  Nemat-Nasser, S., H. Abe and S.
           Hirakawa.(eds.) Martinus NijhoffPub. The Hague. 413-430.
    Rogowski, A.S., W.C. Moldenhauer and D. Kirkham. 1968. Rupture parameters of
           soil aggregates. Soil Sci. Soc. Am. Proc. 32:720-724.
    Ruina, A. 1978. Influence of coupled deformation-diffusion effects on the
           retardation of hydraulic fracture. In: Proc. U.S. Sym. Rock Mech., Rapid
           City, SD. 19:274-282.
    Rummel, F. and O. Kappelmeyer. 1982. The Falkenberggeothermal frac-project:
           Concepts and experimental results.  Hydraulic Fracturing and Geothermal
           Energy. Nemat-Nasser, S., H. Abe and S. Hirakawa.(eds.) Martinus Nijhoff
           Pub. The Hague. 59-74.
    Scheidegger, A.E. 1960. On the connection between tectonic stresses and well
           fracturing data. Geofisica Pura e Applicata. 46:66-76.
                                      276
    

    -------
    Scheidegger, A.E. 1965. Discussion of paper by P.E. Gretener, "Can the state of
          stress be determined from hydraulic fracturing data?". J. Geop. Res. 70:6213.
    
    Schlichting, H. 1960. Boundary Layer Theory. McGraw Hill. New York. 647.
    
    Schmidt, R.A. 1977. Fracture mechanics of oil shale: Unconfined fracture
          toughness, stress corrosion cracking, and tension test results. In: Proc. 18th
          U.S. Sym. Rock Mech. 1-6.
    Scott, P.P., W.G. Bearden and G.C. Howard. 1953. Rock rupture as affected by fluid
          properties. Petr. Trans. AIME. 198:111-124.
    
    Secor, D.T. and D.D. Pollard. 1975. On the stability of open hydraulic fractures in
          the earth's crust. Geophysical Research Letters. 2:510-513.
    
    Settari, A. 1980. Simulation of hydraulic fracturing processes. Soc. Petr. Eng. J. 487-
          500.
    
    Settari, A 1983. A new general model of fluid loss in hydraulic fracturing, SPE
          Paper 11625. In: Proc. SPE Sym. Low Permeability, Denver, CO. 219-228.
    
    Settari, A. 1985. Quantitative analysis of factors influencing vertical and lateral
          fracture growth, SPE Paper 13862. In: Proc. Low Perm. Gas Res. Sym.,
          Denver, CO.(May) 117-128.
    
    Settari, A. 1988. Quantitative analysis of factors influencing vertical and lateral
          fracture growth. SPE Production Engineering. 3:315-322.
    
    Settari, A. and M.P. Cleary. 1982a. Development and testing of a pseudo-three-
          dimensional model of hydraulic fracture geometry, SPE Paper 10505. In:
          Proc. 6th SPE Sym. Res. Sim., New Orleans, LA. 187-215.
    
    Settari, A. and M.P. Cleary. 19825. Three-dimensional simulation of hydraulic
          fracturing, SPE Paper 10504. In: Proc. 6th SPE Sym. Res. Sim., New Orleans,
          LA.  159-186.
    
    Settari, A. and M.P. Cleary. 1984. Three-dimensional simulation of hydraulic
          fracturing. J. Petr. Tech., Trans. AIME. 1177-1190.
    
    Settari, A. and M.P. Geary. 1986. Development and testing of a pseudo-three-
          dimensional model of hydraulic fracture geometry. SPE Production
          Engineering. 449-466.
    
    Settari, A. and H.S. Price. 1984. Simulation of hydraulic fracturing in low
          permeability reservoirs. Soc. Petr. Eng. J. 24:141-150.
    
    Sherard, J.L. 1973.  Embankment dam cracking. Embankment Dam Engineering.
          Hirschfeld, R.C. and S J. Poulos.(eds.) John Wiley & Sons. New York. 271-
          353.
                                      277
    

    -------
     Shockey, DA, K.C. Dao and R.L Jones. 1979. Effect of grain size on the static and
           dynamic fracture behavior of titanium. Mechanisms of Deformation and
           Fracture. Easterling, K.E.(ed.) Permagon Press. Oxford. 77-85.
    
     Shoji, T. and S. Takenouchi. 1982. Hydraulic fracturing in geological processes: A
           review. Hydraulic Fracturing and Geothermal Energy. Nemat-Nasser, S., H.
           Abe and S. Hirakawa.(eds.) Martinus Nijhoff Pub. The Hague. 175-190.
    
     Smart, P. and J.W. Dickson. 1979. Deformation and shear of normally consolidated
           flocculated kaolin. Mechanisms of Deformation and Fracture. Easterling,
           K.E.(ed.) Permagon Press. Oxford. 129-136.
    
     Smith, E. 1981. Some observations on the viability of crack tip opening angle as a
           characterising parameter for plane strain crack growth in ductile materials.
           Int. J. Fracture. 17:443-448.
    
     Smith, M.B., W.K. Miller, and J. Haga. 1987. Tip screenout fracturing: A technique
           for soft unstable formations. SPE Production Engineering. 2:95-103.
    
     Smith, S. 1989. Roqk fracturing methods: Their, development and use. Water Well
           J.(Feb.) 41-47.
    
     Sneddon, I.N. 1946. The distribution of stress in the neighborhood of a crack in an
           elastic,solid. Proc. Royal Soc. London. 187:229-260.
    
     Snyder, VA 1980. Theoretical Aspects and Measurement of Tensile  Strength in
           Unsaturated Soils. Ph.D. Cornell University. 94.
    
     Snyder, VA and R.D. Miller. 1985. Tensile strength of unsaturated soils. Soil Sci.
           Soc. Am. J. 49:58-65.
    
     Spence, DA and D.L Turcotte. 1985.  Magma-driven propagation of cracks. J.
           Geop. Res. 90:575-580.
    
     Stewart, G.W. 1974. Hydraulic fracturing of crystalline rock stimulates yield of two
           test wells drilled in New Hampshire. Ground Water. 12:46-47.
    
    Stewart, G.W. 1978. Hydraulic fracturing of drilled water wells in crystalline rocks
           of New Hampshire.  New Hampshire Dept. of Resources and Econ. Devel.,
           and Water Resources Research Center. New Hampshire. 161.
    
    Stierman, DJ. 1988. Geophysical monitoring of shallow hydraulic fractures,
           Unpublished draft report. Univ. of Toledo, Geology Dept.
    
    Stow, S.H., S. Haase and H.O. Weeren. 1985. Waste disposal by hydrofracture and
              ication of the technology to the management of hazardous wastes. In:
             jc. Int. Conf. New Frontiers Haz. Waste Manag. EPA/600/9-85/025:138-
          144.
                                      278
    

    -------
     Streeter, V.L. 1971. Fluid Mechanics. 5th McGraw-Hill Book Co. New York. 755.
    
     Strickland, F.G. 1985. Reasons for production decline in the diatomite, Belridge Oil
           Field: A rock mechanics view. J. Petr. Tech., Trans. AIME. 521-526.
    
     Strubhar, M.K., W.L. Medlin, S.M. Nabi and F.S. Andreani. 1984. Fracturing results
           in diatomaceous earth formations, South Belridge Field, California. J. Petr.
           Tech., Trans. AIME. 495-502.
    
     Sun, Y. and C. Ting. 1988. Introduction to a new apparatus for hydraulic fracturing
           tests. Geotech. Test. J. 11:288-292.
    
     Tada, H., P.C. Paris and G.R. Irwin. 1985. The Stress Analysis of Cracks Handbook.
           2nd Paris Productions Inc. St. Louis. 475.
    
     Tavenas, F.A., G. Blanchette, S. Laroueil, M. Roy and P. La Rochelle. 1975.
           Difficulties in the in-situ determination of Ko in soft sensitive clays. In: Proc.
           ASCE-GED Spec.  Conf. In-situ Meas. Soil Prop., Raleigh, NC 1:450-476.
    
     Telelman, A.S. and AJ. McEvily. 1967. Fracture of Structural Materials.  John
           Wiley & Sons, Inc.  New York. 697.
    
     Thiem, A. 1906. Hydrologische Methoden.  Gebhardt. Leipzig. 56.
    
     Thiercelin, M., R.G. Jeffrey and K.B. Naceur. 1989. Influences of fracture toughness
           on the geometry of hydraulic fractures. SPE Production Engineering. 4:435-
           442.
    
     Tyler, L.D. and W.C Vollendorf. 1975. Physical observations and mapping of cracks
           resulting from hydraulic fracturing in situ stress measurements, SPE Paper
           5542. In: Proc. 50th Ann. Fall Meeting SPE-AIME, Dallas, TX. 1-16.
    
     Vandamme, L., R.G. Jeffrey and J.H. Curran. 1988. Pressure distribution in three-
           dimensional hydraulic fractures. SPE Production Engineering. 3:181-186.
    
    van Eekelen, H.A. 1980. Hydraulic fracture geometry-Fracture containment hi
           layered formations, SPE Paper 9261. In: Proc. 55th Ann. SPE Fall Tech.
           Conf. Exh., Dallas,  TX. 1-9.
    
    van Poolen, H.K., J.M. Tinsley and CD. Saunders. 1958. Hydraulic fracturing:
          Fracture flow capacity vs. well productivity. Petr. Trans. AIME. 213:91-95.
    
    Veatch, R.W. 1983a. Overview of current hydraulic fracturing design and treatment
          technology-Part 1.  J.  Petr. Tech., Trans. AIME. 677-687.
    
    Veatch, R.W. 1983b. Overview of current hydraulic fracturing design and treatment
          technology-Part 2.  J.  Petr. Tech., Trans. AIME. 853-"
                                      279
    

    -------
     Vomocil, JA and WJ. Chancellor. 1967. Compressive and tensile failure strengths
           of three agricultural soils. Trans. ASAE. 771-774,779.
    
     Waltz, J. and T.L. Decker. 1981. Hydro-fracturing offers many benefits. Johnson
           Driller's Journal. 53:4-9.
    
     Waltz, J.P. 1988. Hydro-frac Basics. Ground Water Age. 22:26-29.
    
     Warpinski, N.R. 1983. Investigation of the accuracy and reliability of in-situ stress
           measurements using hydraulic fracturing in perforated cased holes. In- Proc
           24th U.S. Sym. Rock Mech., Golden, CO. 773-786.
    
     Warpinski, N.R. and P.T. Branagan. 1989. Altered-stress fracturing. J. Petr. Tech
           Trans. AIME. 41:990-997.
     Warpinski, N.R., J.A. Clark, R A Schmidt and C.W. Huddle. 1982. Laboratory
           investigation on the effect of in-situ stresses on hydraulic fracture
           containment. Soc. Petr. Eng. J. 333-340.
    
     Weertman, J. 1980. The stopping of a rising, liquid-filled crack in the earth's crust by
           a freely slipping horizontal joint. J. Geop. Res. 85:967-976.
    
     Weidmann, G.W. and W. Doell. 1979. Crazes and the fracture of glassy
           thermoplastics.  Mechanisms of Deformation and Fracture.  Easterling
           K.E.(ed.) Permagon Press. Oxford. 385-391.
    
     Williams, B.B. 1970. Fluid loss from hydraulically induced fractures. J. Petr Tech
           Trans. AIME. 882-888.
    
     Williams, J.G. 1984. Fracture Mechanics of Polymers.  John Wiley and Sons New
           York.  302.
    
     Williamson, W.H. and D.R. Woolley. 1980. Hydraulic fracturing to  improve the
           yield of bores hi fractured rock. Australian Water Resources Council
           Technical Paper. 55:1-76.
    
     Wolff, R.G., J.D. Bredehoeft, W.S. Keys and E. Shuter. 1975. Stress determination
           by hydraulic fracturing in subsurface waste injection. Water Tech. Res., J.
    
    
     Wong, H.Y. and I.W. Farmer.  1973. Hydrofracture mechanisms in rock during
           pressure grouting. Rock Mechanics. 5:21-41.
    
     Yuster, S.T. and J.C. Calhoun. 1945. Pressure parting of formations in water flood
           operations. The Oil Weekly.
    
     Zhang, Z.M. 1989. Application of hydrofracture principals to grouting in deep
           foundation. Pile Buck. 1-12.
    
    Zoback, M.D. and D.D. Pollard. 1978. Hydraulic fracture propagation and the
           interpretation of pressure-time records for in-situ stress determinations. U S
           Symposium of Rock Mechanics. Kim, Y.S.(ed.) Univesity of Nevada. Reno.
           14-22.
    
                                       280
    

    -------
    Zoback, M.D., F. Rummel, R. Jung and C.B. Raleigh. 1977. Laboratory hydraulic
          fracturing experiments in intact and pre-fractured rock. Int. J. Rock Mech.
          Min. Sri. Geomech. Abs. 14:49-58.
                                     281
    

    -------
                          APPENDIX
    RECORDS OF PRESSURE, INJECTED VOLUME, SURFACE UPLIFT AND
               SURFACE TILT AS FUNCTIONS OF TIME.
                            282
    

    -------
        150
    
    *«  125
    
        100
    
         75
    
         50 £
    
         25
    £
    OJ
           E
           E
           8
           O
                Well 14
                Begin 15:28
    
                • Estimated
                 4
    
    
                 3
    
    
                 2
    
    
                 1
    
    
                 0*
                                   0
    
    
                                   5
                                          Time (minutes)
                                                   10
    Well * 4
    Begin 15
    
    O 6' east of well
    A at well
                                                10
    
    
                                                15
        A
    /     \
                                      A       x°OO\
                                   A         O        OO0
    
    
                                O  O  O
                                                                                      a
                                                                                     •S
                                                                                      o
    
                                                  10
                                                                   15
                                           283
    

    -------
                                                    10
    /u
    *S! e/\
    Q. wU t
    £ 50
    3
    0>
    | 40
    oZ
    c 30
    o
    1 20
    •E1
    10
    fit
    Pressure
    A A
    
    „ . -
    
    Well #5
    Begin 16:50
    .
    
    • Estimated
    ^ — -^^"^
    	 ^O — Volume
    i i mr* i 	 > 	 1 	 > 	 • 	 • 	 1 	 '•
    /w
    60
    WV/
    ^%
    50 "o
    W1
    ^^^
    ^^^
    40 |
    30 |
    
    20
    
    10
    0
                            Time (minutes)
                                                    10
         1000
               Well #5
    
              . Begin 16:53.5
        -2000
                          Time (minutes)
     0)
    
    
     i
    "5.
    
    a
    

    -------
              Well #6
    
              Begin 18:34.5
    -1500
                              Time (minutes)
                 06* east of well
                 A at well
                          285
    

    -------
        100
     CO
     Q.
     s
     0)
     c
     o
     *=
     O
     
    o
    J3
    a.
    M
    
    5
      10
    Well
    
    Begin 12:57
    o 61 west of well
    • at well
    
    A 6' east of well
               O '60-0-0-0 •
                    5          10          15
    
    
                            Time (minutes)
    
    
                                286
    

    -------
    
    /^N
    *W
    Q.
    Q>
    GO
    «0
    £
    Q_
    C
    to
    "o
    —
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    
    150
    125
    
    100
    
    
    75
    50 ^
    25
    0<
    ' Well #4 /°
    Begin 15:28 /. .
    ' / Volume -
    : /
    /V°
    /
    25
    0
    0 5 10 15 20
    Time (minutes)
    4
    
    _ *
    ^^ _
    +* 2
    ^^v *••
    o>
    ^i^
    «
    o 1
    o
    Q-
    0)
    *° 0.
    u t
    1
    
    ,....,....,..
    : wen #4 /A\ ;
    • Begin 15:30 * A
    1 ^ ^^ '
    
    A Ai
    /
    A OOO
    / / \
    A O OCX
    //
    ./
    ^A O
    
    
    
    
    
    
    i
    
    
    >
    
    
    
    
    
    5           10
         Time (minutes)
         287
    15
    

    -------
    /u
    en
    oU ^
    *w
    ,5 50
    CO
    1 40
    CO
    
    288
    

    -------
    3
    CO
    CO
    0>
    o
    =8
    a>
    175
    
    
    150
    
    
    125
    
    
    100
    
    
     75
    
     50
    
    
     25
                                                           1 175
          06
            0
                Well
                Begin 18:30
                                 10
                          Time (minutes)
                                                                  o
                                                                  o>
    a>
    
    ^_^
    £
    £
    "c
    0)
    F
    k*
    o>
    o
    o
    n.
    CO
    b
    
    
    O
    5
    
    4
    3
    
    2
    
    
    OS
    — 1
    
    • We!! #6
    Begin 18:35
    •
    .
    
    •
    * '^^
    & ^ A
    .X ex
    r—fr — UU U
    1 . i
    A
    J\
    oi
    f \
    1 «
    • * * '
    | QA
    4 oo-
    /]
    o
    
    1 . a
                                      5            10
                                   Time (minutes)
                                289
    

    -------
    
    I
    0)
    0)
    V
    o
    1
    (0
    90
    Of\
    80
    70
    60.
    50
    40
    30
    20
    4 f\
    10
    ft*
    _ 	 , 	 , 	 , 	 , 	 1 	 1 	 1 	 1 	 1 	 1 	 1 	 1 	 « 	 1 ' ;
    ; ^^^^"^ Pressure
    m ^^^^^^^^
    i-"""" 	
    . Well #7
    Begin 18:12
    :
    : Volume
    L^ — i — , — i — . — ft^- — « • ' — • — « — • — '
    90
    80
    70
    60 ?
    QI
    50 7
    40 |
    30 ^
    20
    10
    0
                              345
    
                              Time (minutes)
    8
                        r ^
    
    
    
                        E
                        0>
    
    
                        0>
                        o
                        g
    
                        Q.
                        M
                                 Well #7
    
                                 Begin 18:15
                                     1234
    
                                       Time (minutes)
                                    290
    

    -------
     CO
     0.
    £
    3
    m
    m
    i_
    Q_
    C
    O
    9)
    550
    500
    450
    400
    350
    300
    250
    200
    150
    100
    50'
    0<
    , i i i i ••|»«-«|«-««|«...|.»..j....|,..
    : Well #8 A, -
    • Begin 17:07 / A-
    / ,
    A :
    /
    1
    Pressure L i
    /^ i
    / j
    A ._— ._ 	 A. -
    
    »^ \ff\\i tnr\o '•
    
    550
    500
    450
    400
    350
    300
    250
    200
    150
    100
    50
    n
    O
    0>
     !
    o —l
    O
    Q.
    .2
    ^^
    0 -2
    3
    
    Well #8
    • Begin 17:12
    
    
    
    AAA— A— — AAA— A— AAA A— A 0Wh
    _^_ M — i • M i u &* KJU
    
    
    o 6' east of well
    * at well
    
    
    0 5 10 15 20 25 30 35 40 45
    Time (minutes)
    291
    

    -------
    150
    '55   125
    Q.
    100
            Well #10
            Begin 14:44
                                                          5)150
                                                      - 125
                              10
                         Time (minutes)
    
    
    T
    t,
    -M
    q>
    
    o
    o
    *0
    0)
    0
    J
    2
    
    1
    
    Di
    Ul
    
    
    -1
    
    -2
    3
    Well #10
    Begin 14:46
    
    K
    /\
    
    
    ^v
    X^^
    N.
    o 6' west of well >.
    ^ at well \
    A
                       5           10
                           Time (minutes)
                                                  15
                               292
    

    -------
    200 -
           Well #9
           Begin 16:12
                      5              10
                         Time (minutes)
    
    
    
    KM
    
    "c
    Q>
    E
    0)
    o
    o
    Q.
    Of)
    5
    
    
    
    7
    6
    
    
    5
    4
    
    
    3
    
    
    2
    
    
    1
    
    Oi
    i
    • Well #9
    Begin 16:12
    
    
    ; o 6' west of well
    L * at well
    
    
    
    
    
    
    
    
    
    ^A-A— A— A— A^.^ A fc
    i 	 ChVo— O O O_
    
    A i
    AO A
    | / \ \
    I/ °\ A\
    // O A~— A
    b o
    if
    I/
    if •
    H
    •
    1
    1
    f
    O
    ^7
    
    — o-o :
                                    10
                       Time (minutes)
                            293
    15
    

    -------
    CD
    Q.
    3
    m
    m
    Q)
    c
    o
    tp
    a
    a>
    90
    80
    70
    60
    50
    40
    30
    20
    10
             Well
             Begin 11:16
    ' Estimated
                                       Volume
                                        Pressure
    90
    80
    70
    60
    50
    40
    30
    20
    10
    0
                            Time (minutes)
                                 294
    

    -------
    400
     « 350
     o.
    ^ 300
     | 250
     Q. 200
     c •
    S 150
    I 100
        50
                Well #13
                Begin 10:15
            Pressure
                                        Volume
                           10        15
                       Time (minutes)
    u
    4
    
    1* ^
    1 2
    
    "c 1
    Q)
    E Oc
    § u<
    1-1
    Q
    -3
    -4
    -5
    •j\>
    
    O O
    • f»
    1
    7
    /
    
    Well #13
    Begin 10:15
    
    
    \ :
    \
    o
    \
    \
    0
    
    \ •
    \ :
    0 -
    	 ,
                           10        15
                        Time (minutes)
                            295
                                                  20
    

    -------
    

    -------