105
Rock Mechanics Laboratory Tests Relevant to the in-situ Stress Measurements at the 4850-level, Homestake Mine, Lead, South Dakota by Peter Vigilante A thesis submitted in partial fulfillment of the requirements for the degree of Master of Science (Geological Engineering) at the UNIVERSITY OF WISCONSIN-MADISON 2017

by Peter Vigilante A thesis submitted in partial

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: by Peter Vigilante A thesis submitted in partial

Rock Mechanics Laboratory Tests Relevant to the in-situ Stress

Measurements at the 4850-level, Homestake Mine, Lead, South Dakota

by

Peter Vigilante

A thesis submitted in partial fulfillment of

the requirements for the degree of

Master of Science

(Geological Engineering)

at the

UNIVERSITY OF WISCONSIN-MADISON

2017

Page 2: by Peter Vigilante A thesis submitted in partial

i

Abstract

Laboratory rock mechanics experiments were conducted on cores taken from the

Sanford Underground Research Facility (SURF) (formerly of Homestake Mine) in Lead,

South Dakota. Strength properties and their anisotropy were investigated in order to aid the

interpretation of the hydraulic-fracturing stress measurements conducted in boreholes drilled

down from the 4850’ level (1478 meters depth) of the mine. Brazilian disc tests, uniaxial and

triaxial tests, and laboratory hydraulic fracturing tests were conducted on cores recovered

from the kISMET (Permeability (k) and Induced Seismicity Management for Energy

Technologies) project as well as those previously obtained as part of the Deep Underground

Science and Engineering Laboratory (DUSEL) program. Brazilian disc tests revealed

significantly higher tensile strength when samples were loaded normal to the foliation

compared to when the samples were loaded parallel to the foliation, indicating clear tensile

strength anisotropy. Uniaxial tests of the cores produced UCS results with an average value

of 107.4 MPa for samples loaded parallel to foliation, and 88.1 MPa for samples loaded

perpendicular to foliation. Triaxial dynamic moduli measurements conducted under

confinement equivalent to in-situ stress magnitudes indicated a 18% increase in dynamic

Young’s modulus between samples loaded parallel to foliation compared to perpendicular.

Laboratory hydrofracture tests were conducted under various triaxial stress states to see

whether potential shear failure along weak foliation planes, promoted by differential stress,

could affect the apparent hydrofracture breakdown pressure. While the observed breakdown

pressures did not indicate any clear influence of the rock anisotropy, some samples subject to

Page 3: by Peter Vigilante A thesis submitted in partial

ii

differential stress showed lower breakdown pressure possibly caused by leak-off into

fractures before peak pressure is reached.

Page 4: by Peter Vigilante A thesis submitted in partial

iii

Acknowledgements

I would first like to thank my advisor, Hiroki Sone. Being his first graduate student,

he showed immense patience and support throughout my two years at the University of

Wisconsin-Madison. He was always willing to take the necessary time out of his schedule to

help me with my classwork, research, and teaching. I would not have been able to complete

my Master’s without his support. I would also like to thank my other committee members,

Herb Wang and Bezalel Haimson. They both were incredible resources for my research, and

were always willing to help out when needed. I would also like to thank Seiji Nakagawa for

all his help on laboratory experiments.

I would also like to thank both the Geological Engineering and Geoscience

departments at the University of Wisconsin-Madison. I spent endless hours in both

Engineering Hall and Weeks Hall, and both departments helped me complete my degree.

Also, would like to thank the Department of Energy and the kISMET project for financial

support of my research.

Finally, endless thanks to my friends and family for the support the last two years.

My family, although far away, was always there for me when I needed them in some stress-

filled times. Friends and colleagues in the graduate school also made it a memorable two

years in Madison. Specifically, Ben Heinle and Elliott Andelman, who kept me sane at all

times.

Page 5: by Peter Vigilante A thesis submitted in partial

iv

Contents

Abstract ..................................................................................................................................... i

Acknowledgements ................................................................................................................ iii

Contents .................................................................................................................................. iv

List of Figures ...........................................................................................................................v

List of Tables .......................................................................................................................... ix

1. Introduction ..........................................................................................................................1

2. New Laboratory Measurements of Rock Properties at kISMET Field Site...................2

2.1 Brazilian Disc Test ...................................................................................................5

2.2 Triaxial Dynamic Moduli Test...............................................................................19

2.3 Uniaxial Moduli and Unconfined Compressive Strength (UCS) Test ...................24

2.4 Laboratory Hydrofracture Test ..............................................................................36

3. Conclusions and Future Work ..........................................................................................52

4. References ...........................................................................................................................55

Appendix .................................................................................................................................59

Page 6: by Peter Vigilante A thesis submitted in partial

v

List of Figures

Figure 1. a) Brazilian Disk sample preparation schematic b) Brazilian Disk schematic

representing post experiment. Black lines show rock foliation, and red line shows tensile

fracture created from the testing c) Displacement vs Load Brazilian Test results from (Wang

and Xing, 1999) ........................................................................................................................6

Figure 2. Grouped Brazilian Disc Samples plotting Normalized Load (kN/cm) vs axial

displacement (mm). The schist samples were grouped together according to what core run

and associated depth they were recovered from, as well as quartz and rhyolite samples

grouped together separately .......................................................................................................8

Figure 3. Post testing images of two separate schist Brazilian disc samples. On the left,

sample 41A2 shows the expected vertical fracture (tensile crack) within the disc. On the

right, sample 65B3 shows unexpected fracture angled away from the axial load (which is

vertical in both photos) ............................................................................................................10

Figure 4. Plot representing the angle formed between the vertical axial load of the Brazilian

Disc test and the foliation of the disc sample vs the tensile strength of each sample clarifies

(MPa). Solid blue data points represent true tensile strengths of samples, while blue unfilled

data points represent lower bounds of those samples tensile strength. All samples are

Poorman Schist, and raw data are shown in Table 2 ...............................................................13

Figure 5. Anisotropy of Ultimate Strength in Uniaxial Direct-Pull Tension of Lyons

Laminated Sandstone. (Kwaśniewski, 2009; original data from Youash, 1966) .....................17

Page 7: by Peter Vigilante A thesis submitted in partial

vi

Figure 6. Anisotropy of Ultimate Strength in Uniaxial Direct-Pull Tension of Morrow Point

mica schist. T is the tensile strength (MPa) and B is the Angle Between the Loading

Direction and the Samples Foliation from (Kwaśniewski, 2009; original data from Olsen,

1967) ........................................................................................................................................18

Figure 7. Seismic P-wave velocity measurements taken (a) around the circumference of

samples and (b) along the length of samples from the kISMET field site. Velocity anisotropy

is apparent for the entire length of the cores from (Oldenburg et al., 2016) ...........................23

Figure 8. Stress vs Strain curves for all eight Schist triaxial samples tested. Note: no radial

strain data was acquired for sample 56H (bottom left) prior to bringing sample to failure

during testing. Also, “depth” is depth in feet from the 4850’ level of the Homestake Mine ..29

Figure 9. Images of triaxial samples taken after they were brought to failure in UCS testing.

Samples with foliation normal to sample axis are on the top row, and samples with foliation

relatively parallel to sample axis are the bottom row ..............................................................30

Figure 10. Anisotropy of Unconfined Compressive Strength (Kwaśniewski 2009) ...............33

Figure 11. Dependence of Differential Stress (y-axis) at Shear Failure on the Orientation of

the Weakness Plane (x-axis) in a Sample of Anisotropic Rock Model from (Paterson and

Wong, 2005; Produced by Jaeger, 1960) .................................................................................35

Figure 12. (above) Cross-section of lab hydrofracture sample and (below) detailed

description of stainless steel high-pressure tubing inserted into lab hydrofracture samples

(also seen in Figure 13). O-ring was attached to tubing pieces using the o-ring groove, and

Page 8: by Peter Vigilante A thesis submitted in partial

vii

then the tube was epoxied into the 0.125” holes of the samples to seal the hydrofracture

testing zone to the high-pressure tubing in the system ............................................................37

Figure 13. Schematic of lab hydrofracture set up. Not depicted to scale, but dimensions of

hydrofracture sample and tubing detailed ................................................................................38

Figures 14-20. Schematic of lab hydrofracture set up. Not depicted to scale, but dimensions

of hydrofracture sample and tubing detailed ..................................................................... 41-44

Figure 21. Axial Stress (Max. Principal Stress) Applied to Hydraulic Fracturing Samples vs

The Breakdown Pressure of that Sample and Fracture Reopening Pressure for Samples

Recovered from Core Run 64 ..................................................................................................45

Figure 22. Axial Stress (Max. Principal Stress) Applied to Hydraulic Fracturing Samples vs

The Breakdown Pressure and Fracture Reopening Pressure of that Sample for Samples

Recovered from Core Run 53 ..................................................................................................46

Figure 23. Reproduced with Expected Breakdown Pressures (Pb) shown as Overlaid Red

Box ...........................................................................................................................................49

Figure 24. Sample 53-B Pressure vs Time Plot Showing Pressure Increase in Test Zone

Deviates from Linearity (Red Line) to Help Explain Outlier Breakdown Pressure Value .....50

Figure 25. Map View of Cross Section A to A’ (red line) which is shown in cross sectional

view in Figure 26 From (ARUP, 2015) ...................................................................................60

Figure 26. Cross sectional view and formations of the Homestake mine. 4850L represented

by solid red line. From (Denver Region Exploration Geologists’ Society) .............................61

Page 9: by Peter Vigilante A thesis submitted in partial

viii

Figure 27. General Geology Overview of 4850L. Blue Dashes are General Foliation

Orientation and Red Dashes are Rhyolite Dikes. From (Golder Associates Final, 2010a) .....62

Figure 28. Range of UCS values for rock types at 4850L from (Lachel Felice, 2009) ...........69

Figure 29. Comparison of Ultimate Joint Shear Strength (Represented as Red Triangle Data

Points) to Intact Rock Uniaxial Compression and Brazilian Tensile Strength (Represented as

Blue and Green Curves, Respectively) From (RESPEC 2010) ...............................................73

Figure 30. Image of borehole breakout at 4850 level of Homestake Mine. From (Golder

Associates, 2010a) ...................................................................................................................84

Figure 31. Inclination of Breakout in Borehole Drills from the Ventilation Drift (Borehole

Advanced at an Azimuth of Approximately 310°) from (Golder Associates, 2010b) ............85

Figure 32. Vertical Stress Gradients vs Depth in the Homestake Mine, SD. Sources from

Pariseau (1985), Johnson et al. (1995), and Golder Associates (2010a) .................................88

Figure 33. Stress Data Points from Bond (1970), Hooker et al. (1972), Johnson et al. (1995)

and Golder Associates (2010a) at Respective Depths, as well as Gradients from Pariseau

(1985) .......................................................................................................................................89

Figure 34. Estimated Minimum Borehole Pressure Needed at 4850L for Hydraulic Fracturing

(1985) .......................................................................................................................................92

Page 10: by Peter Vigilante A thesis submitted in partial

ix

List of Tables

Table 1. Record of all arbitrary sample names, borehole sample was recovered from,

borehole size, what test was conducted on the sample, the depth along the recovered core

length the sample was taken from, as well as a brief description of the core at the recovered

depth. (All samples are Poorman schist unless designated *=quartz samples and **=rhyolite

samples) ....................................................................................................................................3

Table 2. Brazilian Disc Samples tested. * indicates Schist samples. “Qtz” indicates samples

are quartz. “Rhy” indicates samples are rhyolite samples. “>” indicates samples which did

not fracture vertically and centered, thus not reaching the samples true maximum tensile load

..................................................................................................................................................11

Table 3. Breakdown of average normalized load (kN/cm) for the angle between foliation and

vertical axial load for 0°, 45°, and 90° .....................................................................................14

Table 4. P-wave and S-wave velocities for triaxial samples tested, calculated dynamic

moduli, and UCS data (discussed in the next section). No velocity data were recorded for

sample 56H prior to breaking the sample. M (P-wave modulus), G (Shear), were calculated

from acquired velocity measurements. ....................................................................................21

Table 5. All uniaxial samples with UCS values along with descriptions ...............................25

Table 6. Average UCS for samples of perpendicular foliation. “V” Samples have foliation-

parallel to cylindrical sample axis. “H” Samples have foliation-normal to the cylindrical

sample axis. ..............................................................................................................................26

Page 11: by Peter Vigilante A thesis submitted in partial

x

Table 7. Calculated Young’s Modulus and Poisson’s Ratio for all eight triaxial samples. Both

calculated using stress vs strain curves provided in Figure X (the stress strain curves) .........27

Table 8. Raw data for the successful lab hydrofracture tests. All tests had a minimum

principal stress of 22 MPa, and the fracture reopening pressures are an estimate due to

multiple fracture re-openings for the tests ...............................................................................46

Table 9. Average unconfined compressive strength (UCS) values for rock types at 4850 level

from (Lachel Felice, 2009) ......................................................................................................68

Table 10. Original Anisotropic Rock Property Data (*C=Compressive Strength (UCS),

T=Tensile Strength; The 1- and 3-directions are parallel to the schistosity; the 2-direction is

perpendicular to the schistosity. From (Pariseau, 1985) .........................................................70

Table 11. Modified Anisotropic Rock Property Data (*C=Compressive Strength (UCS),

T=Tensile Strength; The 1- and 3-directions are parallel to the schistosity; the 2-direction is

perpendicular to the schistosity). From (Pariseau, 1985) ........................................................70

Table 12. Compressive and Tensile Strength for Yates amphibolite and Rhyolite from

(RESPEC 2010) .......................................................................................................................71

Table 13. Comparison of Intact and Residual Mohr-Coulomb Strength Parameters for Joints

in Amphibolite from (RESPEC 2010) .....................................................................................72

Table 14. Modified Anisotropic Rock Property Data (*The 1- and 3-directions are parallel to

the schistosity; the 2-direction is perpendicular to the schistosity. All units are labeled MPa

except for Poisson’s ratio (unitless)). E=Young’s Modulus; v= Poisson’s Ratio; G= Shear

Modulus. from (Pariseau, 1985) ..............................................................................................75

Page 12: by Peter Vigilante A thesis submitted in partial

xi

Table 15. In-situ stress data at Homestake Mine, including data sources, stresses, magnitudes,

and directions based on Bond (1970), Hooker et al. (1972), and USBM (1984) ....................77

Table 16. Principal stresses, orientations, and elastic properties (SI units) where SM-02 to

SM-07 was in Yates amphibolite unit and SM-08 and 09 are results from rhyolite dike. From

(Golder Associates, 2010a) ......................................................................................................80

Table 17. Vertical and horizontal components, octahedral, and deviatoric stresses (SI Units)

where SM-02 to SM-07 was in Yates amphibolite unit and SM-08 and 09 are results from

rhyolite dike along 4850L. From (Golder Associates, 2010a) .................................................81

Table 18. Summary of In Situ Stress Measurements from (Golder Associates, 2010b) .........82

Table 19. Data used in Equation 9 to Determine Estimated Minimum Borehole Pressure

Needed at 4850L for Hydraulic Fracturing ..............................................................................90

Table 20. Estimated Minimum Borehole Pressure Needed at 4850L for Hydraulic Fracturing.

..................................................................................................................................................91

Page 13: by Peter Vigilante A thesis submitted in partial

1

1. Introduction

The primary section of this report details the laboratory tests run on samples from the

4850-foot level (4850L) of the Homestake Mine, South Dakota, in contribution to the

collaborative research project kISMET (Permeability (k) and Induced Seismicity

Management for Energy Technologies). The four tests conducted for this report were the

Brazilian disc tensile strength test, triaxial dynamic moduli test, uniaxial moduli and

unconfined compressive strength (UCS) test, and the laboratory hydrofracture test for peak

breakdown pressure under different in-situ lab conditions (Vigilante et al., 2017). The main

focus of these laboratory tests was to observe the effect of anisotropy on the strength of the

rocks recovered from the kISMET field site. These results will help support conclusions

made from the kISMET stress measurements, made at the 4850L in August, 2016 (Wang et

al., 2017). These laboratory results may also be helpful in interpretation of the in-situ stress

state from wellbore failures and future field measurements made at the Homestake Mine.

The appendix of this thesis, is a critical review prepared for the kISMET project. It is

a compilation of previous rock mechanics tests and in-situ stress measurements relevant to

the Sanford Underground Research Facility (SURF) in the Homestake Mine, South Dakota.

The goal of the kISMET field work at SURF was to conduct in-situ hydraulic fracturing

experiments which allowed for interpretation of the stress field (Haimson, 1978), learning

how the local rock fabric affects hydraulic fracturing, and to use geophysical methods to

locally monitor the hydraulic fracturing process at close distance compared to typical field

operations. Results from the August, 2016 kISMET field work can be found in Oldenburg et

Page 14: by Peter Vigilante A thesis submitted in partial

2

al. (2016) and Wang et al. (2017). The critical review was prepared to aid the design of these

field experiments.

2. New Laboratory Measurements of Rock Properties at kISMET Field

Site

In conjunction with the field testing, I conducted laboratory rock mechanics tests on

the cores retrieved from the kISMET field site (Homestake Mine, SD). The intent of this

portion of the project was to measure the retrieved rock’s properties and influence of

anisotropy in order to aid the interpretation of the hydraulic-fracturing stress measurements.

The borehole (kISMET-003) where the cores were recovered penetrated Poorman Formation,

which is a dark grey banded foliated phyllite to schist formation, and described in more detail

in Section 2 of this thesis (Oldenburg et al., 2016). Furthermore, kISMET-003 is the borehole

in which hydraulic fracture stress measurements were conducted (Wang et al., 2017). Quartz

veins, ranging in size from a few millimeters to centimeters, were also apparent sporadically

throughout the recovered core. The strong texture detailed from the recovered core suggests

the presence of significant mechanical and strength anisotropy of the rocks in the Poorman

Formations.

The four laboratory tests I conducted are detailed in this section of the report.

Brazilian disc test for tensile strengths, triaxial tests for dynamic moduli, uniaxial tests for

static moduli and unconfined compressive strengths, and hydraulic fracturing tests for

hydraulic breakdown pressure. The four lab experiments used cores recovered from the

Page 15: by Peter Vigilante A thesis submitted in partial

3

kISMET field site, as well as cores recovered from the Deep Underground Science and

Engineering Laboratory (DUSEL) program, a previous project which took place at the 4850’

level of the Homestake Mine. A main emphasis for the Brazilian disc and triaxial and

uniaxial tests was on how the orientation of the foliation in the schist relative to the loading

direction affects the respective strengths of the rocks. As for the lab hydrofracturing test, the

primary goal was to perform a laboratory stimulation of the breakdown pressures and

fracture reopening pressures of all the samples.

Due to multiple rock tests conducted on rocks recovered from the same field site,

Table 1 contains the arbitrary names given to all the samples, the borehole the sample was

recovered from, the depth in the borehole the sample was taken from (not the depth in the

mine), what test was conducted on the sample, and a brief description of the core where the

sample was taken from. The arbitrary names often include a number which corresponds to a

core run of the borehole it was recovered from. However, for future referencing the depth at

which the sample was taken from is recorded.

Table 1. Record of all arbitrary sample names, borehole sample was recovered from,

borehole size, what test was conducted on the sample, the depth along the recovered

core length the sample was taken from, as well as a brief description of the core at the

recovered depth. (All samples are Poorman schist unless designated *=quartz samples

and **=rhyolite samples)

Arbitrary

Sample

Name

Borehole

Sample was

Recovered

from

Diameter

of

Borehole

Test

Conducted

on Sample

Depth

(along

core

length) Description

56A1

kISMET-

003 NQ2

Brazilian 278’

No intense microfolding.

Plenty of brown/gold

Page 16: by Peter Vigilante A thesis submitted in partial

4

56A2

kISMET-

003 NQ2

Brazilian 278’

coloration. Foliation

consistent throughout.

56B1

kISMET-

003 NQ2

Brazilian 278’

56B2

kISMET-

003 NQ2

Brazilian 278’

41A2

kISMET-

003 NQ2

Brazilian 202.5’

Lighter portion of the core.

Whitish hue.

Intense folding and possible

quartz veins present.

41C1

kISMET-

003 NQ2

Brazilian 202.5’

41C3

kISMET-

003 NQ2

Brazilian 202.5’

41C2

kISMET-

003 NQ2

Brazilian 202.5’

63A1

kISMET-

003 NQ2

Brazilian 312.5’

Not as apparent foliation.

Homogenous looking portion

of the core.

63A3

kISMET-

003 NQ2

Brazilian 312.5’

63B2

kISMET-

003 NQ2

Brazilian 312.5’

65B1

kISMET-

003 NQ2

Brazilian 320’

Darker gray portion of the

core.

Intense microfolding.

65C3

kISMET-

003 NQ2

Brazilian 320’

65C2

kISMET-

003 NQ2

Brazilian 320’

65B3

kISMET-

003 NQ2

Brazilian 320’

Qtz2* DUSEL J HQ Brazilian N/A Pure quartz vein.

No foliation to test

anisotropy.

Qtz4* DUSEL J HQ Brazilian N/A

Qtz5* DUSEL J HQ Brazilian N/A

rhy2** DUSEL J HQ Brazilian 226.5’ Pure rhyolite portion of core.

Whitish-gray with darker

specks.

No foliation to test

anisotropy.

rhy3** DUSEL J HQ Brazilian 226.5’

rhy4** DUSEL J HQ Brazilian 226.5’

rhy5** DUSEL J HQ

Brazilian 226.5’

41V

kISMET-

003

NQ2 Triaxial

and

Uniaxial 204’

Lighter portion of core.

Intense folding. 41H

kISMET-

003

NQ2 Triaxial

and

Uniaxial 204’

49V

kISMET-

003

NQ2 Triaxial

and

Uniaxial 242.5’

Foliation is folding but

consistent. Dark bands in

foliation.

Page 17: by Peter Vigilante A thesis submitted in partial

5

49H

kISMET-

003

NQ2 Triaxial

and

Uniaxial 242.5

56V

kISMET-

003

NQ2 Triaxial

and

Uniaxial 279.5’

Foliation relatively

perpendicular to axis. Not

much folding. 56H

kISMET-

003

NQ2 Triaxial

and

Uniaxial 279.5’

63V

kISMET-

003

NQ2 Triaxial

and

Uniaxial 313’

Foliation relatively

consistent. Normal Poorman

grayish color. 63H

kISMET-

003

NQ2 Triaxial

and

Uniaxial 313’

64-E

kISMET-

003

NQ2 Lab

Hydrofrac 318.5’

Foliation ~20° off vertical.

No intense folding.

64-C

kISMET-

003

NQ2 Lab

Hydrofrac 319’

Some folding, but foliation

about 10°-20° off vertical

53-C

kISMET-

003

NQ2 Lab

Hydrofrac 262.3’

Intense microfolding.

Foliation relatively normal to

vertical axis.

53-D

kISMET-

003

NQ2 Lab

Hydrofrac 262’

Intense microfolding. Shiny

white/silver areas.

53-B

kISMET-

003

NQ2

Lab

Hydrofrac 262.6’

Intense microfolding.

Foliations goes from

horizontal to vertical

orientation.

531111-A

kISMET-

003

NQ2 Lab

Hydrofrac 263.3’

Not as clear foliation.

Foliation dips ~10°-40°

531-A

kISMET-

003

NQ2 Lab

Hydrofrac 264’

Not as clear foliation.

Foliation dips ~10°-40°

2.1 Brazilian Disc Test

Methods - Brazilian Disc Test

The initial laboratory test conducted on the Homestake Mine cores was the Brazilian

Disc test. To prepare the Brazilian disc samples, the first step was to core 2.54 cm diameter

cylinders from the recovered rock samples. The 2.54 cm. specimens were cored parallel to

Page 18: by Peter Vigilante A thesis submitted in partial

6

the general foliation direction on both HQ sized core (63.5 mm or 2.5 in. diameter core) and

NQ2 sized core (50.5 mm or 1.99 in diameter core). These 2.54 cm diameter and 5.08 cm –

6.35 cm length cores were then sliced into disc shaped test specimens using a rock saw

(depicted in red in Figure 1a). The suggested thickness to diameter ratio for the Brazilian

Disc test is 0.5 to 0.6 (Ulusay and Hudson, 2007). Due to constraints on the amount and size

of the rock cores, the average thickness of the 22 Brazilian disc samples we tested was 1.14

cm or 0.45 in.

Figure 1. a) Brazilian Disk sample preparation schematic b) Brazilian Disk

schematic representing post experiment. Black lines show rock foliation, and red line

shows tensile fracture created from the testing c) Displacement vs Load Brazilian Test

results from (Wang and Xing, 1999)

The Brazilian disc test allows the study of anisotropy effects by varying the angle

between the force applied on the discs and the foliation within the rock disc. This is depicted

in Figure 1b, where the angle between the diametrically applied force on the disc shaped

sample and the foliation in the rock can be varied between 0°-90°. The “Triaxial Platens” are

Page 19: by Peter Vigilante A thesis submitted in partial

7

the surfaces that the sample sits on (the bottom platen) and the surface that exerts the axial

load on the sample (top platen). An example of how the Brazilian disc test allows the study

of anisotropy is that a sample with foliation parallel to the vertical axial load of the testing

apparatus would have 0° angle between axial load direction and foliation, and a sample with

foliation normal to the vertical axial load would have a 90° angle between axial load

direction and foliation. With the recovered kISMET cores, as well as the previously drilled

DUSEL cores, we were able to produce groups of Brazilian Disc samples from four separate

depths in the Poorman Formation (foliated phyllite to schist), one group of disc samples from

the quartz veins, and one more group of samples from the DUSEL cores which penetrated a

rhyolite dyke. In total, 22 Brazilian disc samples were tested. A GCTS RTR-1000 Triaxial

Testing System was used for all of the Brazilian Disc testing. Force-displacement data were

recorded during the experiment and failure of the sample was identified as the frame load

dropped suddenly to a residual value. After the destructive testing, samples were then

observed to identify the failure plane.

Results - Brazilian Disc Test

Plots of the Brazilian disc tests conducted in this report are in Figure 2. The results

are presented in a Normalized load (kN/cm) vs piston displacement (mm) curve for all of the

samples. The normalized load is defined by the force from the triaxial machine imposed on

the disc sample divided by the thickness of the sample. Calculating the normalized load was

necessary for eventually calculating tensile strength using the Brazilian test formula. This

formula calculates the tensile strength (using the parameters peak applied load (P),

Page 20: by Peter Vigilante A thesis submitted in partial

8

diameter of the disc sample (D), and thickness of the sample (L), and is shown below in

equation 1 (Ulusay and Hudson, 2007):

𝜎𝑡 = 2 ∗ 𝑃/(𝜋 ∗ 𝐷 ∗ 𝐿) (𝟏)

Figure 2. Grouped Brazilian Disc Samples plotting Normalized Load (kN/cm) vs axial

displacement (mm). The schist samples were grouped together according to what core

run and associated depth they were recovered from, as well as quartz and rhyolite

samples grouped together separately.

Page 21: by Peter Vigilante A thesis submitted in partial

9

These load-displacement curves were then grouped together by rock type and depth

of sample recovery. Core runs 41, 56, 63, and 65 are all Poorman Schist sample groups, and

the Quartz group and Rhyolite group samples are separate. The load-displacement curve is a

good indication that the elastic stiffness (represented by the slope of the curve) is anisotropic.

This anisotropic effect can be seen clearly in CoreRun 56. In CoreRun 56 the Schist seems

stiffest loaded parallel to foliation (0°), and then decreases in stiffness as the angle between

the vertical axial load and the foliation increases to 30°, 60°, and then 90°. These load

displacement curves are also necessary in calculating the tensile strength, because the curves

give us the peak applied load (or P in equation 1 above).

Figure 3 below includes two of the Poorman schist Brazilian disc samples post

failure. These images show the difference in how the Brazilian disc samples fractured. The

sample on the left, 41A2, fractured vertically almost perfectly along the axial load axis. This

type of fracture indicates the sample fractured in tension so that the peak load represents the

true tensile strength of the sample. However, many of the samples fractured similarly to the

sample shown in the right of Figure 3. In this image, the axial load axis is still represented as

a vertical line through the center of the sample, however, the fracture is angled and off

center. Samples which fractured along such diagonal plane likely underwent shear failure

before the tensile stress at the center of the sample reached the true tensile strength of the

sample. Therefore, the tensile strengths inferred form the peak load observed during such

tests are only lower bounds of the actual tensile strength of the rocks.

Page 22: by Peter Vigilante A thesis submitted in partial

10

Figure 3. Post testing images of two separate schist Brazilian disc samples. On the left,

sample 41A2 shows the expected vertical fracture (tensile crack) within the disc. On the

right, sample 65B3 shows unexpected fracture angled away from the axial load (which

is vertical in both photos)

Table 2 shows information from all of the Brazilian Disc samples and tests. Table 2

also shows the normalized peak load (kN/cm) taking into account the thickness of all the

samples, and the tensile strength of the sample (MPa) determined by the maximum peak load

reached during the Brazilian Test. Because a tensile fracture at the center of the disc was not

produced in many samples, the tensile strength is written as a lower limit using the “>”

symbol for those samples that did not produce a load-parallel tensile fracture.

Page 23: by Peter Vigilante A thesis submitted in partial

11

Table 2. Brazilian Disc Samples tested. * indicates Schist samples. “Qtz”

indicates samples are quartz. “Rhy” indicates samples are rhyolite samples. “>”

indicates samples which did not fracture vertically and centered, thus not reaching the

samples true maximum tensile load.

Sample Borehole Depth Description Axial

load to

foliation

angle

(deg)

Normalized

Peak Load

(kN/cm)

Tensile

Strength

(MPa)

56A1* kISMET-

003

278’ No intense

microfolding.

Plenty of

brown/gold

coloration.

Foliation

consistent

throughout.

0 1.9 4.1

56A2* kISMET-

003

278’ 45 >2.3 >4.9

56B1* kISMET-

003

278’ 90 >3.7 >7.9

56B2* kISMET-

003

278’ 90 >5.6 >12.0

41A2* kISMET-

003

202.5’ Lighter

portion of the

core. Whitish

hue.

Intense

folding and

possible

quartz veins

present.

0 1.5 2.7

41C1* kISMET-

003

202.5’ 30 >1.3 >2.5

41C3* kISMET-

003

202.5’ 60 2.2 4.7

41C2* kISMET-

003

202.5’ 90 >2.7 >5.8

63A1* kISMET-

003

312.5’ Not as

apparent

foliation.

Homogenous

looking

portion of the

core.

0 2.4 4.7

63A3* kISMET-

003

312.5’ 45 >2.8 >6.3

63B2* kISMET-

003

312.5’ 90 3.6 7.6

Page 24: by Peter Vigilante A thesis submitted in partial

12

65B1* kISMET-

003

320’ Darker gray

portion of the

core.

Intense

microfolding.

0 1.9 4.8

65C3* kISMET-

003

320’ 0 1.8 3.9

65C2* kISMET-

003

320’ 45 >3.4 >7.9

65B3* kISMET-

003

320’ 90 >3.5 >8.6

Qtz2 DUSEL

J

4850’ Pure quartz

vein.

No foliation

to test

anisotropy.

N/A 2.8 6.0

Qtz4 DUSEL

J

4850’ N/A 5.8 15.8

Qtz5 DUSEL

J

4850’ N/A 4.2 9.1

rhy2 DUSEL

J

4850’ Pure rhyolite

portion of

core.

Whitish-gray

with darker

specks.

No foliation

to test

anisotropy.

N/A 8.6 19.6

rhy3 DUSEL

J

4850’ N/A 10.3 22.9

rhy4 DUSEL

J

4850’ N/A 11.6 24.7

rhy5 DUSEL

J

4850’ N/A 9.7 22.2

In Figure 4, the data in Table 2 above are plotted to represent both the “true” tensile

strength data points, as well as the lower bound tensile strength points. The solid blue circle

data points represent the “true” tensile strength data points, meaning the disc sample broke in

tension along a near vertical fracture. The blue unfilled circle data points represent the lower

bound tensile strength samples. For these samples, they often fractured at an angle away from

Page 25: by Peter Vigilante A thesis submitted in partial

13

vertical, and thus were considered to be a lower limit of the tensile strength of the sample,

and not the true, higher tensile strength of that sample.

Figure 4. Plot representing the angle formed between the vertical axial load of

the Brazilian Disc test and the foliation of the disc sample vs the tensile strength of each

sample clarifies (MPa). Solid blue data points represent true tensile strengths of

samples, while blue unfilled data points represent lower bounds of those samples tensile

strength. All samples are Poorman Schist, and raw data are shown in Table 2.

Table 3 below summarizes the anisotropy of tensile strengths observed from the

results in Table 2 and Figure 4. Table 3 shows the average peak normalized load (kN/cm),

average tensile strength (MPa), and standard deviation of tensile strength data points of the

Brazilian disc samples of each rock type, and how the foliation affected these averages.

However, for these values, the “lower limit” values were not included in the averages. For

0

2

4

6

8

10

12

14

0 20 40 60 80

Ten

sile

Str

engt

h (

MP

a)

Axial Load to Foliation Angle

Axial Load to Foliation Angle of Brazilian Disc Samples vs Tensile Strength (MPa)

Actual Tensile Strengths Lower Bound Tensile Strengths

Page 26: by Peter Vigilante A thesis submitted in partial

14

the Poorman Schist samples, multiple samples were conducted where the foliation was

angled at 0°, 45°, and 90° to the applied load. Although only a limited number of samples

produced actual tensile strength measurements, the average tensile strength of the disc

samples increased as the angle between the axial load and the foliation increased. The

average tensile strength increased from 4.1 MPa to 8.4 MPa as the angle between the

foliation and applied load increased from 0° to 90°. Both the rhyolite and quartz samples did

not show strength anisotropy. The average tensile strength of the rhyolite samples was the

highest of the three rock types at 22.3 MPa, and the quartz vein samples were stronger than

the Poorman Schist but weaker than the rhyolite, with an average tensile strength of 10.3

MPa.

Table 3. Breakdown of average normalized load (kN/cm) for the angle between foliation

and vertical axial load for 0°, 45°, and 90°

Axial Load to

Foliation Angle -

Rock Type

# of

Successful

Samples

Avg. Peak

Normalized

Load (kN/cm)

Avg. Tensile

Strength (MPa)

Std. Dev.

(MPa)

0° - Poorman Schist 5 1.9 4.1 0.9

45° - Poorman Schist 0 N/A N/A N/A

90° - Poorman Schist 1 3.8 7.6 N/A

N/A - Rhyolite 4 10.0 22.3 2.1

N/A -Quartz 3 4.3 10.3 5.0

Page 27: by Peter Vigilante A thesis submitted in partial

15

Discussion - Brazilian Disc Test

The Brazilian disc test and its calculated indirect measurement of tensile strength is

important because the failure point during hydrofracturing tests is directly affected by the

tensile strength of the rock in the borehole. The minimum tangential stress around a vertical

borehole is given by:

𝜎𝜃𝜃𝑚𝑖𝑛 = 3𝑆ℎ𝑚𝑖𝑛 − 𝑆𝐻𝑚𝑎𝑥 − 2𝑃0 − ∆𝑃 (𝟐)

Where Shmin is the minimum horizontal stress, SHmax is the maximum horizontal stress,

P0 is the pore pressure (which is ~zero at the 4850L), ∆P is the difference between the

injected fluid pressure in the borehole and the formation pore pressure (Zoback, 2007). Note

that here we do not take into account the thermal stress. In a laboratory triaxial setting which

we conduct our hydraulic fracturing tests under, the equation can be simplified to the

Equation 3 below where Pb is the minimum pressure needed to create a vertical fracture in the

borehole, PC is confining pressure and T is tensile strength:

𝑃𝑏 = 2𝑃𝐶 + 𝑇 (𝟑)

The Laboratory results showed the apparent Brazilian disc tensile strengths ranged

between 3-8.5 MPa when samples were loaded foliation-parallel, whereas the strengths were

about 8 MPa for the one successful sample loaded foliation-normal. Comparing these results

to Pariseau (1985) tensile strength results, both are in a similar range of tensile strength

values. Pariseau (1985) cited two sets of anisotropic rock property data, an original data set

and a modified rock property data set. The tensile results in this report are closer to

Pariseau’s (1985) original data set, which produced tensile measurements in three orthogonal

directions with values of 11.9 MPa, 6.9 MPa, and 14.0 MPa. Their modified data set

produced tensile measurements in three orthogonal directions of 20.6 MPa, 5.7 MPa, and

Page 28: by Peter Vigilante A thesis submitted in partial

16

13.2 MPa, which are slightly higher than our tensile results on average compared to their

original data set. RESPEC (2010) also conducted tensile strength tests, but these were

conducted on the Yates amphibolite and the rhyolite rock. The mean tensile strength for 13

rhyolite samples RESPEC tested was 10 MPa, while our 4 rhyolite disc samples averaged

22.3 MPa for tensile strength. This shows a clear discrepancy between previous RESPEC

(2010) rhyolite tensile strength results and this reports results.

The data of the tensile strengths associated with their foliation direction compared to

the axial load clearly indicate tensile strength anisotropy in the Poorman schist rocks. The

best supporting argument for this is looking at all of the apparent tensile strength (MPa)

measurements, including the samples which produced tensile fractures as well as shear

failure. For samples with a 0° between the vertical axial load and the foliation plane is 4.1

MPa, for 45° samples there was an average tensile strength of 6.4 MPa, and for 90° samples

there was an average tensile strength of 8.4 MPa. This clear increase in average tensile

strength as the angle between the vertical axial load and the foliation of the sample increases

indicates strong strength anisotropy in the Poorman schist.

Previous tensile strength tests on foliated rocks have also been conducted using

uniaxial direct tensile measurement techniques. These results can show possible similarities

or differences between Brazilian testing and direct testing results. Youash (1966) tested well

bedded and laminated sandstones, shales, and gneiss samples while altering the direction of

loading to the foliation angle in his direct tensile strength tests. Youash found that samples

tested with an angle between 0°-30° from the loading direction and foliation fractured across

the foliation and had the highest tensile strengths, and samples tested with angles between

45°-90° between the loading direction and foliation fractured along planes of weakness, and

Page 29: by Peter Vigilante A thesis submitted in partial

17

had lower tensile strength values (Youash, 1966). These results are shown in Figure 5, and

agree with the general trend of tensile strength anisotropy of our results from the Homestake

Mine samples. However, the one difference which must be taken into account when viewing

the data sets is that our samples were tested using the Brazilian disc method which causes

tensile failure under compressional forces, rather than direct tension. So, the data will look

flipped between the two data sets, but the same tensile strength anisotropy is apparent.

Figure 5. Anisotropy of Ultimate Strength in Uniaxial Direct-Pull Tension of

Lyons Laminated Sandstone. T is the tensile strength (MPa) and is the Angle

Between the Loading Direction and the Samples Foliation from (Kwaśniewski, 2009;

original data from Youash, 1966)

Olsen (1967) also conducted a similar direct tensile strength test on Morrow Point (II)

mica schist. Olsen varied the angle between the loading direction and the foliation of the rock

Page 30: by Peter Vigilante A thesis submitted in partial

18

from 15°-65°, and similar trend in tensile strength anisotropy was consistent with our data

and Youash’s (1966) data. As Olsen increased the angle between the loading direction and

the foliation of the sample, the apparent tensile strength of the samples decreased. This is

shown in Figure 6.

Figure 6. Anisotropy of Ultimate Strength in Uniaxial Direct-Pull Tension of

Morrow Point mica schist. T is the tensile strength (MPa) and is the Angle Between

the Loading Direction and the Samples Foliation from (Kwaśniewski, 2009; original data

from Olsen, 1967)

Page 31: by Peter Vigilante A thesis submitted in partial

19

2.2 Triaxial Dynamic Moduli Test

Methods – Triaxial Dynamic Moduli test

The second laboratory test conducted was the triaxial tests to obtain dynamic moduli.

The cores retrieved from the kISMET project, as well as the DUSEL project, show intense

folding and microfolding within its foliation. Ideally, dynamic elastic properties should be

measured at various loading directions against the foliation plane, but the small core diameter

and intense variability of the rock texture posed difficulty in producing multiple triaxial

samples from similar depths with consistent rock texture. Therefore, the focus in preparation

of this test was to obtain a pair of 2.54 cm diameter cylindrical cores from adjacent portions

of the core with perpendicular foliation to one another.

Similar to the initial step in producing the Brazilian disc samples, coring

perpendicular to the axis of the NQ2 (50.5 mm or 1.99 in. diameter) core was used to obtain

some of the triaxial samples. To obtain sets of samples with perpendicular foliation to one

another, other samples were often cored axially to obtain cores in the opposite orientation.

Four pairs of cylindrical samples were obtained after coring from the original NQ2 sized

cores. For each pair of cylindrical samples from the same depth, the axis of the cylindrical

sample was either parallel or normal to the foliation. We refer to the former as a vertical (V)

foliation samples, and the latter as a horizontal (H) foliation samples in the tables and figures

of this report. Once coring of these samples was completed, ends of the cylindrical cores

were ground using a surface grinder to produce parallel surfaces, necessary for the triaxial

tests and subsequent uniaxial testing.

Page 32: by Peter Vigilante A thesis submitted in partial

20

Same as for the Brazilian disc test, the GCTS RTR-1000 Triaxial Testing System was

used for the triaxial testing. The test began with an initial pressurization of the confining

pressure to 21 MPa, which is representative of the in-situ minimum horizontal stress

measured at the 4850’ level of the Homestake Mine (Oldenburg et al., 2016). Axial

differential stress of 21 MPa was then applied, held for a three-hour creep step while

ultrasonic measurements were taken in order to obtain dynamic elastic moduli, and then the

axial differential stress was finally unloaded. Axial differential stress of 21 MPa was chosen

to match the approximate value estimated from the stress measurements (Oldenburg et al.,

2016).

Results – Triaxial Dynamic Moduli Test

Dynamic and static elastic properties were measured under triaxial stress conditions.

The average density of the Poorman schist samples was 2.75 g/cm3, which was used for

calculating dynamic moduli from the p-wave and s-wave velocity measured during the

triaxial testing. Table 3 shows the p- and s-wave velocities (in m/s). The average p-wave

velocity for the eight samples was 4772 m/s, and the average s-wave velocity for the eight

samples was 2848 m/s. The anisotropy effect was not as clearly shown in the velocity

measurements when comparing the foliation direction of the samples. The p wave velocity

average was 4672 m/s for foliation-normal samples (vertical samples), and 4847 m/s for

foliation-parallel samples (horizontal samples). The s-wave velocity average was 2860 m/s

for foliation-normal samples, and 2838 m/s for foliation-parallel samples. Using the velocity

measurements, Shear (G) and M modulus were also calculated. The average shear (G)

modulus for the eight samples was 22.5 GPa. The raw velocity data and the calculated

Page 33: by Peter Vigilante A thesis submitted in partial

21

moduli are in Table 4 below. Lawrence Berkeley National Laboratory (LBNL) also produced

velocity data on cores retrieved from the kISMET field site. However, they used a different

testing technique which utilized immersion transducers to calculate ultrasonic P wave

velocity measurements in a submerged water tank environment. This data and its significance

are discussed further below in the uniaxial/triaxial discussion section.

Table 4. P-wave and S-wave velocities for triaxial samples tested, calculated

dynamic moduli, and UCS data (discussed in the next section). No velocity data were

recorded for sample 56H prior to breaking the sample. M (P-wave modulus), G (Shear),

were calculated from acquired velocity measurements.

Sampl

e

Borehol

e

Dept

h Description

p wave

velocit

y (m/s)

s wave

velocit

y (m/s)

M

Modulus(GPa

)

Shear

Modulu

s (GPa)

41V kISMET

-003 204’ Lighter

portion of

core. Intense

folding.

4962 2909 69.2 23.8

41H kISMET

-003 204’ 4470 2929 56.4 24.2

49V kISMET

-003

242.5

Foliation is

folding but

consistent.

Dark bands

in foliation.

4709 2747 60.6 20.6

49H kISMET

-003 242.5 4988 2896 68.0 22.9

56V kISMET

-003

279.5

Foliation

relatively

perpendicula

r to axis. Not

much

folding.

4495 2628 54.9 18.8

56H kISMET

-003

279.5

’ N/A N/A N/A N/A

63V kISMET

-003 313’ Foliation

relatively 4561 2603 56.8 18.5

Page 34: by Peter Vigilante A thesis submitted in partial

22

63H kISMET

-003 313’

consistent.

Normal

Poorman

grayish

color.

4559 2757 55.7 20.4

Discussion – Triaxial Dynamic Moduli Test

The triaxial portion of the test was when velocity measurements were taken to

calculate dynamic moduli on the samples. Stress-strain data were acquired when the

additional axial load was applied. These data are stored in the GCTS RTR-1000 machine and

could be interpreted for static modulus. The velocity measurements did not show as clear of

an anisotropic effect as the UCS rock strength data. The average P-wave velocity for

foliation-parallel samples was 4682 m/s, while the average P-wave velocity for foliation-

normal samples was 4672 m/s. The P-wave velocity measurements for foliation-parallel

samples ranged 467.35 m/s, while the foliation-normal samples ranged 517.46 m/s. The S-

wave velocities were also quite similar in magnitude, with foliation-parallel samples

producing an average S-wave velocity of 2722 m/s, and foliation-normal samples producing

an average S-wave velocity of 2860 m/s. The range for S-wave velocities for foliation-

parallel samples was 305.5 m/s, while foliation-perpendicular samples was 172.3 m/s.

Lawrence Berkeley National Laboratory (LBNL) also conducted velocity

measurements on cores retrieved from the kISMET site (Oldenburg et al., 2016). Differing

form our velocity measurements, clear anisotropy was seen in the data produced by LBNL.

Figure 7 produced by LBNL shows the velocity measurements of two core types, long core

and short core, tested for velocity around the circumference of the samples, and along the

length of the samples. Both show clear anisotropy in that the velocity measurements are not

Page 35: by Peter Vigilante A thesis submitted in partial

23

constant around and along the samples. However, this anisotropy which is not apparent in our

data set could be due to the confining pressure applied during our testing. Our samples were

under 21 MPa confining pressure within a triaxial set up while velocity measurements were

recorded, and LBNL’s samples were tested immersed in a water bath with no confining

pressure. With confining pressure, foliation planes may close, and thus diminish the

anisotropic effect that may be prevalent in an unconfined situation.

Figure 7. Seismic P-wave velocity measurements taken (a) around the circumference of

samples and (b) along the length of samples from the kISMET field site. Velocity

anisotropy is apparent for the entire length of the cores from (Oldenburg et al., 2016)

Due to low contrast between the velocity measurements for foliation-parallel and

foliation-perpendicular samples, calculated shear modulus (G) results were similar between

both types of foliated samples. Preliminary full waveform sonic logs with calculated dynamic

elastic constants were produced during the kISMET field work, and presented in the

Page 36: by Peter Vigilante A thesis submitted in partial

24

“kISMET Field Work Results” portion of this report. The field results produced P wave

velocities ranging from around 4500 m/s to 5500 m/s, and S wave velocities ranging from

around 2750 m/s to about 3375 m/s (Oldenburg et al., 2016). Our results for both P-wave and

S-wave velocities fall into a similar range of values, thus showing consistency between the

two velocity data sets for field and lab data.

2.3 Uniaxial Moduli and Unconfined Compressive Strength (UCS) Test

Methods – Uniaxial Moduli and Unconfined Compressive Strength (UCS) Test

The same samples used for the triaxial test to measure the dynamic moduli were also

used for the uniaxial static moduli as well as the UCS test. This was done because in one test

using the triaxial set up, the sample could undergo the triaxial conditions and have the

velocity measurements recorded, and then all the pressures applied to the sample could be

released, and the axial stress could be increased until sample failure. This second “step” in

the test run was the Uniaxial static moduli and UCS test.

Following the triaxial test, the sample had no confining pressure or axial pressure

applied to the sample. With this unconfined stress state, UCS was then measured by applying

axial load with no confining pressure until rock failure. While the axial load was increased,

strain gauges which were applied to all the samples recorded strain data which then produced

stress strain curves below in Figure 8. The final step in the uniaxial test was loading until

sample failure occurred, and this is the data used for our UCS measurements.

Page 37: by Peter Vigilante A thesis submitted in partial

25

Results - Uniaxial Moduli and Unconfined Compressive Strength (UCS) Test

For the strength of the samples, unconfined compressive strength tests were

conducted on all eight triaxial samples after the ultrasonic velocities were measured at

triaxial stress conditions. Table 5 shows the maximum stress each sample endured prior to

failure. These values are thus the UCS of each of the eight samples. The average UCS of the

eight samples of the Poorman schist was 96.4 MPa. The anisotropy effect on the strength of

the samples was also obtained by comparing the strength of the samples with different

foliation directions. The average UCS for foliation-parallel vertical samples was 107 MPa,

and the average UCS for foliation-normal horizontal samples was 88.1 MPa. These data are

shown in Table 6. In general, the samples with foliation parallel to the sample axis had a

higher average UCS than the samples with foliation normal to the sample axis.

Table 5. All uniaxial samples with UCS values along with descriptions

Sample

Borehole

Depth

Description

UCS

(MPa)

41V

kISMET-

003

204’ Lighter portion

of core. Intense

folding.

89.5

41H

kISMET-

003

204’

94.0

49V

kISMET-

003

242.5’ Foliation is

folding but

consistent. Dark

bands in

foliation.

88.5

49H

kISMET-

003

242.5

89.8

Page 38: by Peter Vigilante A thesis submitted in partial

26

56V

kISMET-

003

279.5’ Foliation

relatively

perpendicular to

axis. Not much

folding.

121.9

56H

kISMET-

003

279.5’

69.9

63V

kISMET-

003

313’ Foliation

relatively

consistent.

Normal

Poorman

grayish color.

110.7

63H

kISMET-

003

313’

98.6

Table 6. Average UCS for samples of perpendicular foliation. “V” Samples have

foliation-parallel to cylindrical sample axis. “H” Samples have foliation-normal to the

cylindrical sample axis.

Foliation of Sample Avg. UCS (MPa)

"V" samples 107.4

"H" samples 88.1

Young’s modulus and Poisson’s ratio were also calculated for the samples under

uniaxial conditions. The stress (MPa) vs strain (E) curves for the samples (Figure 8 below)

were used to make these calculations. Results of these calculations are presented in Table 7

below. The average Young’s Modulus for the eight samples was 58.0 MPa, and the average

Poisson’s ratio was 0.22. Anisotropy seemed to have an effect on the Young’s modulus of

the samples, but not as much on the Poisson’s ratio. The average Young’s modulus for

foliation-normal horizontal samples was 52.3 MPa, while the average for foliation-parallel

vertical samples was 63.4 MPa. The average Poisson’s ratio of foliation-normal horizontal

Page 39: by Peter Vigilante A thesis submitted in partial

27

samples was 0.22, and that of foliation-parallel vertical samples was 0.22. The difference in

Young’s Modulus for the different foliation angles was a 18% difference, while the Poisson’s

ratio was only 2.7%, a significantly smaller difference.

Table 7. Calculated Young’s Modulus and Poisson’s Ratio for all eight triaxial samples.

Both calculated using stress vs strain curves provided in Figure X (the stress strain

curves).

Sample Young's Modulus

(GPa)

Poisson's

Ratio

41V 87.2 0.20

41H 45.1 0.28

49V 57.0 0.22

49H 60.1 0.21

56V 53.1 0.20

56H 56.0 N/A

63V 57.3 0.24

63H 47.9 0.18

Page 40: by Peter Vigilante A thesis submitted in partial

28

Page 41: by Peter Vigilante A thesis submitted in partial

29

Figure 8. Stress (MPa) vs Strain (E) curves for all eight Schist triaxial samples

tested. Note: no radial strain data was acquired for sample 56H (bottom left) prior to

bringing sample to failure during testing. Also, “depth” is depth in feet from the 4850’

level of the Homestake Mine.

Figure 8 are images of all eight samples tested under uniaxial/triaxial conditions for

this thesis. All of these pictures show the samples post-testing, so they are post-failure

samples. The top row of the samples is the “H” samples, or samples with foliation normal to

the vertical applied load for UCS tests. The bottom row is the “V” samples, of the samples

with foliation parallel to the vertical applied load for UCS tests. The rock appears shiny or

slightly saturated because they all came into contact with some amount of confining oil post-

testing. Fractures are clearly seen in all eight of the samples, however, their orientations and

lengths vary significantly.

Page 42: by Peter Vigilante A thesis submitted in partial

30

Figure 9. Images of triaxial samples taken after they were brought to failure in

UCS testing. Samples with foliation normal to sample axis are on the top row, and

samples with foliation relatively parallel to sample axis are the bottom row.

Discussion – Uniaxial Moduli and Unconfined Compressive Strength (UCS) Test

The unconfined compressive strength (UCS) destructive test also indicated a certain

degree of anisotropic behavior. Strength anisotropy was shown in the UCS test data, and

static and elastic property anisotropy was shown in the triaxial portion of the test. It is also

important to note that the foliation in these pairs of samples was not perfectly perpendicular

to one another. With limited core to work with, the sample pairs in the uniaxial tests were a

best approximation for samples with foliation perpendicular to one another.

Page 43: by Peter Vigilante A thesis submitted in partial

31

The rock samples tested in the UCS test produced data which were broken down to

compare samples with foliation-parallel and foliation-normal orientation to the vertical axial

load to show strength anisotropy. The average strength for the samples with foliation-parallel

to the vertical axial load was 107.4 MPa, and the average strength for the samples with

foliation-normal to the vertical axial load was 88.1 MPa. This showed that the average

strength of samples with foliation-parallel to the vertical axial load was on average almost 20

MPa higher than the triaxial samples with foliation-normal orientation. This shows clear

strength anisotropy in the eight samples which we tested for this portion of the test. However,

the range for foliation-parallel samples was 33.4 MPa and the range for foliation-normal

samples was 28.7 MPa, both quite significant. If a higher number of Poorman Schist UCS

tests were run focusing on the varying foliation between samples, it would help confirm this

significant strength anisotropy.

Figure 9 shows images of all eight samples post UCS test and fracturing. The top row

shows the foliation-normal samples, and the bottom row shows the foliation-parallel samples.

An observation on the foliation-normal samples is that fractures often were at a higher angle

(sometimes close to 45°) compared to the foliation-parallel samples, and show rough non-

planar fractures which mostly do not extend through the entire sample and to the bottom of

the samples. The foliation- normal samples seemed to chip off corners of the cylindrical

sample upon fracturing. The foliation-parallel samples (which had a higher average UCS)

tended to show failure planes which are more planar and extend through the entire length of

the samples. The foliation clearly affected the morphology of the fractures and thus also the

resulting UCS of the samples.

Page 44: by Peter Vigilante A thesis submitted in partial

32

A possible source of error in our UCS data may come from the 2:1 ratio of

differential stress introduced to the samples during the creep portion of the triaxial tests. This

could have potentially damaged the samples to an extent, which could have biased the UCS

to lower values. However, the maximum horizontal stress applied to the samples during this

creep step was 42 MPa, which is less than half the average UCS of our samples. The

Poorman schist is a relatively hard rock, and believed to have been under similar in-situ

conditions before being recovered from the mine. Thus, hopefully possible weakening of the

samples did not occur before the UCS was measured.

Previous research has also conducted similar UCS testing to determine the effect of

anisotropy. Kwaśniewski (2009) conducted UCS tests on Zloty Stok crystalline mica schist.

Similar to the previously cited papers who focused on direct tension tests, Kwaśniewski also

altered the angle between the direct compressional load of the UCS test and the foliation of

cylindrical triaxial samples. Kwaśniewski altered the angle between the load and foliation of

the sample between 0°-90° by 15° increments. What he found was the highest UCS values

when the foliation of the cylindrical samples were 0° from the compressional load (or the

foliation was parallel to the load direction), then the UCS values decreased to a certain angle

(about 30° for his data) and then increased again to 90°, creating a parabolic curve. This

curve for mean UCS values at differing angles between the load and foliation is shown below

in Figure 10.

Page 45: by Peter Vigilante A thesis submitted in partial

33

Figure 10. Anisotropy of Unconfined Compressive Strength (C of Zloty Stok schist as

, the angle between the axial load and samples foliation, is altered between 0°-90° from

(Kwaśniewski, 2009)

However, the average UCS at 90° was about 57 MPa lower than at 0° for this data set.

This ratio of samples with an angle of 90° between the axial load and sample foliation and

samples with an angle of 0° between the axial load and sample foliation was 0.66 (111.4

MPa/168.6 MPa). Due to constraints on cores used for testing and the intense folding within

those cores, we were only able to test samples on the two extremes of 0° between loading

direction and foliation and 90° between loading direction and foliation. Because of this we

cannot confirm that our samples acted in the same parabolic UCS curve with varying the

angle between axial load and foliation, but we can confirm that samples with 0° between

axial load and foliation had a higher average UCS than samples with 90° between axial load

Page 46: by Peter Vigilante A thesis submitted in partial

34

and foliation. From our UCS data set, samples with 0° between the axial load and sample

foliation averaged 107.4 MPa while samples with 90° between the axial load and sample

foliation averaged 88.1 MPa. This ratio of UCS averages for the two end points is 0.82,

which is higher than Kwaśniewski (2009) results of 0.66.

Although both our data and Kwaśniewski (2009) showed greater UCS values at 0°

between the sample foliation and axial load compared to 90°, past theories predict that the

two end members of 0° and 90° in UCS tests should produce similar UCS values. Jaeger

(1960) assumed that the plane of weakness in a rock sample would have one set of values for

cohesion and internal friction, and that any other plane in the rock would have another set of

values for cohesion and internal friction. Using these assumptions and the Coulomb theory to

“calculate the resistance to failure on the plane of weakness and on the most favored plane

intersecting it, in terms of the principal stress” led to the theory that compressive strength is

dependent on orientation, and produced Figure 11 below (Paterson and Wong, 2005).

Page 47: by Peter Vigilante A thesis submitted in partial

35

Figure 11. Dependence of Differential Stress (y-axis) at Shear Failure on the

Orientation of the Weakness Plane (x-axis) in a Sample of Anisotropic Rock Model from

(Paterson and Wong, 2005; Produced by Jaeger, 1960)

Using the stress-strain data plotted in Figure 8 of the uniaxial test results, Young’s

modulus and Poisson’s ratio were calculated for the triaxial samples. This stress strain data

was recorded during the uniaxial test portion of the uniaxial/triaxial test. Anisotropy was

greater in the Young’s Modulus compared to the Poisson’s ratio. The preliminary full

waveform sonic logs produced from the kISMET field work also produced Young’s Modulus

and Poisson’s ratio data for the kISMET-003 well in which our samples were recovered from

and the field stress measurements were conducted. The preliminary full waveform sonic logs

produced Young’s modulus results which ranged from about 55 GPa to 85 GPa, and

Poisson’s ratio which ranged from about 0.05 to 0.28. Our results showed an average

Young’s Modulus of 58 GPa and a range between 45 GPa and 87 GPa. This average and

range show consistency with the field moduli results. Our laboratory Poisson’s ratio results

produced an average of 0.22 and a range of values between 0.18 and 0.28, which also fall

between the Poisson ratios determined from the kISMET-003 borehole.

A key observation which supports the consistency of our data is apparent in groups 41

and 49 samples, where the sample with the larger P-wave velocity measurement (samples

41V and 49H) also had the higher Young’s Modulus result compared to the groups other

samples (41H and 49V). This same trend could have been true for sample group 56, but lack

of data made it impossible to compare. Sample group 63 produced P-wave velocity

measurements which were nearly identical, but their moduli differed, so we did not see the

Page 48: by Peter Vigilante A thesis submitted in partial

36

agreement for this sample group. This general agreement between the static moduli and

dynamic moduli support the credibility of our data set. Also, the consistency and range of

Poisson’s ratio values throughout the samples is promising as well.

2.4 Laboratory Hydrofracture Test

Methods - Laboratory Hydrofracture Test

The fourth test conducted was the lab hydraulic fracturing test (Avasthi, 1981). The

primary goal of the lab hydrofracture testing was to determine if premature shear failure

along weak foliation planes affect the breakdown pressure of the Poorman formation

samples. Another goal of the testing was to document the initial hydraulic breakdown

pressure for all the samples, as well as the fracture reopening pressure, and see how these

compare with the field results.

The same NQ2 (50.8 mm or 1.99 in. diameter) recovered core was used for the lab

hydrofracture test as the Brazilian disc and triaxial test and uniaxial tests. The NQ2 diameter

core was cut with a rock saw, and although the sample lengths varied, the average

hydrofracture sample was 7.874 cm length, and 5.05 cm diameter. A total of twelve

hydrofracture samples were initially produced. The next step was to remove a smaller

diameter hole in the center of the samples, which would eventually be used to inject the

water for the testing. A 0.3175 cm coring bit was used to core out a 4.1275 cm long hole

centered in the middle of one end of the sample, as shown in Figure 12.

High pressure steel tubing pieces 5.0165 cm long and slightly smaller than 0.3175 cm

diameter were inserted into the small borehole. These tubing pieces were manufactured with

Page 49: by Peter Vigilante A thesis submitted in partial

37

a groove at one end, which would be used to hold an o-ring in place to seal off the injection

zone for the hydrofracture testing. Low viscosity epoxy was also poured between the steel

tubing and the borehole to secure the o-ring and seal off the injection zone. A schematic of

the cross section of the hydrofracture sample and the high-pressure steel tubing pieces are

depicted below in Figure 12.

Figure 12. (above) Cross-section of lab hydrofracture sample and (below) detailed

description of stainless steel high-pressure tubing inserted into lab hydrofracture

samples (also seen in Figure 13 below). O-ring was attached to tubing pieces using the o-

Page 50: by Peter Vigilante A thesis submitted in partial

38

ring groove, and then the tube was epoxied into the 0.3175 cm holes of the samples to

seal the hydrofracture testing zone to the high-pressure tubing in the system.

Following steps to introduce the steel tubing and epoxy into the drilled-out hole,

which included chamfering the opening of the borehole and applying some silicone grease to

allow the tubing to fit, the epoxy was poured and settled, and the tubing was now attached

into the hole and the injection zone was sealed off. The remaining tubing sticking out of the

sample would be inserted into the sample holder platen where it is sealed off with another o-

ring.

Figure 13. Schematic of lab hydrofracture set up. Not depicted to scale, but dimensions

of hydrofracture sample and tubing detailed.

Page 51: by Peter Vigilante A thesis submitted in partial

39

The lab hydrofracture test was again conducted in the GCTS RTR-1000 Triaxial

Testing System. The lab hydrofracture testing set up included a ISCO 100DM syringe pump

to supply the constant water flow for the testing, a vacuum to eliminate any air bubbles in the

high-pressure tubing setup. A schematic of the lab hydrofracture set up is depicted above in

Figure 13.

Specialty platens were made for the lab hydrofracture testing. The platens were 5.08

cm in diameter, slightly larger than the 5.08 cm diameter samples. The bottom platen was

manufactured with an injection hole on the side to carry the high-pressure water to the

sample (seen in Figure 13). The base of the platen, or the portion of the bottom platen where

the sample is placed on, is equipped with a 0.3175 cm hole and another o-ring, used to seal

off the other end of the high-pressure tubing which was epoxied into the samples. With this

set up complete, making sure all connections were sufficiently tightened for high-pressures,

the samples were then ready to be tested.

The tests were run in a triaxial testing set up to be able to replicate the in-situ

conditions of the samples. The first step of the lab hydrofracture testing was to increase the

confining pressure to 22 MPa. This value was chosen because it was representative of the

minimum horizontal stress near the 4850’ level of the Homestake mine (Oldenburg et al.,

2016). The next step was to increase the axial load to a desired force. This desired force was

then converted into additional pressure applied axially to the sample, and was deemed as the

axial stress on the sample. This pressure was varied, to conduct different samples under

different in-situ conditions. The reason for this was to see if there was any correlation in

varying the axial stress and the samples breakdown pressures.

Page 52: by Peter Vigilante A thesis submitted in partial

40

Before all test runs, to eliminate any air bubbles in the high-pressure tubing and to

ensure the testing zone was filled purely with water, the pore pressure lines were vacuumed.

Once the high-pressure tubing system was vacuumed and the desired in-situ conditions for a

given sample were met within the triaxial testing system, the ISCO syringe pump was

operated and monitored for pressure to determine sample breakdown pressures. For all tests,

the pump was run at a constant flow rate of 0.5 mL/min. The LabVIEW software produced a

plot showing the increasing hydraulic pressure in the system with increasing time as the flow

rate stayed constant. Using this real-time plot, it helped decipher when the sample fractured

due to the significant drop in pressure in the system. Upon initial breakdown and obtaining

the peak breakdown pressure for the sample, the fluid pressure was released to complete the

first injection cycle, and then multiple injection cycles were performed using the same flow

rate to observe the fracture re-opening pressure.

Results - Laboratory Hydrofracture Test

Of the 13 prepared hydrofracture samples, 7 of them produced results which were

acceptable. The samples which were not considered acceptable were altered by unstable axial

load controls as well as leakage in the injection fluid. Similar to both the Brazilian and

uniaxial/triaxial tests, the hydrofracture samples were grouped together depending on what

depth in the core they were recovered from. In the title of all of the hydrofracture plots, an

initial number indicates what core run the sample was taken from. The subscript numbers and

letters in the sample names were solely for naming purposes when handling the cores. The

axial stress the samples were tested under are in parentheses for all samples. The pressure

(MPa) vs time (s) plots for all 7 successful test runs are shown below.

Page 53: by Peter Vigilante A thesis submitted in partial

41

0

10

20

30

40

50

60

0 50 100 150 200 250

Pre

ssu

re, M

Pa

Time (s)

64-E (Max. Principal Stress 27 MPa)

Test Zone

0

10

20

30

40

50

60

0 500 1000 1500 2000 2500 3000 3500

Pre

ssu

re,

MP

a

Time (s)

Sample 64-C (Max. Principal Stress 44 MPa)

TestZone

Page 54: by Peter Vigilante A thesis submitted in partial

42

0

10

20

30

40

50

60

0 200 400 600 800 1000

Pre

ssu

re,

MP

a

Time (s)

53-C (Max. Principal Stress 23 MPa)

Test Zone

0

10

20

30

40

50

60

0 200 400 600 800 1000

Pre

ssu

re,

MP

a

Time (s)

53-D (Max. Principal Stress 33 MPa)

TestZone

Page 55: by Peter Vigilante A thesis submitted in partial

43

0

10

20

30

40

50

60

0 200 400 600 800

Pre

ssu

re,

MP

a

Time (s)

53-B (Max. Principal Stress 44 MPa)

TestZone

0

10

20

30

40

50

60

0 200 400 600 800 1000

Pre

ssu

re,

MP

a

Time (s)

Sample 531111-A (Max. Principal Stress 55 MPa)

TestZone

Page 56: by Peter Vigilante A thesis submitted in partial

44

Figures 14-20. Pressure (MPa) vs Time (s) plots for all 7-successful lab

hydrofracture experiments. Title describes what core run the sample came from and

the axial stress on the sample during hydrofracture testing.

Although we only have 7 samples with data deemed acceptable, 5 of those samples

came from one testing group and thus all were tested with different axial stress magnitudes.

The other 2 samples were from a different core depth, and the axial stress was altered

between those two samples as well. The breakdown pressure vs axial stress plots are below in

Figures 21 and 22. In these figures, the fracture reopening pressure for these samples was

also plotted. The hydraulic breakdown pressure is plotted in blue, and the fracture reopening

pressure is in orange for both figures. The raw data for breakdown pressures and fracture

reopening pressures are below in Table 8.

0

10

20

30

40

50

60

0 200 400 600 800 1000

Pre

ssu

re,

MP

a

Time (s)

Sample 531-A (Max. Principal Stress 66 MPa)

Test Zone

Page 57: by Peter Vigilante A thesis submitted in partial

45

Figure 21. Axial Stress (Max. Principal Stress) Applied to Hydraulic Fracturing

Samples vs The Breakdown Pressure of that Sample and Fracture Reopening Pressure

for Samples Recovered from Core Run 64

0

10

20

30

40

50

60

0 10 20 30 40 50

Pre

ssu

re,

MP

a

Max. Axial Stress (MPa)

Run 64 Breakdown Pressure and Fracture Reopening Pressure vs Max. Axial Stress

Max. HydraulicPressure (MPa)

FractureReopeningPressure (MPa)

Linear (Max.Hydraulic Pressure(MPa))

Linear (FractureReopeningPressure (MPa))

0.00

10.00

20.00

30.00

40.00

50.00

60.00

70.00

0 10 20 30 40 50 60 70

Pre

ssu

re,

MP

a

Max. Axial Stress (MPa)

Run 53 Breakdown Pressure and Fracture Reopening Pressure vs Max. Axial Stress

Max. HydraulicPressure (MPa)

Fracture ReopeningPressure (MPa)

Linear (Max.Hydraulic Pressure(MPa))

Linear (FractureReopeningPressure (MPa))

Page 58: by Peter Vigilante A thesis submitted in partial

46

Figure 22. Axial Stress (Max. Principal Stress) Applied to Hydraulic Fracturing

Samples vs The Breakdown Pressure and Fracture Reopening Pressure of that Sample

for Samples Recovered from Core Run 53

Table 8. Raw data for the successful lab hydrofracture tests. All tests had a minimum

principal stress of 22 MPa, and the fracture reopening pressures are an estimate due to

multiple fracture re-openings for the tests.

Sample

Borehole

Depth

Description

Max.

Principal

Stress

(MPa)

Max.

Hydraulic

Pressure

(MPa)

Fracture

Reopening

Pressure

(MPa)

64-E

kISMET-

003

318.5’

Foliation

~20° off

vertical. No

intense

folding. 27 43.3 ~21

64-C

kISMET-

003

319’

Some

folding, but

foliation

about 10°-

20° off

vertical 44 48.5 ~27.6

Run 53-C

kISMET-

003

262.3’

Intense

microfolding.

Foliation

relatively

normal to

vertical axis. 23 47.4 ~21.5

Run 53-D

kISMET-

003

262’

Intense

microfolding.

Shiny

white/silver

areas. 33 52.9 ~23.5

Page 59: by Peter Vigilante A thesis submitted in partial

47

Run 53-B

kISMET-

003

262.6’

Intense

microfolding.

Foliations

goes from

horizontal to

vertical

44 34.8 ~22.7

Run 531111-A

kISMET-

003

263.3’

Not as clear

foliation.

Foliation

dips ~10°-

40° 55 58.4 ~22.2

Run 531-A

kISMET-

003

264’

Not as clear

foliation.

Foliation

dips ~10°-

40° 66 50.0 ~21.9

Discussion - Laboratory Hydrofracture Test

The laboratory hydrofracture experiment was on a different scale than the field

hydrofracturing test, so comparing results of the two tests is not entirely possible. With the

relatively small hydrofracture samples which were produced due to limited core size, the goal

was to maximize the difference between the borehole diameter and the sample diameter to

have the confining pressure act as far field stress. We were still able to create a test where an

interval of the rock was isolated and allowed for the buildup of hydraulic pressure to induce

fracturing. In this environment, it provided the capability to differ the stresses on the samples

to see if this had any effect on the samples.

The average hydraulic breakdown pressure for the 2 core-run 64 samples was 45.9

MPa, with the sample tested under 27 MPa axial stress having a breakdown pressure of 43.3

MPa, and the other sample tested under 44 MPa axial stress having a breakdown pressure of

48.5 MPa. There is a trend of increasing hydraulic breakdown pressure with increasing axial

Page 60: by Peter Vigilante A thesis submitted in partial

48

stress here, but with a small sample size of only two samples it is difficult to conclude that

this is a general trend.

The average hydraulic breakdown pressure for the five core run 53 samples was 48.7.

These five samples with varying axial stresses resulted in hydraulic breakdown pressures of

47.4 MPa, 52.9 MPa, 34.8 MPa, 58.4 MPa, and 50.0 MPa in the increasing order of axial

stress magnitudes. There does not seem to be a clear trend in the breakdown pressure against

axial stress, however, a linear fit to the data does show a slightly positive slope, meaning the

average breakdown pressure may possibly be increasing with increasing axial load.

As for the fracture reopening pressures, for both group of samples from core run 64

and 53, no clear trend with increasing axial stress is observed and all values are close to 22

MPa, which is the confining pressure used for all samples tested. In general, from these

observations, we see that there is no strong dependence of the breakdown and reopening

pressures with the axial stress applied on these samples.

As stated previously in equation 3, the expected breakdown pressure at the borehole

wall in a laboratory triaxial is Pb=2PC+T. This is a simplification of the equation for

breakdown pressure in the field (Pb=3Shmin-SHMax+T) because Shmin and SHMax are equal to

one another in a triaxial set up, due to the confining pressure being equal in all directions

besides the vertical axial load. Using the equation for breakdown pressure in the borehole for

a laboratory setting, we expected to see a range of breakdown pressures between 47 MPa and

56 MPa. The confining pressure used for all the lab hydraulic fracturing tests was 22 MPa, so

that is PC, and we know tensile strength (T) ranged between 3 MPa to 12 MPa from our

Brazilian disc results. Plugging these values in is how the expected range of breakdown

Page 61: by Peter Vigilante A thesis submitted in partial

49

pressures is calculated. Below in Figure 23 it shows the expected range of breakdown

pressures (47-56 MPa) as a red box for the core run 53 sample group. This shows that four of

the five samples fall into the expected range or are very close, while one sample is a clear

outlier and shows a much lower breakdown pressure.

Figure 23. Reproduced with Expected Breakdown Pressures (Pb) shown as Overlaid

Red Box.

The clear outlier of the core run 53 samples was sample 53-B with a breakdown

pressure of about 35 MPa, and was conducted under 44 MPa of axial stress. This figure is

presented again below in Figure 24. Looking at the sample 53-B, and comparing it to the other

successful hydrofracture lab samples, it is apparent that the increase of pressure in the sample’s

test zone was non-linear. This is shown in Figure 24 by showing how the increase in pressure

in the test zone up until the breakdown pressure deviates from linearity (shown as a red line

Page 62: by Peter Vigilante A thesis submitted in partial

50

here). This shows that there was a clear dissipation of pressure in the test zone prior to the

breakdown pressure. This could have been due to fluid leakage into a preexisting or shear-

induced fracture at the test zone. This shows evidence that if there is a source for premature

leak off of the test zone, then the breakdown pressure could be lowered significantly compared

to a sample that overcomes the true breakdown pressure anticipated for a completely intact

rock.

Figure 24. Sample 53-B Pressure vs Time Plot Showing Pressure Increase in Test Zone

Deviates from Linearity (Red Line) to Help Explain Outlier Breakdown Pressure Value

Back to Figure 23 (axial load vs pressure) above it is also important to note that the

fracture reopening pressures average around 22 MPa. This is shown by the orange data points

and the near horizontal trend line between all five of the fracture reopening pressure values.

0

10

20

30

40

50

60

0 200 400 600 800

Pre

ssu

re,

MP

a

Time (s)

53-B (Max. Principal Stress 44 MPa)

TestZone

Page 63: by Peter Vigilante A thesis submitted in partial

51

Although the breakdown pressures mostly fell into the predicted range of values (besides the

one outlier explained above), these fracture reopening pressure are not consistent with what

we had expected from our laboratory measurements. The equation for fracture reopening

pressure should be the same as the breakdown pressure in the lab (Pb=2PC+T) but without the

tensile strength. So, our expected fracture reopening pressure would thus be 44 MPa (Pb=

2*22 MPa = 44 MPa). However, the average fracture reopening pressure is about 22 MPa,

which is equal to the confining pressure being applied to all the samples. We interpret that

the fluid pressure is reopening the fracture that has propagated beyond the stress

concentration produced around the borehole, thus equation (3) is not valid in this case to

predict the fracture reopening pressure.

Studying the hydrofracture orientations, and how the fractures form and propagate

from the borehole is also imperative in interpreting the laboratory and field hydrofracture

experiments. Although we have not successfully image the fractures from the laboratory

samples, we did see other aspects of our results which may indicate how the Poorman schist

altered hydrofracture results. The tensile strength anisotropy which we clearly saw from the

Brazilian disc test is the main indicator from our results that the hydrofracture orientations

may have been altered due to the rock properties and local foliation. Similar to Figure 3 (two

samples with actual and lower bound tensile strength), which saw evidence that certain

samples seemed break at pressures lower than the pressure necessary to cause a true tensile

fracture vertically down the sample, the laboratory and field hydrofractures also may have

formed before overcoming the pressure needed to form a vertical hydraulic fracture which

had to overcome the true tensile strength contributing toward the breakdown pressure. This

Page 64: by Peter Vigilante A thesis submitted in partial

52

would suggest that hydrofractures may have formed along a weak plane, thus showing a

lower apparent breakdown pressure.

Contributing factors which may indicate that hydraulic fractures formed prematurely

is that the foliation is not consistent throughout our core of Poorman schist. With the foliation

changing drastically throughout the core, it is difficult to determine how the fracture formed

relative to the foliation. The hydrofracture is supposed to form in the direction of SHMax and

perpendicular to the Shmin, but if the local foliation alters this and the hydrofracture forms

along a weak plane, then the fracture may not form perfectly in this orientation.

3. Conclusions and Future Work

Past laboratory research had been conducted on cores recovered from multiple

locations of the Homestake mine. However, the cores tested for this thesis were recovered

directly from the kISMET field site, and thus support rationalizing the field results. All three

of the laboratory tests presented in this report were chosen to acquire specific rock property

data to compare with past laboratory data and the field data to see similarities or

discrepancies. The updated and spatially relevant rock property data in this thesis may be

used for future research that may take place at the kISMET site location.

One of the main conclusions from our laboratory work was clear strength anisotropy

from both tensile and UCS strength. Brazilian disc data showed that as the samples were

rotated from foliation parallel to the axial load to foliation normal to the axial load, the

apparent tensile strength of the samples increased. For UCS, samples with foliation parallel

to the axial load had a significantly higher average strength than foliation normal to the axial

Page 65: by Peter Vigilante A thesis submitted in partial

53

load. More samples for both tests would certainly increase our understanding of how strong

these anisotropic effects actually are. Future work for both the Brazilian and UCS test would

include samples tested at more foliation angles than in our testing. For example, for the UCS

test, it would be beneficial to test samples with foliation between foliation parallel and

foliation normal.

The Brazilian disc and UCS test also showed evidence that the strength anisotropy

affected the orientation of the failure planes. As pointed out for the Brazilian disc test, many

of the samples seemed to fracture at weak planes which were not vertical, tensile fractures.

These premature fractures were assumed to break at a lower load than they would have if it

fractured vertically in the middle of the disc sample. These weak planes in the sample thus

affected the apparent tensile strength of samples. Similar to the tensile strength, the UCS

samples of different orientation also fractured differently. Foliation parallel samples

produced planar fractures which extended through the ends of the samples, while most of the

foliation perpendicular samples seemed to chip or fracture at a weak plane. The results from

the Brazilian disc test and the uniaxial and triaxial tests which did not show a strong

anisotropic effect was the elastic and dynamic moduli. Other studies conducted on the

Poorman schist by Lawrence Berkeley National Laboratory showed stronger anisotropic

effect

Our hydrofracture results show the local stress state may not affect the breakdown

pressure all that much. As we saw in our results, as we altered the axial stress on the samples

drastically, the breakdown pressure remained quite constant. This may mean that a factor as

to why the samples breakdown when they do may be controlled more by local foliation or

Page 66: by Peter Vigilante A thesis submitted in partial

54

other intrinsic qualities. Again, future research may be able to delve into this and hone in on

the mechanisms that truly control the hydraulic fracturing breakdown during these tests.

More work needs to be done on the hydrofracture samples in the future to determine a

greater understanding of how these samples fractured, and why certain samples may have a

higher breakdown pressure than others. Determining whether the breakdown was occurring

due to pure tensile fracturing, or a possible mix of shearing as well may indicate how the

fractures formed in the stress measurement field tests as well. To do this, future work would

include imaging the fractures. CT scans were attempted on the hydrofracture samples, but did

not result in successful imaging of the fractures. Other methods that could be used to image

the fractures could be scanning electron microscopy or injecting fluorescent epoxy into the

sample.

Another possibility for future work could be testing the size effect of the samples.

With larger Poorman schist samples, and a different hydrofracture test set up to

accommodate the larger sample. The purpose of testing larger samples would be to see if the

results are altered, and determine if there is any size effect on the testing. An example of how

size effect may affect the results is a larger sample is more likely to have natural fractures,

and thus could decrease the average breakdown pressures of the larger samples. This could

help determine how to interpret our results from this thesis.

Page 67: by Peter Vigilante A thesis submitted in partial

55

4. References

ARUP USA “FA/E Service for Site Investigation in Support of the LBNF Far Site

Conventional Facilities Project - Geotechnical Interpretive Report.” Draft 2,

prepared for South Dakota Science and Technology Authority, Feb. 2015.

Avasthi, J. “Hydrofracturing in Homogeneous, Anisotropic and Fractured Rocks.” University

of Wisconsin-Madison, Ph.D. Thesis, 1981.

Bachman, R.L., and S. W. Caddey “The Homestake Mine, Lead, South Dakota: An

Overview” in Metallogeny of Gold in the Black Hills, South Dakota: A Guidebook

Prepared for the Society of Economic Geologists Field Conference, September 5-9,

1990. Ed. C.J. Paterson, and A.L., Lisenbee, The Society of Economic Geologists

Guidebook Series, 7 1990, pp. 89-94.

Bond, P.H. “The directions and magnitudes of the principal stresses at the 6200 foot level of

the Homestake Mine, Lead, South Dakota: Rapid City, S. Dak.” South Dakota

School of Mines and Technology, M.S. Thesis, 1970, pp. 33.

Caddey, S.W., R.L. Bachman, T.J. Campbell, R.R. Reid, and R.P. Otto “The Homestake gold

mine, an early Proterozoic iron-formation-hosted gold deposit, Lawrence County,

South Dakota.” Geology and Resources of Gold in the United States, 1857-J, 1991,

pp. 1–67.

Campbell, T. J. “Characteristics of the Yates Unit Amphibolite.” South Dakota School of

Mines and Technology, 2004,

homestake.sdsmt.edu/Geology/Characteristics%20of%20the%20Yates%20Unit%20

Amphibolite.pdf. Accessed 15 Dec. 2015.

Page 68: by Peter Vigilante A thesis submitted in partial

56

Duncan-Fama, M.E., and M.J. Pender. “Analysis of the hollow-Inclusion technique for

measuring in-Situ rock stress.” Journal of Rock Mechanics and Mining Sciences, vol.

17)3, 1980, pp. 137–146.

Golder Associates “In Situ Stress Measurement Deep Underground Science and Engineering

Laboratory December 2009.” prepared by Golder Associates Inc. for DUSEL, Dec.

2009

Golder Associates “Geotechnical Engineering Services In-Situ Stress Measurement Deep

Underground Science and Engineering Laboratory.” prepared by Golder Associates

Inc. for South Dakota School of Mines and Technology, Jan. 2010a

Golder Associates “Golder Associates Preliminary Design Final Report.” prepared by Golder

Associates for Deep Underground Science and Engineering Laboratory, Oct. 2010b

Haimson, B. C. “The Hydraulic Fracturing Stress Measurement Method and Recent Results.”

Int. Jour. Rock Mech, vol. 15, 1978, pp. 167–176.

Hooker, V. E., D.L. Bickel, and J.R. Aggson “In situ determination of stresses in

mountainous topography.” Report of Investigation, RI 7654, U.S. Department of the

Interior, Bureau of Mines, 1972.

Jaeger, J. C. “Shear Failure of Anisotropic Rocks.” Geological Magazine, vol. 97(1),1960.

Johnson, J.C., W.G. Pariseau, D.F. Scott, and F.M. Jenkins “In Situ Stress Measurements

Near the Ross Shaft Pillar, Homestake Mine, South Dakota.” Report of

Investigations, RI 9446, U.S. Department of the Interior, Bureau of Mines, 1993.

Kwaśniewski, M. “Testing and Modeling of The Anisotropy of Tensile Strength of Rocks.”

Proceedings of the International Conference on Rock Joints and Jointed Rock

Page 69: by Peter Vigilante A thesis submitted in partial

57

Masses, Tucson, Arizona, USA, Jan. 7-8, 2009.

Lachel Felice and Associates “Geotechnical Engineering Services Final Report for 4850

Level Mapping.” prepared for South Dakota School of Mines and Technology, Sept.

2009.

Lisenbee, A., and M. Terry. “Development of a 3-D Structural Geology Model of

Homestake's 4100 to 5000 Levels at the Proposed Location of the Large Cavities.”

South Dakota School of Mines and Technology Contract #09-05, Consulting Report,

May 2009.

Oldenburg, C. M., et al. “Intermediate-Scale Hydraulic Fracturing in a Deep Mine kISMET

Project Summary 2016.” Oct. 2016,

drive.google.com/file/d/0B3KTfL4IfhRFWDRGcnVuRkxRWjA/view.

Olsen, O. J. “Beobachtungen über die Anisotropie von Glimmerschiefern bei Triaxial- und

anderen Untersuchungen an Gesteinskernen.” in Bericht uber das 8. Landertreffen

des IBG, Eds. G. Bilkenroth, and K.H. Hofer, Berlin: Akademie-Verlag,1967, pp.

98-108.

Pariseau, W.G. “Research Study on Pillar Design for Vertical Crater Retreat (VCR) Mining.”

Final Report, U.S. Department of the Interior, Bureau of Mines, Contract JO215043,

1985.

Paterson, M.S., and T Wong “Experimental Rock Deformation - The Brittle Field.”, second

edition, Springer-Verlag Berlin Heidelberg, 2005.

Ulusay, R., and J. A. Hudson. “Suggested Methods for Determining Tensile Strength of Rock

Materials.” in The Complete ISRM Suggested Methods for Rock Characterization,

Page 70: by Peter Vigilante A thesis submitted in partial

58

Testing and Monitoring: 1974-2006. ISRM Turkish National Group, 2007.

Vigilante, P.J., H. Sone, H.F. Wang, B. Haimson, and T.W. Doe “Anisotropic Strength of

Poorman Formation Rocks, kISMET Project” 51st US Rock Mechanics /

Geomechanics Symposium, San Francisco, California, USA, 25-28 Jun. 2017, paper

766.

Wang, H., M. Lee, T. Doe, B. Haimson, C. Oldenburg, and P. Dobson. “In-Situ Stress

Measurement at 1550-meters depth at the kISMET TEST SITE in Lead, S.D.” 51st

US Rock Mechanics / Geomechanics Symposium, San Francisco, California, USA,

25-28 Jun. 2017, paper 651.

Wang, Q., and L. Xing. “Determination of fracture toughness KIC by using the flattened

Brazilian disk specimen for rocks.” Engineering Fracture Mechanics, 64, 1999, pp.

193-201.

Worotnicki, G. “CSIRO Triaxial Stress Measurement Cell.” in Comprehensive Rock

Engineering, Volume 3, Rock Testing and Site Characterization, Eds. J.A. Hudson,

1993, pp. 329–394.

Youash, Y.Y. “Experimental Deformation of Layered Rocks.” Houston Geological Soc.

Bull., vol. 8, no. 10, 1966, p. 24.

Zoback, M.D. “Reservoir Geomechanics” Cambridge University Press, 2007.

Appendix

Homestake Mine (Lead, SD) Geology and Previous Lab Data

General Geology

Page 71: by Peter Vigilante A thesis submitted in partial

59

The Homestake Mine, located in the northern portion of the Black Hills, South

Dakota, is a part of a Precambrian core complex associated with an elongate domal Laramide

orogenic uplift structure that is 100 km long and 60 km wide. The region has undergone

multiple deformation events as well as continued uplift and erosion of bedrock materials

(Lachel Felice, 2009). The core complex is primarily metamorphosed greenschist or upper

amphibolite facies (Caddey et al., 1991). The metamorphosed portions of the region are

interpreted to be greater than 1.84 billion years old (ARUP, 2015). The Homestake Mine is

composed primarily of three different rock types. Oldest to youngest, these are the Poorman,

Homestake, and Ellison Formations. The gold mineralization, which was what the

Homestake Mine produced when active, occurred solely in the iron-rich Homestake

Formation. The rock formations in the Homestake are intruded by rhyolite and phonolite

dikes, which are dated to be around 53 million years old.

Figure 25 below shows a bird’s eye geologic map of the Homestake Mine and

surrounding regions. Detailed in red on the figure is a bird’s eye perspective of a cross-

sectional line which runs from arbitrary point A to point A’. The cross section of this red

cross-sectional line is depicted in Figure 26 below. The red line is depicted along the 4850’

level of the Homestake Mine in Figure 26, which is where the field work took place. Figure

27 is a view of the 4850’ level, with labeled names of locations on level, as well as overlaid

blue dashes showing the general foliation in areas, as well as red dashes which show rhyolite

dike locations. Figure 27 also shows where the DUSEL holes were drilled along the level, as

well as where the kISMET field work took place.

Page 72: by Peter Vigilante A thesis submitted in partial

60

Figure 25. Map View of Cross Section A to A’ (red line) which is shown in cross

sectional view in Figure 2 From (ARUP, 2015)

Page 73: by Peter Vigilante A thesis submitted in partial

61

Figure 26. Cross sectional view and formations of the Homestake mine. 4850L

represented by solid red line. From (Denver Region Exploration Geologists’ Society)

Page 74: by Peter Vigilante A thesis submitted in partial

62

Figure 27. General Geology Overview of 4850L. Blue Dashes are General Foliation

Orientation and Red Dashes are Rhyolite Dikes. From (Golder Associates Final, 2010a)

Poorman Formation (Schist)

The upper member of the Poorman is made up of metasediments. The planned

locations for both the physics laboratories at the Homestake Mine and the kISMET location

are within these metasediments, referred to as the Poorman Formation Schist (ARUP, 2015).

Page 75: by Peter Vigilante A thesis submitted in partial

63

The Poorman Schist’s protoliths include thinly bedded, carbonate-rich siltstones and

claystones, marl, iron formations, and dolomite. It is often described physically as dark grey

banded and laminated, micacious phyllite to schist, and mineralogically as carbonate-rich

with muscovite, biotite, pyrrhotite, graphite and garnet. The four, distinct rock lithologies, in

decreasing abundance, found in or intruding into the Poorman Formation Schist are Sericite

carbonate quartz schist, Graphitic Schist, Rhyolite, and Biotite quartz carbonate schist. Due

to their sedimentary source, the contacts among these lithologies are often gradational,

although the contacts between the rhyolite dikes and the Poorman formation are very abrupt

with little contact metamorphism and alteration (ARUP, 2015). The Poorman Formation is

described in geologic maps along the drifts of the 4850 level as both phyllite and schist. The

Poorman is said to have distinctive banding along the 4850 level, with band widths varying

from 1 to 5 cm thick. Beds of the Poorman Formation can contain “coarse-grained biotite,

with or without garnet, and varying portions of pyrrhotite” and “massive quartz bands with

pyrrhotite are common”. The Poorman is often folded on all scales due to the preservation of

multiple periods of deformation in this region (Lisenbee and Terry, 2009). The kISMET field

hydraulic fracturing stress measurement testing took place within the Poorman (schist)

Formation, and a majority of the laboratory experiments were conducted on the Poorman

cores recovered from the well where the hydrofracturing experiment took place. These

results are summarized in the field measurement and laboratory experiment sections of this

report.

Yates Unit (Amphibolite)

Both the lower (Yates unit) and upper members of the Poorman Formation are

present at the 4850’ level. The lower Yates unit consists of pillow basalts, which have been

Page 76: by Peter Vigilante A thesis submitted in partial

64

metamorphosed into amphibolites with interbedded metasediments, or sedimentary rock that

has been altered by metamorphism. The Yates Unit is described as dark green, fine- to

medium-grained hornblende plagioclase schist and varies in thickness between 600 to 1,200

meters (Bachman and Caddey, 1990). The amphibolite Yates has strong compositional

layering and is fine grained (Lisenbee and Terry, 2009) and is commonly populated with

white calcite veins typically between 2 millimeters and 2 centimeters wide, and is noted to

have conformable contacts and interbeds (Bachman and Caddey, 1990). In summary, the

Yates Unit is described as “metamorphosed tholeiitic basalt with possible back-arc basin

affinities based in part of the structures that have been interpreted as relic pillows and

conformable overlying metasediments with an overall fine-grained character” (Campbell,

2004). Although the kISMET field work did not take place in the Yates Unit amphibolite,

this information may be pertinent to future research in the Homestake Mine and the 4850’

level.

Rhyolite Dikes

Rhyolite dikes have been mapped in both the Yates Unit and Poorman Formations at

the 4850L (Lachel Felice, 2009]. Their apparent widths vary from 2 to 115 feet. Along the

4850L drifts the dikes tended to strike north-northeast dipping to the east or to strike

northwest dipping steeply to the north. With certain measurement limitations, the rhyolite

dike swarm is believed to be 750 feet wide. Contacts between the rhyolite dikes and the

Yates Unit and Poorman Formation are sharp and clear, however the curvature of the

contacts is often undulating. Due to the undulating contacts, projecting the rhyolite pathways

through the formations is difficult.

Page 77: by Peter Vigilante A thesis submitted in partial

65

Structural Geology

Contact Relationships

The Yates unit (amphibolite) contact with the Poorman Formation is exposed in four

separate drifts on the 4100 and 4850 level of the Homestake Mine. The contact zone of the

Yates with the Poorman is a “zone of broken rock” (Lisenbee and Terry, 2009). The contact

between the Yates unit and the rhyolite dikes are annealed, seen on the 4100 level of the

mine (expecting the contact to be similar on the 4850 level). A few inches of recrystallization

are often seen extending into the Yates amphibolite and the outer margins of the rhyolite

dikes are extremely fine-grained (porcelain-like). The contact between the Yates unit and the

rhyolite dike is undulatory, ranging from centimeters to the strike length of the dike. The

contacts seen between the Poorman (schist) formation and the rhyolite dikes are similar to the

contacts seen between the Yates unit and the rhyolite dikes (Lisenbee and Terry, 2009).

Folding

Geologic mapping has produced evidence of a series of synclines and anticlines in the

region (Lachel Felice, 2009). Folding and deformation can be seen both at small scales down

to inches, and also at larger regional scales. Due to multiple events of deformation in this

region, the structural folding is quite complex, potentially contributing to the complication of

stress measurement testing in this region.

The lead anticline plunges to the south-southwest and its axial trace is located 100-

200 feet west of the Yates shaft and bisects the 4850L. The Yates Shaft and Davis Campus of

the mine lay on the east limb of the fold (Lisenbee and Terry, 2009). According to Lachel

Felice’s 2009 mapping of the 4850 level, the fold axis is believed to be in the Exhaust Drift.

Page 78: by Peter Vigilante A thesis submitted in partial

66

The Yates amphibolite is folded on several smaller scales in addition to the major anticline

fold (Golder Associates Final, 2010a). These smaller scale folds ranged from inches to feet

with plunges of 15°-30° to the East-southeast. The Yates amphibolite has much larger

wavelengths within its folding compared to the Poorman Formation (Golder Associates,

2010a).

Faulting

Faults along the 4850L were characterized by offsets across foliations and quartz veins

(Lachel Felice, 2009). Vertical offset of the faults found is often small (order of inches to

feet), and sometimes found to have graphitic inclusions. These graphitic inclusions could

promote weak shear zones, which could amplify the faults effect on local structural stability.

Rock in the region, which had been altered due to shearing displayed lower rock strength

than the surrounding host rock (ARUP, 2015). For the most part, the faults have evidence of

healed contacts with little or no shear zone along contact (Lachel Felice, 2009).

Fractures and Jointing

Discontinuities are an important aspect of the geology of the region around 4850L of

the Homestake Mine due to their effect on rock mass geomechanical properties such as

strength, deformability, and anisotropy, all of which are important for the stability of the

cavern (Lachel Felice, 2009). Lachel Felice (2009) report 1,495 data points collected during

the mapping of the 4850 Level (See Appendix A in Lachel Felice (2009) for full data set) and

three dominant fracture sets plus one random fracture set were identified. Different studies

have assessed the jointing, fracture networks, and faulting in the region for their effect on

different rock properties that are detailed in this paper. The range found for spacing between

discontinuities along the 4850L was less than 1 inch to roughly 3 feet.

Page 79: by Peter Vigilante A thesis submitted in partial

67

Mine Water Level

The water level of the mine has fluctuated over time. Flooding of the Homestake

Mine first began in 2003 after gold production had previously ceased at the mine. Pumping to

maintain the water level was initiated when it reached 5,000 ft. depth and a maximum water

level of roughly 4,500 ft. was reached before levels started to decrease. The pumping then

maintained the water level at 5,700 ft. This water fluctuation is significant because water

levels did exceed the 4850L of the kISMET project. However, the existence of abundant

space and drifts along the 4850L of the mine, modern underground campuses along the

4850L, and sufficient distance from the water level and the physics experiments (over 1,000

ft. away from physics experiments), made the 4850L an ideal location for hydraulic

fracturing experiments in a mine (Oldenburg et al., 2016). The history and current condition

of the water table in the Homestake Mine is pertinent to this research project since it shows

that the pore pressure at the 4850L is zero.

Previous Unconfined Compressive Strength (UCS) and Tensile Strength

Data

Lachel Felice, 2009

UCS values for intact rock were presented in Lachel Felice (2009). Lachel Felice

(2009) notes that these UCS values are used in the absence of laboratory UCS testing of the

different rock units, and thus must not be considered direct rock measurements. The origins

of these UCS values are unclear through the sources used in this report. Appropriate values

Page 80: by Peter Vigilante A thesis submitted in partial

68

for the Yates amphibolite’s rock strength ranged from 152 to 214 MPa with an average 183

MPa, the rhyolite ranged from 96.5 to 234 MPa with an average of 165.5 MPa, and in

domains that are controlled by both, the average rock strength is 172 MPa. The rock strength

values for the Poorman formation averaged 71 MPa. This is significantly less than the

strength of the Yates unit and rhyolite dikes. In transition zones of contact between the Yates

and Poorman Formation, the average rock strength used in analysis is 81.4 MPa, although the

definition of the transition zone of the Yates and Poorman formations is not clearly defined.

Table 9 below presents the UCS for rock types at the 4850L from the Lachel Felice (2009)

report, and Figure 28 shows the range of UCS values for different rock types along the

4850L.

Table 9. Average unconfined compressive strength (UCS) values for rock types at 4850

level From (Lachel Felice, 2009)

Rock Type AVG UCS (MPa) AVG UCS (psi)

Yates 182.7 26,500

Yates Contact with

Poorman

81.7 11,800

Poorman 71.0 10,300

Rhyolite 165.5 24,000

Page 81: by Peter Vigilante A thesis submitted in partial

69

Figure 28. Range of UCS values for rock types at 4850L From (Lachel Felice, 2009)

Pariseau (1985)

Pariseau (1985) performed laboratory testing on drill core taken in three directions

from the Poorman, Homestake, and Ellison formation. Along with other elastic properties,

Pariseau tested for uniaxial and tensile strength in three mutually orthogonal directions in

each formation to ultimately determine the elastic and strength properties necessary to define

an orthotropic material model for each formation. Pariseau’s laboratory studies focused

solely on the Poorman formation; however, data for other formations along the 4850L and

other areas of the Homestake Mine are included for future reference. Previous knowledge on

the high degree of foliation in the Poorman formation suggests anisotropic properties within

the formation. Orthotropic models of the formations were suggested due to this high degree

of foliation in the region. To obtain data to create these orthotropic models, drill cores were

0

50

100

150

200

250

0

5,000

10,000

15,000

20,000

25,000

30,000

35,000

40,000

Yates Yates/Poorman Poorman Rhyolite

)psi

Rock Unit

Range of UCS Values at Homestake Mine [Lachel Felice, 2015]

Page 82: by Peter Vigilante A thesis submitted in partial

70

taken from parallel to the down dip direction, perpendicular to the foliation and parallel to

strike, with the intention of aligning with the three principal material axes.

Table 10. Original Anisotropic Rock Property Data (*C=Compressive Strength (UCS),

T=Tensile Strength; The 1- and 3-directions are parallel to the schistosity; the 2-

direction is perpendicular to the schistosity. From (Pariseau, 1985)

Property* Poorman Formation

(psi)

Homestake

Formation (psi)

Ellison

Formation (psi)

c1 74.5 117.8 86.5

c2 60.6 79.6 78.1

c3 78.5 98.3 83.0

T1 11.9 9.5 13.6

T2 6.8 7.8 4.1

T3 14.0 8.9 11.7

The original data, shown in Table 10 above, showed significant variation in the

results. Pariseau (1985) removed outliers from the original data, combined with further

laboratory testing conducted by the University of Utah (referred to as UU in Pariseau

(1985)), to produce the modified rock strength properties shown in Table 11. These tests

were conducted on specimens from the Homestake, Poorman, and Ellison Formation’s

hoping to somewhat correct the original rock property data.

Table 11. Modified Anisotropic Rock Strength Property Data (*C=Compressive

Strength (UCS), T=Tensile Strength; The 1- and 3-directions are parallel to the

schistosity; the 2-direction is perpendicular to the schistosity). From (Pariseau, 1985)

Page 83: by Peter Vigilante A thesis submitted in partial

71

Property* Poorman

Formation

(MPa)

Homestake

Formation (MPa)

Ellison

Formation

(MPa)

c1 93.9 138.8 78.1

c2 68.9 79.6 78.6

c3 84.5 91.4 56.2

T1 20.6 9.5 16.2

T2 5.6 7.9 4.1

T3 13.2 13.2 11.4

RESPEC’s Geotechnical Engineering Summary for DUSEL (2010)

RESPEC conducted UCS and Tensile strength measurements on samples taken from

boreholes along the 4850L in the Yates amphibolite as well as the rhyolite. Table 12 below

presents tensile and compressive strength values for the average of two amphibolite

boreholes as well as a rhyolite core. Tensile strength data was measured using the Brazilian

test, while the compressive strength was measured in the uniaxial stress test. Due to large

coefficients of variability, the rocks in this report were considered to be isotropic for the

following UCS and tensile strength data.

Table 12. Compressive and Tensile Strength for Yates amphibolite and Rhyolite from

(RESPEC, 2010)

Rock

Type

Uniaxial Compressive Strength

(MPa) Tensile Strength (MPa)

Mean SD Low High # of

Samples Mean SD Low

Hig

h

# of

Samples

Amphibolite 115 52 33 216 36 14 7 0.2 35 36

Page 84: by Peter Vigilante A thesis submitted in partial

72

Rhyolite 111 55 28 223 18 10 4 5 20 13

The RESPEC (2010) report also conducted direct shear strength tests on Yates

amphibolite intact joints to have data to compare intact and residual Mohr-Coulomb strength

parameters. Table 13 below presents the Mohr-Coulomb parameters of cohesion, So, and

angle of internal friction, ϕ, for both the intact and residual amphibolite.

Table 13. Comparison of Intact and Residual Mohr-Coulomb Strength Parameters for

Joints in Amphibolite from (RESPEC, 2010)

Rock Type Type Cohesion (MPa) Friction Angle (o)

Amphibolite Intact 12.0 47

Amphibolite Residual 3.6 42

Using this data, RESPEC (2010) used a set of equations to then calculate compressive

strength (Co) and tensile strength (To) from the cohesion, So, and angle of internal friction, ϕ

data in Table 13. Using these calculations, Figure 29 was produced to plot the initial, or

ultimate, joint failure in the amphibolites, compared to the average uniaxial compressive

strength (Co) and average tensile strength (To).

Page 85: by Peter Vigilante A thesis submitted in partial

73

Figure 29. Comparison of Ultimate Joint Shear Strength (Represented as Red Triangle

Data Points) to Intact Rock Uniaxial Compression and Brazilian Tensile Strength

(Represented as Blue and Green Curves, Respectively) From (RESPEC, 2010)

Using the initial joint failure strength data and the following calculations, the average

uniaxial compressive strength (Co) is 60.9 MPa and average tensile strength (To) is 9.5 MPa.

For the intact amphibolite, the average Co was 115 MPa and the To was 14 MPa. This joint to

intact rock strength ratio is about 70%-75%, which is considerably high (RESPEC, 2010).

This may provide a rough estimate as to the joint strength of the amphibolite rock, however,

the ratios are suspect due to the fact that the rocks are nonlinear across tension-compression.

Page 86: by Peter Vigilante A thesis submitted in partial

74

Previous Elastic Property Data

Pariseau (1985)

Pariseau (1985) also presented elastic property data on the Poorman, Homestake, and

Ellison formation from the Homestake Mine. These tests were conducted on cores in three

directions to obtain three mutually orthogonal directions in each formation. The purpose of

this was to ultimately determine the elastic properties necessary to define an orthotropic

material model for each formation. Similar to Pariseau’s (1985) UCS and tensile strength

results, previous knowledge on the high degree of foliation in the Poorman formation

suggests anisotropic properties within the formation. Orthotropic models of the formations

were suggested due to the high degree of foliation in the region. In the rock property data

provided by Pariseau (1985), the Shear Modulus was estimated rather than tested in the

laboratory. The Shear Modulus data in Table 14 is estimated using equation 1 below

(Pariseau, 1985).

𝑮𝒊𝒋 =𝑬𝒊

𝟐(𝟏 + 𝒗𝒊𝒋)⁄ (𝟒)

Similar to the UCS and tensile strength data presented in Pariseau (1985), the original

elastic property data showed a large amount of variation in the results. By removing outliers

from the original anisotropic rock property data, combined with supplemental data from the

University of Utah (referred to as UU in Pariseau (1985)), scatter of the results was reduced,

and produced the modified rock strength properties. The ambiguity of these results should be

noted for future referencing.

Page 87: by Peter Vigilante A thesis submitted in partial

75

Table 14. Modified Anisotropic Rock Property Data (*The 1- and 3-directions are

parallel to the schistosity; the 2-direction is perpendicular to the schistosity. All units

are labeled MPa except for Poisson’s ratio (unitless)). E=Young’s Modulus; v=

Poisson’s Ratio; G= Shear Modulus. from (Pariseau, 1985)

Property* Poorman Formation

(MPa)

Homestake

Formation (MPa)

Ellison

Formation

(MPa)

E1 93,015 88,192 89,570

E2 49,608 67,522 63,388

E3 94,393 62,010 75,790

v12 0.23 0.14 0.20

v23 0.15 0.18 0.17

v31 0.22 0.19 0.15

G12 26,182 33,072 31,694

G23 26,871 26,871 28,938

G31 38,584 29,627 35,139

Previous Stress Measurements (Prior to Aug. 2016 kISMET Field Work)

Pariseau (1985)

Pariseau conducted a full-scale vertical crater retreat (VCR) case study between the

6950’-7100’ levels of the Homestake mine. The report also utilized past stress measurements

at various levels reported by Bond (1970), Hooker et al. (1972), and the Spokane Research

Center of the U.S. Bureau of Mines (USBM, 1984) for analysis. The purpose of combining

data from the VCR case study in conjunction with the previous stress measurements was to

provide reliable in-situ stress gradients along with rock property data to provide vital

Page 88: by Peter Vigilante A thesis submitted in partial

76

information for the evaluation of safety and stability in the mine. The stress estimations

produced from this study were based on different locations of the mine than the kISMET

project. Due to the complex geology in the region, the stress field could vary between these

levels and the 4850L location for kISMET. This information was taken into account when

using this data set for preparing the hydraulic fracturing experiment.

Reports before Pariseau (1985) estimated in-situ stress at several levels of the

Homestake Mine. These reports include Bond (1970) along the 6200 level and Hooker et al.

(1972) on the 3050’ and the 6200’ levels, and the U.S. Bureau of Mines along the 7400 level

(1984). Pariseau (1985) found from comparing these results that there was rough agreement

in stress gradient magnitudes, but noticeably different information in the principal stress

directions. Table 15 summarizes the in-situ stress measurements sources and their values, as

well as shows magnitudes of normal stresses relative to coordinate axes parallel and

perpendicular to strike, referred to as the vein stresses. Earlier measurements (Bond, 1970;

Hooker et al.,1972) show principal directions aligned with the vertical, parallel to strike and

perpendicular to strike. USBM (1984) data along the 7400’ level show directions skewed

with respect to the vertical or gravity axis and with respect to local structure. This is

explained in the Pariseau (1985) report to be due to the presence of vertical shear stress

component at the 7400’ level. Thus, it is evident that the principal stress orientation at 7400’

level does not follow the trend above.

Table 15. In-situ stress data at Homestake Mine, including data sources, stresses,

magnitudes, and directions based on Bond (1970), Hooker et al. (1972), and USBM

(1984).

Page 89: by Peter Vigilante A thesis submitted in partial

77

Principal Stresses and Directions

Source Stress* Magnitude (MPa) Bearing Dip**

Bond

(1970) -

6200

Level

Major 55.1 -- Vertical

Intermediate 35.8

N50E 0

Minor 20.0 N40W 0

Hooker et

al. (1972) -

3050

Level

Major 21.0 -- Vertical

Intermediate 25.4 N43E 0

Minor 12.7 N47W 0

Hooker et

al. (1972) -

6200

Level

Major 53.2 -- Vertical

Intermediate 36.9 N30E 0

Minor 25.0 N60W 0

U.S.B.M.

(1984) –

7400

Level

Major 55.0 N83W 53°

Intermediate 23.5 N08W 71°

Minor 13.3 N61E 43°

Pariseau (1985) notes that previous studies by Bond (1970) and Hooker et al. (1972)

were in similar locations (both took measurements along the 6200L), and thus can check one

another for consistency. The data from the 7400 level, however, had not been matched with

any other measurements at the time, so confidence in those stress measurements was not as

strong. Because of this, Pariseau ignored the vertical shear stresses encountered in his

overcoring results, and produced equations for the in-situ stresses in the mine, in part by

setting the vertical normal stress according to a conservative gravity gradient of 1.25 psi per

foot depth.

Pariseau thus formulated equations 5-7 below for vertical in-situ stress, as well as

principal in-situ stresses in the strike and dip directions of the foliation, as a function of

depth, h, in feet or meters. Again, it is important to keep in mind that these gradients were

Page 90: by Peter Vigilante A thesis submitted in partial

78

based on measurements taken from different levels of the Homestake Mine than the 4850L

where the kISMET project took place.

σ𝑣 = 1.25ℎ (Vertical) (5)

σℎ1 = 2078 + 0.53ℎ (Dip direction) (6)

σℎ2 = 121 + 0.55ℎ (Strike direction) (7)

Where h = depth (ft), and stress is in psi.

Or, in SI units:

σ𝑣 = 0.02828ℎ (Vertical) (5’)

σℎ1 = 14.33 + 0.01199ℎ (Dip direction) (6’)

σℎ2 = 0.834 + 0.01244ℎ (Strike direction) (7’)

Where h = depth (m), and stress is in MPa

The Pariseau (1985) report has been influential in estimating in-situ stresses in the

Homestake Mine since many reports have used these equations to estimate stress states at

depths in the mine. Pariseau also simplified the stress gradient equations. Due to the basic

nature of these equations, they most likely represent a rough estimation of the exact in-situ

stress at depths in the mine. However, these equations were used prior to the field work for

determining the in-situ stress at the 4850L. Also, noted by Pariseau, along with most reports

Page 91: by Peter Vigilante A thesis submitted in partial

79

regarding the stress state in the Homestake Mine, the complex geology of the region creates

more heterogeneity thus causing in-situ stress characterization to be difficult.

2009 Stress Measurement Program (Golder Associates, Jan. 2010)

The testing summarized in Golder Associates (2010) report was conducted in 2009

along the 4850 level of the mine. Eight new CSIRO hollow-cell overcoring measurements

were taken in three holes at the location of the DUSEL facility. CSIRO overcoring

(Worotnicki, 1993) method is described as “a reliable triaxial method of obtaining stress”

(RESPEC, 2010). Although not the exact location of the kISMET project, this study is closer

in proximity to the kISMET field site compared to previous stress measurements made in the

Homestake Mine.

Six of the reports measurements were testing the Yates amphibolite unit, and two of

the measurements were testing the rhyolite dikes. This is significant because previous stress

reports were lacking measurements taken from the Yates amphibolite and rhyolite units, but

the kISMET field location was located in the Poorman schist formation. However, this

report is still important to our field work due to the close proximity of this reports work to

our field site. This report also presents data on borehole breakouts observed in portions of the

optical televiewer logs as part of Golder’s 2009 coring program. Breakouts can provide data

on stress directions, and can also be used for quantitative stress determination (Zoback, 2007,

Ch. 8).

In their 2010 report, Golder Associates presented and summarized the 2009 stress

measurements executed by the firm, which were also presented in their preliminary report in

December 2009 titled “In-Situ Stress Measurement Deep Underground Science and

Engineering Laboratory”. Tables 16-18 below summarize the stress measurement results

Page 92: by Peter Vigilante A thesis submitted in partial

80

(Golder, Dec. 2009) published in Golder Associates (2010a) reports. The stress

interpretations were calculated using solutions from Duncan-Fama and Pender (1980) who

assumed homogeneous, elastic isotropic rock when calculating stress from overcoring strains.

Tables 16 and 17 summarize principal stresses and Table 18 summarizes vertical and

horizontal components. Hypothesized results using Pariseau (1985) in-situ stress equations

(equations 5-7) for depths through the Homestake Mine are also calculated in Table 17 for

comparison. Figure 7 below is an image of a borehole breakout taken by Golder Associates

along the 4850L. Figure 8 is a photo depicting the inclination of a borehole breakout along

the 4850L. Breakouts and their significance are summarized below in the “Conclusions Made

by Golder Associates” portion of this thesis.

Table 16. Principal stresses, orientations, and elastic properties (SI units) where SM-02

to SM-07 was in Yates amphibolite unit and SM-08 and 09 are results from rhyolite

dike. From (Golder Associates, 2010a)

Principal Stresses Biaxially Measured

Rock Properties

2009

Tests

(4850

Level)

σ1 σ2 σ3 Young’s

Modulus

Poisson’

s Ratio

MPa

(kpsi) tr pl

MPa

(kps

i)

tr pl

MPa

(kps

i)

tr pl GPa

(Mpsi) -

SM-01 - - - - - - - - - 84.6

(12.3) 0.30

SM-02 43.6

(6.32) 15 18

35.2

(5.1

0)

25

6 56

30.8

(4.4

7)

115 28 67.2

(9.7) 0.29

SM-03

76.8

(11.1

4)

20

9 48

48.7

(7.0

6)

31

0 10

41.5

(6.0

2)

49 40 81.3

(11.8) 0.30

SM-04 59.7

(8.66)

22

4 31

47.9

(6.9

5)

44 59

40.3

(5.8

4)

134 0 - -

Page 93: by Peter Vigilante A thesis submitted in partial

81

SM-05 43.6

(6.32)

18

5 21

27.0

(3.9

2)

27

5 1

19.0

(2.7

6)

8 69 93.7

(13.6) 0.25

SM-06 61.1

(8.86) 37 14

38.7

(5.6

1)

29

6 37

31.7

(4.6

0)

145 49 94.1

(13.6) 0.30

SM-07 34.2

(4.96)

13

3 51

27.5

(3.9

9)

35

9 30

24.9

(3.6

1)

255 23 93.0

(13.5) 0.32

SM-08 60.8

(8.82)

25

8 75

34.4

(4.9

9)

14

5 6

26.9

(3.9

0)

54 14 - -

SM-09 59.5

(8.63)

17

5 75

34.8

(5.0

5)

30

8 10

30.7

(4.4

5)

40 11 61.5

(8.9) 0.36

Averag

es

54.9

(7.96)

36.8

(5.3

3)

30.7

(4.4

6)

81.8

(11.9) 0.30

Table 17. Vertical and horizontal components, octahedral, and deviatoric stresses (SI

Units) where SM-02 to SM-07 was in Yates amphibolite unit and SM-08 and 09 are

results from rhyolite dike along 4850L. From (Golder Associates, 2010a)

Vertical and Horizontal

Components Octahedral

Stresses Deviatoric Principal

Stresses

2009 Tests (4850 Level)

σHMax σHmin σv σoct τoct S1 S2 S3

MPa

(kpsi) Dir.

MPa

(kpsi)

MPa

(kpsi)

MPa

(kpsi)

MPa

(kpsi)

MPa

(kpsi)

MPa

(kpsi)

MPa

(kpsi)

SM-02 42.7

(6.19) 17

31.9 (4.62)

35.1 (5.09)

36.5 (5.30)

5.3 (0.76)

7.1 (1.02)

-1.3 (-0.19)

-5.7 (-0.83)

SM-03 57.7

(8.36) 20

48.1

(6.98)

61.2

(8.87)

55.7

(8.07)

15.1

(2.19)

21.1

(3.06)

-7

(-1.01)

-14.2

(-2.05)

SM-04 56.6

(8.20) 44

40.2

(5.83)

51.1

(7.41)

49.3

(7.15)

7.9

(1.15)

10.4

(1.51)

-1.4

(-0.20)

-9

(-1.31)

SM-05 40.2

(5.83) 4

27

(3.92)

22.3

(3.23)

29.9

(4.33)

10.1

(1.47)

13.7

(1.99)

-2.9

(-0.42)

-10.9

(-1.58)

SM-06 50.1

(7.26) 20

44.1

(6.39)

36.1

(5.23)

43.8

(6.36)

12.4

(1.80)

17.3

(2.50)

-5.1

(-0.74)

-12.1

(-1.76)

SM-07 29.8 148 25.7 31.2 28.9 3.9 5.3 -1.4 -4

Page 94: by Peter Vigilante A thesis submitted in partial

82

(4.32) (3.73) (4.52) (4.19) (0.56) (0.77) (-0.20) (-0.58)

SM-08 38.7

(5.61) 141

28.7

(4.16)

58.7

(8.51)

40.7

(5.90)

14.4

(2.09)

20.1

(2.91)

-6.3

(-0.91)

-13.8

(-2.00)

SM-09 35.8

(5.19) 141

31.5

(4.57)

57.7

(8.37)

41.7

(6.04)

12.6

(1.83)

17.8

(2.59)

-6.9

(-1.00)

-11

(-1.59)

Averages

43.9

(6.37)

34.7

(5.02)

44.2

(6.41)

40.8

(5.92)

10.2

(1.48)

14.1

(2.05)

-4

(-0.58)

-10.1

(-1.46)

Pariseau (1985)

gradient

32.1

(4.65)

19.2

(2.79)

41.8

(6.06)

Table 18. Summary of In Situ Stress Measurements from (Golder Associates Final,

2010b)

Summary of In Situ Stress Measurements

from (Golder Associates Final, Oct. 2010)

Test No. Depth m

(ft)

Vertical Stress

σ2/ σ1 ratio σ3/ σ1 ratio Value

(MPa)

Gradient

(MPa/m)

Hooker et al.

(1972)

950

(3050) 26.29 0.028 0.96 0.48

RI9446 – Site

1

1112

(3650) 28.73 0.026 0.71 0.48

RI9446 – Site

2

1112

(3650) 35.53 0.032 0.69 0.38

Hooker et al.

(1972)

1890

(6200) 53.23 0.028 0.69 0.47

Bond (1970) 1890

(6200) 55.16 0.029 0.65 0.36

NIOSH 2256

(7400) 54.92 0.024 0.75 0.59

Average 0.028 0.74 0.46

Golder Associates Stress Results

SM-02 1478

(4850) 35.1 0.024 0.81 0.71

SM-03 1478

(4850) 61.22 0.041 0.63 0.54

SM-04 1478

(4850) 51.06 0.035 0.8 0.68

Page 95: by Peter Vigilante A thesis submitted in partial

83

SM-05 1478

(4850) 22.3 0.015 0.62 0.44

SM-06 1478

(4850) 38.14 0.026 0.63 0.52

SM-07 1478

(4850) 31.17 0.021 0.8 0.73

SM-08 1478

(4850) 58.67 0.04 0.57 0.44

SM-09 1478

(4850) 57.65 0.039 0.58 0.52

Borehole Breakouts (Golder Associates, Jan. 2010)

Borehole breakouts are fractures along the borehole caused by compressive stress

concentration (Golder Associates, 2010a). The bands formed along borehole walls and

breakouts are often on opposing sides of the borehole in the direction of minimum stress

normal to the borehole axis. Thus, breakouts can be a useful indicator for stress direction, and

were seen in one of the Golder Associates stress measurements holes. Based on the 2009

testing, the breakouts seemed to occur only in the Yates amphibolite close to rhyolite

contacts or in rock mass, which contains Yates amphibolite and rhyolite. An example of a

borehole breakout seen in this study is seen in Figure 30.

To further investigate which of the two sub-horizontal directions is the minor stress

direction, televiewer images were investigated in the boreholes along the 4850 level. The

televiewer images showed evidence of borehole breakouts in the holes drilled at trends near

310°, and no borehole breakouts in the boreholes oriented at trends near 40°. This would

suggest that the borehole breakouts would occur while drilling in the general direction of the

intermediate stress, which would be 310°. It is assumed that borehole breakouts would not

Page 96: by Peter Vigilante A thesis submitted in partial

84

occur while drilling parallel to the minor stress axis, so the general trend of 40° can be

assumed to be the minor stress direction. These televiewer breakout images would thus

suggest that the minor principle stress is in an inferred northeast orientation (Golder

Associates, 2010b). Figure 31 below shows an image of the inclination of a borehole

breakout along a borehole, which had been drilled at an azimuth of approximately 310°.

Figure 30. Image of borehole breakout at 4850 level of Homestake Mine. From (Golder

Associates, 2010b)

Page 97: by Peter Vigilante A thesis submitted in partial

85

Figure 31. Inclination of Breakout in Borehole Drills from the Ventilation Drift

(Borehole Advanced at an Azimuth of Approximately 310°) from (Golder Associates

Final, 2010b)

Conclusions made by Golder Associates (Jan. 2010(a) and Oct. 2010(b) Reports)

Below are conclusions derived from the reports by Golder Associates. It is important

to emphasize again that these conclusions are based on results from tests conducted in the

Yates amphibolite and rhyolite units of the Homestake Mine, while the kISMET project site

was located in the Poorman schist. Although units being tested differed, these conclusions

were made in regions in close proximity to the kISMET project site.

Page 98: by Peter Vigilante A thesis submitted in partial

86

• Stress varied with space in the mine. The Yates amphibolite stress

measurements varied greatly, showing large degrees in variability from tests

within 1 meter of each other. The two rhyolite dikes measured showed

consistency with one another, however, was a smaller sample set than the

amphibolite.

• Data quality and Uncertainties are prevalent in this data set because the

rhyolite data is consistent because the rock is likely isotropic and the

assumptions used in analyzing the overcoring strains hold true. However, the

amphibolite data is variable because they didn’t fully characterize the

anisotropic elastic properties and strain was not properly interpreted.

• Heterogeneity Effects may also affect the rock properties determined in this

report. Heterogeneities could potentially occur on the scale of centimeters or

less, which could cause variability on a stress measurement cell. Larger scale

heterogeneities could affect results spatially. Differing rock types can also

cause stress variations due to contrasts in elastic properties.

• Comparison with Lithostatic Stress and other Stress Indicators can also be an

indicator of the quality of the data measured in the field. The vertical stress at

the 4850 level would be 40.6 MPa for rock with density of 2800 kg per cubic

meter. Pariseau’s (1985) hypothetical calculations would estimate a vertical

stress of 41.8 MPa. Golder Associates testing in 2009 produced an average

vertical stress of 44.2 MPa, and includes both the Yates unit and rhyolite dike

measurements. The rhyolite dike data had vertical stress results higher than

Page 99: by Peter Vigilante A thesis submitted in partial

87

lithostatic and near vertical σ1. Alone, the Yates amphibolite’s average

vertical stress was 39.5 MPa, very close to lithostatic gradient values.

Considering horizontal stress components, figure 10 shows that orientations of

maximum horizontal stress components, σHMax, trend northeast for Yates unit

measurements except SM-07 which trends northwest along with the two

rhyolite measurements. An average of all σHMax values are close to lithostatic

stress, while σhmin are about 80% of lithostatic.

• Borehole Breakouts were seen in one of the Golder Associates boreholes

drilled for overcoring. The direction of breakouts in the borehole suggest that

the 40° was the minor stress axis direction, while 310° was the general

direction of the intermediate stress.

Preparing for kISMET Field Work

For predictive purposes and to approximate what equipment the kISMET project

would need at the 4850L to perform hydrofracture experiments, the figures below in this

preparation section represent previous results and hypotheses for vertical, maximum

horizontal, and minimum horizontal stress throughout the Homestake Mine. These gradients

are extrapolated from the available data presented in this report. This previously published

data is from Pariseau (1985), Johnson et al. (1993), and Golder Associates (Jan. 2010).

Figure 32 shows extrapolated gradients for vertical stress against depth in the Homestake

Mine, and Figure 33 shows specific stress data points from Bond (1970) along the 6200

level, Hooker et al. (1972) along the 3050 and 6200 level, Johnson et al. (1995) along the

3650 level in the Homestake and Poorman Formations, and Golder Associates (Jan. 2010)

Page 100: by Peter Vigilante A thesis submitted in partial

88

along the 4850 level in the Yates Unit and Rhyolite dikes, as well as Pariseau’s (1985)

hypothesized stress gradients for horizontal stress measurements.

Figure 32. Vertical Stress Gradients vs Depth in the Homestake Mine, SD. Sources from

Pariseau (1985), Johnson et al. (1995), and Golder Associates (2010)

0

1000

2000

3000

4000

5000

0 2000 4000 6000 8000

De

pth

(ft

)

Vertical Stress (psi)

Vertical Stress (psi) vs Depth (ft) at Homestake Mine

Vertical Stress-Avgs. From[GolderAssociates,2010] in psi

Vertical Stress-from [Pariseau,1985] in psi

Lithostatic(density=2800kg/m3)

Vertical Stress-Avgs. from[Johnson et al.,1993] in psi

Page 101: by Peter Vigilante A thesis submitted in partial

89

Figure 33. Stress Data Points from Bond (1970), Hooker et al. (1972), Johnson et al.

(1995) and Golder Associates (2010) at Respective Depths, as well as Gradients from

Pariseau (1985)

With tabulated stress data throughout the Homestake Mine, measuring minimum

hoop stress and predicting injection pressure for hydraulic fracturing was done. The

minimum tangential stress around a vertical borehole is given by,

𝜎𝜃𝜃𝑚𝑖𝑛 = 3𝑆ℎ𝑚𝑖𝑛 − 𝑆𝐻𝑚𝑎𝑥 − 2𝑃0 − ∆𝑃 − 𝜎∆𝑇 (𝟓)

Page 102: by Peter Vigilante A thesis submitted in partial

90

where Shmin is the minimum horizontal stress, SHmax is the maximum horizontal stress, P0 is

pore pressure which is zero at 4850L due to water level maintenance, and ∆P is the

difference between the injection in the borehole and formation pressure (Zoback, 2007)

Change to Zoback. By setting the minimum compressional stress in the above equation to the

tensile strength of the rock, one can equate for the pressure in the borehole necessary to

fracture the rock by solving the above equation. Table 19 presents the maximum and

horizontal stress values used in calculating predictions for minimum borehole pressure at the

4850 level of Homestake Mine. Table 20 presents the minimum borehole pressure necessary

for hydraulic fracturing at the 4850 Level for several different instances, and Figure 34

represents this in bar chart form.

Table 19. Data used in Equation 9 to Determine Estimated Minimum Borehole Pressure

Needed at 4850L for Hydraulic Fracturing

Data Source SHmax - MPa (psi) Shmin - MPa (psi)

Golder Associates (2010a)- Averages

43.9 (6,370)

34.7 (5,030)

Golder Associates (2010a) – SM-02

42.7 (6,193)

31.9 (4,627)

Golder Associates (2010a) – SM-03

57.7 (8,369)

48.1 (6,980)

Golder Associates (2010a) – SM-04

56.6 (8,209)

40.2 (5,831)

Golder Associates (2010a) – SM-05

40.2 (5,831)

27.0 (3,916)

Golder Associates (2010a) – SM-06

50.1 (7,266)

44.1 (6,396)

Page 103: by Peter Vigilante A thesis submitted in partial

91

Golder Associates (2010a) – SM-07

29.8 (4,322)

25.7 (3,727)

Golder Associates (2010a) – SM-08

38.7 (5,613)

28.7 (4,163)

Golder Associates (2010a) – SM-09

35.8 (5,192)

31.5 (4,569)

Pariseau (1985) (using gradient)

27.7 (4,020)

14.7 (2,130)

Table 20. Estimated Minimum Borehole Pressure Needed at 4850L for Hydraulic

Fracturing. (SHmax and Shmin values used for calculations are presented in Table 19

above)

Data Source Minimum Borehole Pressure

needed at 4850L – MPa (psi)

Minimum Borehole Pressure needed

at 4850L with Assumed 10 MPa

Tensile Rock Strength – MPa (psi)

Golder Associates

(2010a) - Average

60.2 (8,720)

70.2 (10,170)

Golder Associates

(2010a) – SM-02

53 (7,787)

63 (9,140)

Golder Associates

(2010a) – SM-03

86.6 (12,564)

96.6 (14,011)

Golder Associates

(2010a) – SM-04

64 (9,282)

74 (10,732)

Golder Associates

(2010a) – SM-05

40.8 (5,918)

50.8 (7,368)

Golder Associates

(2010a) – SM-06

82.2 (11,922)

92.2 (13,373)

Page 104: by Peter Vigilante A thesis submitted in partial

92

Golder Associates

(2010a) – SM-07

47.3 (6,860)

57.3 (8,311)

Golder Associates

(2010a) – SM-08

47.4 (6,875)

57.4 (8,325)

Golder Associates

(2010a) – SM-09

58.7 (8,514)

68.7 (9,964)

Pariseau (1985) 16.4

(2,370) 26.4

(3,820)

Figure 34. Estimated Minimum Borehole Pressure Needed at 4850L for Hydraulic

Fracturing

From these results, it shows the Golder Associates data requires a much higher

borehole pressure to fracture the rock at the 4850L. Due to uncertainty in all previous stress

0

20

40

60

80

100

120

Min

. Bo

reh

ole

Pre

ssu

re N

ee

de

d a

t 4

85

0L

(M

Pa

)

Data Source

Predicted Minimum Borehole Pressures Needed for Hydraulic Fractuing at 4850L of Homestake

Mine, SD

Min. PressureNeeded @4850L (WithAssumed 10MPa TensileRockStrength)

Min. PressureNeeded @4850L

Page 105: by Peter Vigilante A thesis submitted in partial

93

measurements done in the Homestake Mine, we prepared for the “worst case scenario” in

which the equipment was suited for the highest pressures. This is why we honored Golder

Associates data which was conducted most recent of the three data sets along the 4850 Level.