You are on page 1of 143

University of Nevada, Reno

Application of Index Test Methods for Intact Rock Strength


Assessment

A thesis submitted in partial fulfillment of the requirements for the degree of


Master of Science in Mining Engineering

By

Pedram Roghanchi

Dr. Raj R. Kallu/Thesis Advisor

August, 2013
THE GRADUATE SCHOOL

We recommend that the thesis


prepared under our supervision by

PEDRAM ROGHANCHI

entitled

Application of Index Test Mehods for Rock Strength Assessment

be accepted in partial fulfillment of the


requirements for the degree of

MASTER OF SCIENCE

Dr. Raj R. Kallu, Advisor

Dr. Jaak J. Daemen, Committee Member

Dr. Robert Watters, Graduate School Representative

Marsha H. Read, Ph. D., Dean, Graduate School

August, 2013
i

Abstract:

Uniaxial compression strength of intact rock is important for engineering geology and

geotechnics, because it is an important design parameter for mines, tunnels, slopes, and

rock foundations. It is also used as input parameter in most the rock mass classification

systems. The difficulties associated with performing direct compression strength tests on

rock leads to indirect test methods for the rock strength assessment. Indirect test methods

are widely used because they are simple, more economical, less time consuming, and

easily adaptable to the field. The main aim of this study is to define correlations between

direct and indirect test methods for core samples from a gold mine in NV. The indirect

test methods are divided in two groups of (1) destructive indirect test methods and (2)

non-destructive indirect test methods. In the destructive methods, point load index (PLI)

tests, splitting tensile strength (Brazilian) tests, and block punch index (BPI) tests are

conducted. In the non-destructive methods, Schmidt hammer and ultrasonic pulse

velocity tests are performed. The results demonstrate that the block punch index test is

not applicable to these rock types. Eleven correlations between the direct and indirect

compression strength tests are developed using linear and nonlinear regression analysis

methods. To evaluate the performance of each regression equation, coefficient of

correlation (R2), variance accounted for (VAF), root mean square error (RMSE) and

mean absolute error (MAE) were calculated. The results show that the splitting tensile

strength has the best correlation relation with the uniaxial compression strength.

Furthermore, the Poisson‟s ratio has no correlation relation with any of the direct and

indirect test results.


ii

Acknowledgement:

I would like to gratefully thank Dr. Raj R. Kallu, assistant professor at UNR, who

without his support and guidance, this was not possible.

I would like to express my deepest appreciation to Dr. Jaak J. Daemen, professor at UNR,

for all of his advice and help for building this thesis. Without his guidance, advice, and

persistent help this thesis would not have been possible.

I would like to acknowledge Mr. John Leland, development technician at UNR, for his

great ideas and suggestions, and his time. Mr. Leland helped me to build my thesis with

his great ideas about laboratory testing methods. I would like to thank him for his

technical assistance during the laboratory testing.

I would like to thank my committee members, Dr. Raj R. Kallu, Dr. Jaak J. Daemen, and

Dr. Robert Waters, for their guidance during my Mater‟s studies at UNR.

In addition, I would like to acknowledge Sean N. Warren and Evan Keffeler, PhD

students at UNR, who provided the core specimens for me. I would like to acknowledge

Dr. Siavash Soroushian, PhD at UNR, for his outstanding ideas about data analysis and

data reductions. I would like to thank him for his support and care.
iii

Table Contents

Introduction

1.1. Intact Rock Strength Test Methods:....................................................................................................... 1


1.1.1. Uniaxial Compressive Strength (UCS) Test: .................................................................................. 2
1.1.2. Splitting Tensile Strength (Brazilian) Test: .................................................................................... 2
1.1.3. Point Load Index (PLI) Test: .......................................................................................................... 3
1.1.4. Schmidt Rebound Hammer (SRH) Test: ........................................................................................ 4
1.1.5. Ultrasonic Pulse Velocity Test: ...................................................................................................... 4
1.2. Objective and Background:.................................................................................................................... 5

Sample Preparation and Testing Results

2.1. Density, Water content, and Slake Durability Index of the Intact Rock: .......................................... 9
2.1.1. Density: ..................................................................................................................................... 9
2.1.2. Water (Moisture) Content of the UCS Specimens: ................................................................. 10
2.1.3. Slake Durability Test on the UCS Specimens: ....................................................................... 11
2.2. Index Test Methods: ....................................................................................................................... 13
2.2.1. Non-Destructive Indirect Test Methods:................................................................................. 13
2.2.1.1. Schmidt hammer test: ......................................................................................................... 13
2.2.1.2. Ultrasonic Pulse Velocity Test:........................................................................................... 15
2.2.2. Destructive Indirect Tests: ...................................................................................................... 16
2.2.2.1. Point Load Index Test: ........................................................................................................ 16
2.2.2.1.1. The Point Load Tester .................................................................................................. 17
2.2.2.1.2. Point load test specimen ............................................................................................... 18
2.2.2.2. Splitting Tensile Strength Test:........................................................................................... 21
2.3. Uniaxial Compression strength tests: .............................................................................................. 24

Correlations between Direct and Indirect Test Methods

3.1. Introduction:......................................................................................................................................... 28
3.2. Regression Analysis: ............................................................................................................................ 29
3.2.1. Simple Linear Regression: ............................................................................................................ 29
3.2.2. Nonlinear Regression: ................................................................................................................... 29
3.2.3. Proposed Regression Models Performance: ................................................................................. 30
iv

3.3. Correlations between the Indirect and Direct Compression Strength Test Methods ........................... 31
3.3.1. Ultimate compression strength (UCS) and Young‟s modulus (E): ............................................... 31
3.3.2. Ultimate compression strength (UCS) and Schmidt rebound hammer (SRH) number: ............... 33
3.3.3. Ultimate compression strength (UCS) and splitting tensile strength (BRZ):................................ 34
3.3.4. Ultimate compression strength (UCS) and Ultrasonic Pulse velocity (P-wave):.......................... 35
3.3.5. Ultimate Compression Strength (UCS) and Point Load Index (Is(50)): .......................................... 36
3.3.6. Young‟s Modulus (E) and Point Load Index (Is(50)): ..................................................................... 37
3.3.7. Young‟s Modulus (E) and Splitting Tensile Strength (BRZ): ...................................................... 37
3.3.8. Splitting Tensile Strength (BRZ) and Point Load Index (Is(50)): ................................................... 38
3.3.9. Splitting Tensile Strength (BRZ) and Schmidt Rebound Hammer (SRH) Number: .................... 39
3.3.10. Young‟s Modulus (E) and Schmidt Rebound Hammer (SRH) Number:.................................... 40
3.3.11. Young‟s Modulus from Uniaxial Compression Test (EUCS) and Young‟s Modulus from
Ultrasonic Velocity Test (EP-wave): .......................................................................................................... 41

The Effect of Cylindrical Core Sample Length on Schmidt Rebound Hammer

(SRH) Value

4.1. Introduction:......................................................................................................................................... 43
4.2. Principal of Schmidt hammer test: .............................................................................................. 44
4.3. Hammer type:....................................................................................................................................... 45
4.4. Significant and use: .............................................................................................................................. 46
4.4. Specimen requirements: ....................................................................................................................... 47

Numerical Simulation of Block Punch Index (BPI)

5.1. Block Punch Index Test: ...................................................................................................................... 52


5.2. Block Punch Index Limitations: .......................................................................................................... 56
5.3. Numerical Simulation of Block Punch Index Test: ............................................................................. 57
5.3.1. Compression only support at the bottom of the specimen: ........................................................... 60
5.3.2. Fixed support at bottom and top of the specimen: ........................................................................ 61
Conclusions:................................................................................................................................................ 64
References ................................................................................................................................................... 68
Appendix-1: Specimen Preparation Report ................................................................................................ 72
Appendix-2: Strain Gages Results ............................................................................................................ 104
v

List of Tables:

Table 2.1. Indirect tensile strength and direct compression specimens density ......................................... 10
Table 2.2. Water content of the UCS specimens ........................................................................................ 11
Table 2.3. Slake durability of the UCS specimens ..................................................................................... 13
Table 2.4. Schmidt rebound hammer (SRH) number of the core specimens ............................................. 14
Table 2.5. Results of the ultrasonic pulse velocity tests on the UCS core specimens ................................ 16
Table 2.6. Results of the point load index test on the available samples ................................................... 21
Table 2.7. The results of the splitting tensile strength................................................................................ 24
Table 2.8. The results of the uniaxial compression strength test................................................................ 27

Table 3.1. Performance of the ultimate strength-Young‟s modulus equation ............................................ 32


Table 3.2. Performance of the SRH- ultimate strength equation ............................................................... 33
Table 3.3. Performance of the splitting tensile strength-ultimate strength equation .................................. 34
Table 3.4. Performance of the P-wave-ultimate strength equation ............................................................ 35
Table 3.5. Performance of the point load index-ultimate strength equation .............................................. 36
Table 3.6. Performance of the point load index-Young‟s modulus equation ............................................. 37
Table 3.7. Performance of the splitting tensile strength-Young‟s modulus equation ................................ 38
Table 3.8. Performance of the splitting tensile strength-point load index equation ................................... 39
Table 3.9. Performance of the splitting tensile strength-SRH equation ..................................................... 40
Table 3.10. Performance of the splitting tensile strength-SRH equation ................................................... 41
Table 3.11. Performance of EUCS-EP-wave equation...................................................................................... 42

Table 4.1. Critical core sample length according to ASTM and ISRM standards procedures ................... 51

Table 5.1. Isotropic properties of the specimens for numerical simulation of BPI test ............................. 59
vi

List of Figures

Fig. 2.1. Slake durability testing machine .................................................................................................. 12


Fig. 2.2. Specimen W-5U before test and after the second cycle ............................................................... 13
Fig. 2.3. Point load tester (left) and point load platen (right) ..................................................................... 18
Fig. 2.4. Diamteral point load test on the specimen W-5U......................................................................... 19
Fig. 2.5. Example of invalid point load test (specimen 7U) ....................................................................... 19
Fig. 2.6. Example of valid point load test (specimen 7U) .......................................................................... 20
Fig. 2.7. Prepared specimens for splitting tensile strength test ................................................................... 22
Fig. 2.8. Brazilian specimen after the splitting tensile strength test ........................................................... 23
Fig. 2.9. Specimen W-1U before and after the uniaxial compression test (4 bonded rosettes) .................. 25
Fig. 2.10. Bonded rosettes on the specimens W-1U after the uniaxial compression test ........................... 25
Fig. 2.11. Specimen W-6U before and after the uniaxial compression test (2 bonded rosettes) ................ 26
Fig. 2.12. Specimens W-6U after the uniaxial compression test ................................................................ 26

Fig. 3.1. Specimen 7U (left) and 5U (right) after the failure ...................................................................... 32
Fig. 3.2. Correlation relation between ultimate strength and Young‟s modulus ........................................ 32
Fig. 3.3. Correlation relation between ultimate strength and SRH ............................................................. 33
Fig. 3.4. Correlation relation between ultimate strength and splitting tensile strength .............................. 34
Fig. 3.5. Correlation relation between ultimate strength and P-wave ......................................................... 35
Fig. 3.6. Correlation relation between ultimate strength and point load index ........................................... 36
Fig. 3.7. Correlation relation between Young‟s modulus and point load index ......................................... 37
Fig. 3.8. Correlation relation between Young‟s modulus and splitting tensile strength ............................. 38
Fig. 3.9. Correlation relation between splitting tensile strength and point load index ............................... 39
Fig. 3.10. Correlation relation between splitting tensile strength and SRH ............................................... 40
Fig. 3.11. Correlation relation between splitting tensile strength and SRH ............................................... 41
Fig. 3.12. Correlation relation between EUCS and EP-wave ............................................................................ 42

Fig. 4.1. Required V-shape steel base for Schmidt rebound hammer test .................................................. 45
Fig. 4.2. Schmidt hammer test requirements based on ASTM standard (a) hammer axis is
perpendicular to the surface test surface, (b) specimen is securely clamped to a steel base....................... 49
Fig. 4.3. Core sample length against SRH value based on ASTM standard procedure .............................. 50
Fig. 4.4. Core sample length against SRH value based on ISRM standard procedure ............................... 50
Fig. 4.5. Cylindrical core samples with different length of 3 cm t 20 cm .................................................. 51
Fig. 4.6. The Effect of Cylindrical Core Sample Length on Schmidt Hammer Hardness Value
(Method for critical core sample length determination) ............................................................................. 51

Fig. 5.1. Block Punch Index Machine (Modified GCTS PLI-100 Model) ....................... 52
Fig. 5.2. Prepared specimens for BPI test ......................................................................... 54
Fig. 5.3. Valid (left) and invalid (right) block punch index tests...................................... 57
vii

Fig. 5.4. Many of the block punch index tests were invalid (according to ISRM suggested
method) since the middle part of the specimen which is under compression load was
broken in many parts. ........................................................................................................ 58
Fig. 5.5. Splitting the top and bottom surfaces of the body sketch to model the supports
and loading plate ............................................................................................................... 59
Fig. 5.6. Mesh refinement at the two parallel planes at the top and bottom of the specimen
........................................................................................................................................... 60
Fig. 5.7. Mohr-Coulomb safety factor at the bottom (left) and top (right) for specimen W-
5U at the 2.5 MPa Pressure with compression support at the bottom of the body ........... 61
Fig. 5.8. Maximum principal stresses (MPa) at bottom (left) and top (right) for the
specimen W-5U with compression support at the bottom of the body ............................. 61
Fig. 5.9. Mohr-Coulomb safety factor at the bottom (left) and top (right) for specimen W-
5U at the 2.5 MPa applied stress with fixed supports ....................................................... 62
Fig. 5.10. Maximum principal stresses (MPa) at bottom (left) and top (right) for the
specimen W-5U with fixed supports................................................................................. 62
1

Introduction:

Rock strength measurement is important for the design of structures in rock as well as for

the strength classification of rock materials. By a better understanding of the rock mass

strength, it is possible to reduce stability problems that may occur due to deeper mining.

One of the most common ways of determining the rock mass strength is by a failure

criterion. The rock mass failure criteria are stress dependent and often include one or

several parameters that describe the rock mass properties. These parameters are based on

classification or characterization systems (Edelbro, 2003). Intact rock strength is an

important input in rock mass classification systems. Strength determinations on rock

usually require careful test setup and specimen preparation, and the results are highly

sensitive to the method and style of loading. An index is useful only if the properties are

reproducible from one laboratory to another and can be measured inexpensively. Indirect

test methods such as point load index (PLI), Schmidt hammer, ultrasonic pulse velocity,

and splitting tensile strength (Brazilian) tests are widely used because they are simple,

more economical, less time consuming, and easily adaptable to the field (Aksoy et al.,

2011).

1.1. Intact Rock Strength Test Methods:

Strength characteristics of rock are very important parameters for rock mass classification

and design of the structures in rock. Tests developed to measure rock strength properties,

whether direct or indirect tests are limited by sample availability. In some cases, core

samples having a sufficient height cannot be obtained because of the rock mass
2

properties. For this reason, many testing methods have been proposed for indirect

unconfined strength determination.

1.1.1. Uniaxial Compressive Strength (UCS) Test:

The UCS is an important input parameter in rock mass classifications. Unconfined

compression is the most frequently used strength test for rocks; however, it is difficult to

perform and results can vary as procedures are varied. The test specimen should be a rock

cylinder of length-to-diameter ratio in the range 2 to 2.5 with flat, smooth, and parallel

ends cut perpendicularly to the cylinder axis. Procedures are given in ASTM D-2938

standard (2010) and by Bieniawski and Bernede (Goodman 1989). Measurement of rock

strength requires testing which must be undertaken on test specimens of particular size in

order to meet testing standards. A standard UCS test requires high quality core samples.

Recommended core samples cannot be obtained particularly from “weak, stratified and

fractured rock.” Often, the drilling process breaks up the weaker core pieces, and they are

too thin or fragmented to be used in either uniaxial compressive or splitting tensile

strength tests (Sulukcu & Ulusay, 2001).

1.1.2. Splitting Tensile Strength (Brazilian) Test:

The Brazilian test, described in the ASTM D-3967 standard (2008), is convenient for

gaining an estimate of the tensile strength of rock. “It has been found that a rock core

about as long as its diameter will split along the diameter and parallel to the cylinder axis

when loaded on its side in a compression machine” (Goodman, 1989). The reason for this

can be demonstrated by examining the stress inside a disk loaded at opposite sides of a

diametric plane (Goodman, 1989). Theoretically, the tensile failure occurs along the

loaded diameter, splitting the disc (or cylinder) into two halves. However, in many cases,
3

the fractures do not go through the center and separate the disc in two halves, as the

simple theory predicts. Also, in many cases, the influence of orientation angle is much

larger than that of foliation-loading angle (Quac et al. 2013).

1.1.3. Point Load Index (PLI) Test:

In the point load test, a rock is loaded between hardened steel cones, causing failure by

the development of tensile cracks parallel to the axis of loading. The test is an outgrowth

of experiments with compression of irregular pieces of rock in which it was found that

the shape and size effects were relatively small and could be accounted for, and in which

the failure usually was caused by induced tension (Goodman 1989).

Tests are done on pieces of drill core at least 1.4 times as long as the diameter. In practice

there is a strength/size effect; so a correction must be made to reduce results to a common

size (Goodman 1989, Schrier 1988). Although shortcomings related to this method have

been reported in many papers, it is being still used to predict UCS (Deere & Miller 1966,

Fener et al. 2005).

Shortcomings, limitations and problems related to the point load index test are as follows:

“(a) tested specimens are generally anisotropic and heterogenic, but tests are applied in

very small area; (b) irregular failures (invalid test results) frequently occur and cause a

requirement of too many rock specimens; (c) the specimen may move during loading;

and (d) micro-fissures may cross the conical platens. The main problem is sourced from

the more or less heterogeneity or anisotropy, thus the point load index test should be

preferably conducted on at least ten tests. If the rock is heterogeneous or anisotropic, the

test number should be more than ten. Modes of failures are also important for a valid test.

If the rock is broken as invalid shape, the test should be rejected” (Yalmiz 2009).
4

1.1.4. Schmidt Rebound Hammer (SRH) Test:

Schmidt hammer has been used worldwide as an index test for a quick rock strength and

deformability characterization due to its rapidity and easiness in execution, simplicity,

portability, low cost and non-destructiveness. The Schmidt hammer was originally

developed for measuring the strength of hardened concrete but it can also be correlated

with rock compressive strength according to Miller (1965) and Barton and Choubey

(1977). The principle of the test is based on the absorption of part of the spring-released

energy through plastic deformation of the rock surface, while the remaining elastic

energy causes the actual rebound of the hammer. The main shortcomings, limitations and

problems related to this testing method are “(a) anisotropy and heterogeneity of the rocks,

very small test conduction area, (b) roughness on the surfaces where the test is applied,

(c) vibration in the rock during test may set the specimen in motion, (d) test direction, and

(e) there have been a number of different empirical equations proposed for different types

of rocks (Yalmiz, 2009).”

1.1.5. Ultrasonic Pulse Velocity Test:

Ultrasonic pulse velocity testing is a useful and reliable nondestructive tool for assessing

the mechanical characteristics of rock, such as the modulus of elasticity and the

compressive strength. The velocity of ultrasonic pulses traveling in a solid material

depends on the density and elastic properties of that material. The quality of some

materials is sometimes related to their elastic stiffness so that measurement of ultrasonic

pulse velocity in such materials can often be used to indicate their quality as well as to

determine elastic properties (Sharma & Singh 2008).


5

1.2. Objective and Background:

The main objective of this study is to develop correlations between the direct and indirect

test methods. Several correlations have been reported by many authors for different rock

types in various conditions. Sharma et al. (2011) established a statistical relationship

between Schmidt hammer rebound numbers with impact strength index (ISI), slake

durability index (SDI) and P-wave velocity. They determined these index properties in

the laboratory. Each index property was correlated with Schmidt hammer rebound values

for granite, basalt, andesite, sandstone, and quartz. Representative rock mass samples

were collected from the site to carry out other tests in the laboratory. During sample

collection, each block was inspected for macroscopic defects so that it would provide test

specimens free from fractures and joints. The equation of the best fit line and the

coefficient of determination (R2) were determined for each set of test results. It was found

that Schmidt hammer rebound numbers show linear relationship with ISI and SDI, but

shows exponential relationship with P-wave velocity. T-test was done to verify the

correlation between rebound values and other rock index properties. The results from the

T-test show higher calculated values, for each relationship, than the tabulated values.

Hence, they all have significantly strong correlation among themselves.

Cobanoglu & Celik (2008) developed a correlation of uniaxial compressive strength test

with indirect strength test methods. They prepared 150 cores from the sandstone,

limestone and cement mortar, with five different diameters: 54, 48, 42, 30 and 21 mm.

For the uniaxial compressive strength test, the tests were carried out using a loading rate

of 0.5 MPa/s. Five tests were undertaken for each core size of each material type. The

study of the relationship between core diameter and the UCS values showed that the
6

highest value for the limestone was obtained on the 21 mm sample and the lowest on the

54 mm sample. A total of 15 core samples were tested using the L-type Schmidt hammer

and a rock cradle following ISRM and the relationship between Schmidt hammer

rebound number and uniaxial compressive strength was found. The relationship between

the Is(50) and UCS values was obtained in this study for the five core diameters. Simple

regression analyses were performed to define type of the relationship between dependent

and independent parameters by considering linear functions. Also, multiple regression

analysis was done to evaluate the UCS of rock based on five index tests which were

performed in this study. The results are summarized as: UCS = 4.14 Is(50) + 29.8 Vp +

0.54 SHR – 116 (SHR stands for Schmidt hammer rebound number and Vp stands for

sonic velocity). The validity of the proposed equations is limited by the data range and

sample types which were used to derive the equations.

Potro and Hurlimann (2009) compared different direct and indirect compression test

methods to evaluate intact rock strength of volcanic rocks. Schmidt hammer, point load

test, and uniaxial compressive strength were performed on the core samples (or in the

field) from a volcanic area in the North Atlantic Ocean. L-type and N-type Schmidt

hammers were used to estimate rock strength in the thirty four different locations in the

area of interest. An error analysis for both Schmidt hammer tests showed they both have

very similar coefficient of variation distributions. Correlations between unit weight and

(N-type rebound number) RN and (L-type rebound number) RL were proposed. RN

provides very good linear correlation with unit weight; however, R L is less sensitive to

variations in unit weights. In this study, 152 point load tests were performed following

ISRM recommended procedure and ASTM standards. Strength values from point load
7

tests failed to give a relation with unit weight. Uniaxial compressive strength results

showed a good logarithmic relation with unit weight.

Kahraman et al. (2005) defined a correlation between uniaxial compressive strength and

point load index. The authors indicated that the derived ratios between UCS and Is exhibit

a very large range; the ratio for the equations using the zero intercept varies between 8.6

to 29. In this study, 38 different rock types were sampled, 11 of which were igneous, 9

were metamorphic, and 18 were sedimentary. The UCS tests were conducted on samples

with 38 mm diameter and L/D ratio of 2. Each test was repeated at least five times for all

the rock types and the average value was recorded as the UCS. The correlation between

UCS and Is values proposed: UCS=10.91Is+27.41 (R2=0.61). The authors studied the

relation between UCS/Is and n (porosity) for the tested rocks. They could not find any

relation between UCS/Is and n. To study the effect of porosity, the authors divided the

tested rock into two groups according to n values: n<1% and n>1%. The equations of the

correlations between UCS and Is proposed for these two groups were: UCS=24.83 Is-

39.64 (n<1%) and UCS=10.22Is+24.31 (n>1%).

The main aim of this study is to develop correlations between the direct and indirect

compression test methods for the core samples available from a gold mine in NV. There

was no information about the RQD (Rock Quality Designation) and depth of the core

samples. The core samples were basalts and rhyolite. Point load index (PLI), splitting

tensile strength (Brazilian Test), block punch index (BPI), Schmidt hammer test, and

ultrasonic pulse velocity test were performed.

In Chapter 2, the sample preparation, testing procedures, and testing results are described.

Indirect tests are divided into two groups, destructive and nondestructive methods. For
8

each indirect test, the number of tests, the number of valid tests and summary of the

results are given.

Chapter 3 presents the correlations between the direct and indirect test results. Simple

regression analysis is used to relate the direct and indirect test methods. To evaluate the

performance of each regression equation, coefficient of correlation (R2), variance

accounted for (VAF), logarithmic standard deviation (β), root mean square error (RMSE),

and mean absolute error (MAE) are calculated.

In Chapter 4, the effect of the core specimen length on the Schmidt rebound hammer

(SRH) is studied. In order to assess the effect of the length on the SRH value, HQ (63.5

mm diameter) core samples having lengths of 3, 5, 6, 8, 10, 12, 15, and 20 cm, from three

different rocks were prepared. The aim of the study was to validate the results of the

Schmidt hammer number which were used for the correlation relations.

Chapter 5 presents numerical simulation of the block punch index (BPI) test. In this

study, the BPI test was used as an indirect test method. The aim of the chapter is to

simulate the block punch index for moderate to strong intact rock.

Chapter 6 presents the conclusions and recommendations for future work for other

researchers who may wish to investigate the correlation relation between the direct and

indirect test methods.


9

Specimen Preparation and Testing Results:

Samples were prepared from core obtained from a NV gold mine. All attempts were

made to obtain index test specimens right above and/or right below the UCS specimens

for direct comparison. In some cases, one specimen, either indirect tensile strength or

point load strength, was selected either below or above the UCS sample. In some cases,

index test specimens were not available directly above or below the UCS specimen.

However all index specimens were cut from the same core and associated with the closest

UCS specimen. In a few cases, the core samples were not long enough to get enough

specimens for both direct and indirect tests methods. In these cases, the UCS core

samples were not included in the calculations.

2.1. Density, Water content, and Slake Durability Index of the Intact

Rock:

2.1.1. Density:

Prior to the destructive index testing, the density of splitting tensile test specimens and

uniaxial compression strength test specimens was determined. The splitting tensile test

specimens were weighed in an air-dried condition. The average density is considered for

the Brazilian splitting test specimens which came from the same sample ID.

The UCS test specimens were oven-dried to determine the density of the rock. The

volume of the specimen was computed by measuring the diameter of the rock specimen

at three locations and the thickness of the rock specimen at three locations. The average
10

thickness and diameter of the specimen was computed and the volume of the specimen

determined (Table 2.1).

The density of the specimen is simply the weight divided by the volume:

(2.1)

Table 2.1. Indirect tensile strength and direct compression specimens density

Density (g/cm3)
Sample ID UCS specimen Indirect tensile test specimen
2U 2.45 2.08
5U 2.53 2.54
6U 2.35 2.40
7U 2.47 2.55
8U 2.40 2.63
11U 2.25 2.39
W-1U 2.55 2.53
W-2U 2.10 2.08
W-5U 2.07 2.33
W-6U 1.96 2.13

2.1.2. Water (Moisture) Content of the UCS Specimens:

“Water content by mass is the ratio of the mass of water contained in the pore spaces of

soil or rock material, to the solid mass of particles in that material, expressed as a

percentage” (ASTM D2216, 2008). The water content was calculated for all UCS

specimens. The procedure of the water (moisture) content calculation for rock and soil is

described in the ASTM D-2216. For all the specimens, the weights were recorded after

sawing the specimens. The specimens were dried in the oven for at least 12 hours at the

temperature of 105 ± 2 oC. After the drying, weights of the dried specimens were

recorded. This procedure was performed two to three times to get a constant mass for all

the specimens. The water content by mass was recorded to the nearest 1 %. Table 2.2

shows the water content of the UCS specimens. As was expected, the water contents of
11

the UCS specimens were not considerable because the core samples were left in the

laboratory for several days, and most of the specimens were dried in the air-dried

condition.

Table 2.2. Water content of the UCS specimens

Sample ID Moisture content (%)


2U 0.32
5U 0.17
6U 0.79
7U 0.34
8U 0.49
11U 1.67
W-1U 0.67
W-2U 1.91
W-5U 4.34
W-6U 4.14

2.1.3. Slake Durability Test on the UCS Specimens:

The slake durability test, devised by Franklin and Chandra (1972), has proved to be

particularly suitable for evaluating the wide range of rock durability conditions

encountered in the field. (Hoek 1977).

The slake durability test consists of testing ten specimens weighting between 40 g to 60

g. The total test specimen should weigh 450 g to 550 g (ASTM D4644, 2008).
12

Fig. 2.1. Slake durability testing machine

All the specimens for each sample were weighed before the slake durability test, and then

the specimens were placed in a drum made of 2 mm square-mesh. The slake durability

drum rotated at 20 rpm for ten minutes. In the next step, the drum was placed in oven

with temperature of 105 ± 2 oC for 24 hours. The procedure was repeated one more time.

The final oven-dried mass for the second cycle was recorded for each sample. The slake

durability index is calculated as follows:

Id (2) = [(WF – C) / (B – C)] × 100 (2.2)

Where:

Id (2) = slake durability index in the second cycle,

B = Mass of drum plus oven-dried specimen before the first cycle (g),

WF = Mass of drum plus oven-dried specimen retained after the second cycle (g),

C= Mass of drum (g).

Table 2.3 represents the slake durability index for each UCS core samples. As was

expected the slake durability index is not considerable because the specimens are strong.
13

The slake durability test is usually used to estimate the durability of weak rocks. The

procedure of the slake durability index calculation for shale and similar weak rocks is

described in the ASTM D-4644.

Fig. 2.2. Specimen W-5U before test and after the second cycle

Table 2.3. Slake durability of the UCS specimens

Sample ID Slake durability index (%)


2U 99.10
5U 97.39
6U 99.10
7U 97.39
8U 99.51
11U 98.37
W-1U 99.30
W-2U 99.39
W-5U 94.10
W-6U 90.73

2.2. Index Test Methods:

2.2.1. Non-Destructive Indirect Test Methods:

2.2.1.1. Schmidt hammer test:

The Schmidt hammer test is a non-destructive, easy, inexpensive testing method which

can be performed both in the laboratory and in the field. The effects of cylindrical core

sample length on the Schmidt hammer hardness value are comprehensively explained in
14

Chapter 4. The Schmidt hammer tests have been performed on all the UCS core samples.

The test procedure of the Schmidt hammer test is described in the ASTM D-5873 (2005).

The Schmidt Rebound Hammer (SRH) numbers were reduced based on the ASTM D-

5873 standard. A total of 20 readings were recorded from each UCS specimen from

different locations separated by at least the diameter of the piston.

Ten readings were chosen randomly from the total readings. If the difference between a

reading and average of ten reading was more than seven units, that reading was

discarded, and the average of the remaining readings was determined.

Table 2.4 shows the results of the Schmidt hammer tests on the specimens. The SRH

number for specimen 11U seems to be influenced by the weakness plane in this

specimen. The number is relatively high for this specimen as compared with the other

indirect tests and direct compression test. Further, the SRH number for the specimen 5U

is relatively less than the number that was expected.

Table 2.4. Schmidt rebound hammer (SRH) number of the core specimens

Sample ID SRH Number


2U 50
5U 44
6U 46
7U 53
8U 50
11U 48
W-1U 50
W-2U 27
W-5U 18
W-6U 18
15

2.2.1.2. Ultrasonic Pulse Velocity Test:

Compression wave velocity is the dilational wave velocity which is the propagation

velocity of a longitudinal wave in a medium that is effectively infinite in lateral extent

(2008). In this project, the GCTS model ULT-100 was used as an ultrasonic velocity test

system. GCTS recommended using ordinary honey as an acoustic compliant, and this

seems to provide reasonable results. GCTS also recommended applying a slight load to

the platens. There were two NX-size test platens, with dual P- and S- wave capabilities

(CATS Ultrasonic 1.95 User’s Guide and Reference). A steel disc was used which weighs

about 17 lbs. A small amount of honey was applied to a platen as the bottom platen. The

sample was pressed onto the honey, and rotated slightly to spread the honey to a thin

layer. The honey was applied to the free end of the specimen and the other platen onto the

specimen was pressed. The P- and S-wave were measured using the CATS software of

the GCTS Company.

The results from the CATS software usually are not accurate. The first arrival indicator

line was set manually. The selection of the “first arrival time” is slightly different

between the P-wave and the S-wave. For the P-wave, typically the waveform is very

horizontal initially, and then drops off toward a negative peak. For the S-wave the

waveform starts off horizontal as well. However, there are two distinct, very low-

amplitude “wiggles,” before the start of a more substantial peak. The “wiggles” are

apparently artifacts and are not part of the true S-wave signal. For the S-wave, the cursor

should be set beyond the “wiggles” to the start of the first truly distinct peak. Table 2.5

shows the results of ultrasonic velocity test on the UCS specimens.


16

Table 2.5. Results of the ultrasonic pulse velocity tests on the UCS core specimens

Sample ID P-wave (m/s) S-wave (m/s) Poisson’s Ratio Young’s Modulus (Gpa)
2U 5220 2791 -- 54.573
5U 5320 2360 0.38 38.709
6U 4901 2381 0.35 35.839
7U 5089 2696 -- 46.811
8U 4979 2184 0.38 31.812
11U 3812 2100 0.28 25.401
W-1U 4477 2224 0.34 32.581
W-2U 4463 2305 0.32 29.439
W-5U 3134 1758 0.27 16.225
W-6U 3250 1804 0.28 16.322

2.2.2. Destructive Indirect Tests:

2.2.2.1. Point Load Index Test:

The point load test is used as a quick and inexpensive means of obtaining a quantitative

rock strength index while logging core (Hoek, 1977). The point load test was originally

proposed (Broch & Franklin 1972) as a means of providing for destructive strength

testing of hard rock materials with a portable apparatus, such that the tests produced a

field strength index which could be correlated with UCS. Much of the costly laboratory

testing requiring large, stationary machines could be avoided in rock site characterization

(Smith, 1997).

The point load test loading geometry produces a failure mode which closely

approximates a tensile failure, and of course does correlate well with the uniaxial tensile

or the Brazilian tensile test strength (Bieniawski, 1975). Accordingly, correlation of point

load strength with unconfined compressive strength could be expected to closely follow

the tensile strength to unconfined compressive strength correlation for a given material.
17

The advantages of this index test are: (1) Smaller forces are needed so that a small and

portable testing machine may be used. (2) Specimens in the form of core or irregular

lumps are used and require no machining (3) Fragile or broken materials may be tested.

The point load test cannot be used on very soft rocks such as mudstone or claystone or on

soft evaporates. When used on anisotropic rocks such as slate, considerable care has to be

taken to ensure that the loading direction is either parallel to or perpendicular to the

dominant weakness direction. One disadvantage of the point load test, shared with all

other strength tests, is that the core is fractured and this can lead to confusion in

subsequent interpretation of fractures in the core. It is important that photography of the

core and determination of the RQD (Rock Quality Designation which is based upon the

number of intact core pieces of more than 10cm length) be carried out before the core is

used for point load tests (Hoek 1977, Bieniawski 1975, & Broch & Franklin 1972).

2.2.2.1.1. The Point Load Tester

Full description of the point load apparatus is available in the ASTM D5731-08. The

apparatus consists of a loading system typically comprised of a loading frame, platens, a

measuring system for indicating load, P, (required to break the specimen), and a means

for measuring the distance, D, between the two platen contact points. The equipment

should be resistant to shock and vibration so that the accuracy of readings is not

adversely affected by repeated testing. The two platens have 60 degree conical points

with 5 mm point radius. The platens should be of hard material such that they remain

undamaged during testing. A typical load capacity should be more than adequate to fail

the higher strength rocks when testing NX-size (54-mm) core (Smith, 1997; Heidari etal,

2012; Broch & Franklin, 1972).


18

Fig. 2.3. Point load tester (left) and point load platen (right)

2.2.2.1.2. Point load test specimen

In general, four different point load tests can be performed according to specimen shape:

(a) the Diametral Test, (b) the Axial Test, (c) the Block Test, and (d) the Irregular Lump

Test. Usually, core samples are available for laboratory testing; however, block tests can

be performed if such samples are available. “Diametral point load test is the convenient

method of determining the uniaxial compressive strength of rock materials for strength

classification purposes” (Bieniawski, 1975). Previous work by Broch and Franklin (1972)

also revealed that the size and shape effects are very pronounced in the axial and irregular

methods (Heidari et al., 2012). It is recommended that core diameters of less than “BX

size (42 mm diameter) should not be used for point load testing” because for smaller

diameters the loading points cannot be considered as theoretical "points" in relation to the

specimen size (Bieniawski, 1975).

One hundred forty three point load tests have been performed on the core samples to find

out the average point load indexes for the available UCS specimen. The point load test

and the test procedure are comprehensively described in the ASTM D-5731 (2008). In

this work, the specimens were tested in diametral, axial, and irregular configurations.

Forty three diametral, twenty five axial, and one hundred and five irregular point load
19

tests were performed on the different core samples from the available UCS specimens.

The number of point load tests is not the same for all the specimens. For example,

nineteen point load tests were performed on the core sample W-5U. For the specimens

5U and 7U, there is no point load data since the core samples were too strong and the

point load device could not break them.

During the test the load, as applied by the hand pump, is steadily increased such that

failure of the specimen occurs with 10 to 60 seconds. Based on the ASTM D-5731

standard, part of the point load tests were considered as invalid, and rejected. In the end,

the point load tests were reduced to ninety five valid point load tests.

Fig. 2.4. Diamteral point load test on the specimen W-5U

Fig. 2.5. Example of invalid point load test (specimen 7U)


20

Fig. 2.6. Example of valid point load test (specimen 7U)

Broch & Franklin (1972) proposed that the strength index at any available core diameter

should be corrected to a value Is(50) as a reference diameter of 50 mm. This corrected

index is now required by ASTM D-5731 standard.

The uncorrected point load strength is calculated as:

Is= P/De2 (MPa) (2.3)


where:

P = failure load, N,

De = equivalent core diameter:

De2 = D2 for diametral core tests without penetration, mm2,

De2 = 4A/π for axial, block, and lump tests, mm2;

And A=WD (Minimum cross sectional area of a plane through the platen contact points).

Size effects must be taken into account in any strength classification whose function is to

compare test results from a variety of sources. The size correction factor (F) can be

obtained as follows:

F= (De/50)0.45 (2.4)

And the corrected point load strength is:

Is(50)=F×Is (2.5)
21

Table 2.6. Results of the point load index test on the available samples

Sample ID Is(50) (MPa)


2U 1.8
5U --
6U 1.64
7U --
8U 1.99
11U 0.97
W-1U 2.64
W-2U 1.84
W-5U 0.49
W-6U 0.36

2.2.2.2. Splitting Tensile Strength Test:

The Brazilian test is a simple, inexpensive, and desirable alternative test method which

can be used as indirect tensile strength test in rock mechanics studies (Mishra & Basu

2012). Engineers involved in rock mechanics design usually deal with complicated stress

fields, including various combinations of compressive and tensile stress fields. Under

such conditions, “the tensile strength should be obtained with the presence of

compressive stresses to be representative of the field conditions” (Hoek 1977).

The Brazilian strength test (splitting tensile test) is a well-known indirect method of

determining the tensile strength of rocks and other brittle materials using a circular plane

disc under diametral compression. It is widely applied in rock engineering because

specimens are easy to prepare, the test is easy to conduct and uniaxial compression test

machines are quite common.

The contact conditions created under the Brazilian test are the principal means by which

the experimenter exerts an influence on the specimen. The main issue during an

experiment is that it is the contact conditions themselves which determine whether

failure will initiate in the central part of the specimen (in which case the experiment will
22

be valid) or under the loading devices( in which case the experiment should be

considered as invalid) (Andreev 1991).

The Brazilian test method and procedure is described in the ASTM D-3967 (2008). In

this project, An MTS machine is utilized to perform the splitting tensile test on the twenty

two Brazilian specimens from the available core.

Fig. 2.7. Prepared specimens for splitting tensile strength test

Since it is widely used in practice, a uniform load is needed for data to be comparable. A

uniform load is also needed to ensure that the disk specimens break diametrally due to

tensile pulling along the loading diameter. On the other hand, in many cases, the indirect

tensile strength testing was conducted using an apparatus consisting of a hand hydraulic

pump and small load frame. Although the attempt is to apply the load uniformly, the

applied force to the specimen is not completely uniform in comparison with the MTS

machine. The splitting tensile strength test conducted with the MTS machine becomes a

time consuming and a difficult test, since the MTS machine should be set up by an

expert. This reason and the limitations in the core sample length narrow the number of

tests on many of the IDs related to the UCS specimens.

The uniform load is applied to the specimen with a constant rate of 0.16 mm/min for all

the Brazilian specimens. The bottom platen moved upwards under the action of the
23

hydraulic pump. This test method is intended to indirectly measure direct tensile strength

of a rock specimen of regular geometry.

The splitting tensile strength of the specimen is calculated as follows:

σt = 2P/πLD (2.6)

where:

σt = splitting tensile strength (MPa)

P = maximum applied load indicated by the testing machine (N)

L = thickness of the specimen, mm

D = diameter of the specimen, mm

Table 2.7 shows the results of the splitting tensile strength tests. The specimen diameter

and thickness were considered as an average of three measurements.

Fig. 2.8. Brazilian specimen after the splitting tensile strength test
24

Table 2.7. The results of the splitting tensile strength

Sample ID σt (MPa)
2U 14.90
5U 29.35
6U 16.92
7U 22.19
8U 26.00
11U 7.50
W-1U 32.23
W-2U 12.73
W-5U 9.73
W-6U 5.18

2.3. Uniaxial Compression strength tests:

Uniaxial compression strength is one of the key parameters for rock engineering design

and rock mass characterization. On the other hand, ultimate strength, ultimate strain, and

mechanical properties of the intact rock are the most important parameters for numerical

design in many finite element rock mechanics softwares. Uniaxial compression test was

conducted on the ten UCS specimens from the available core samples. The procedures of

the specimen preparation are described in Appendix 1. Specimens 2U and 6U were

shorter than the ASTM D-4543 standard requirement. The ASTM D-7012 requires

specimen length of 2 to 2.5 time of the diameter of the core specimen (L ≥ 2D).

Furthermore, the specimen W-2U did not meet the ASTM D-4543 dimentional

requirement (Side Smoothness and Straightness).

A total of 10 UCS core samples were tested which were divided to three main groups:

 Three core samples with four rosettes (8 strain gages) to assess if the strain gages

(CEA-13-125WT-350 rosettes) are applicable to install on the HQ (63.39 mm)

core samples.
25

 Five core samples with two rosettes to develop correlation relations of the

mechanical properties of the rock with indirect test methods.

 Two core sample without any strain gages.

The results from the strain gage studies are comprehensively explained in the Appendix

2. It was concluded that the CEA-13-125WT-350 rosettes are fairly accurate to apply to

the HQ core specimens.

Fig. 2.9. Specimen W-1U before and after the uniaxial compression test (4 bonded rosettes)

Fig. 2.10. Bonded rosettes on the specimens W-1U after the uniaxial compression test

To calculate the mechanical properties of the intact rock sample (E and ν), the slope of

axial and lateral curves are measured based on the methods suggested by ASTM standard
26

(ASTM D7012, 2010). Two rosettes were bonded to each sample with 180o degree angle

to each other.

Fig. 2.11. Specimen W-6U before and after the uniaxial compression test (2 bonded rosettes)

Fig. 2.12. Specimens W-6U after the uniaxial compression test

The value of the Young‟s modulus and Poisson‟s ration are calculated using three

methods as follow:

 Tangent modulus, at a stress level at 50% of the maximum strength.

 Secant modulus, from 5% of maximum stress to 95% maximum strength.

 Average modulus, slope of the straight line portion of the stress-strain curve (with

the r = 0.95).
27

Comparison between the errors of each method for five different samples indicates that

the “Average Modulus of Linear Portion of Axial Stress-Strain Curve” is the best method

for Young‟s modulus and Poisson‟s ratio calculations.

Furthermore, the recorded deformations with the MTS machine were compared with the

strain gages recorded strain. Comparison between the recorded displacement/strain from

the MTS machine and the strain gages shows that the displacement recorded with the

MTS machine is much larger than the actual strain/displacement from the strain gages.

To find out if there is any (mathematical) relation between the recorded displacement on

the MTS machine and strain gages, the maximum, minimum, and average recorded strain

from the gages are plotted against the maximum, minimum, and average recorded strain

with the MTS machine. A correction factor for the deformation recorded with MTS is

calculated based on the slope of linear curve of all the considered displacements which is

equal to “J = 2.25”. The results of the uniaxial compression strength tests are listed in the

table 2.8.

Table 2.8. The results of the uniaxial compression strength test

Sample ID Number of Ultimate strength Young’s Modulus* Poisson’s Ratio


rosettes (MPa) (GPa)
2U 0 212.89 59.47
5U 4 375.62 62.23 0.24
6U 2 170.68 42.04 0.25
7U 0 199.99 55.39
8U 2 214.67 48.80 0.18
11U 4 91.58 28.81 0.26
W-1U 4 215.39 45.81 0.19
W-2U 2 107.67 28.95 0.20
W-5U 2 39.08 17.40 0.18
W-6U 2 39.59 18.12 0.18
* Young‟s modulus based on “Average Modulus of Linear Portion of Axial Stress-Strain Curve” method
28

Correlations between the Indirect and Direct Compression


Strength Test Methods

3.1. Introduction:

The difficulties associated with performing direct compression strength tests on rock lead

to indirect test methods for the rock strength assessment. Indirect test methods are widely

used because they are simple, more economical, less time consuming, and easily

adaptable to the field. The indirect compression test methods can be divided into two

destructive and non-destructive test methods. There are advantages and disadvantages

when one indirect test method is applied for the intact rock strength assessment. There

are different standards and technical notes on how to use the indirect tests in the project.

Furthermore, there are some uncertainties due to specimen availability, test procedure,

and data analyzing. On top of that, it is very important to study if an indirect test method

is applicable to the rock type and the project. For example, the block punch index (BPI)

test method is an easy and economical test which has been developed as an indirect

strength test. The BPI test is applicable to weak rock in which it is not possible to get a

core specimen long enough for direct compression testing. The point load index (PLI) test

is not applicable to weak rock strength assessment. Furthermore, for very strong rocks,

the point load cannot be used since there is a possibility to damage the point load

machine. A similar problem exists for the Schmidt hammer test; Schmidt hammer is not

recommended to use as an indirect test for evaluation of intact rock strength for weak

rock.

It is important to study how the indirect test methods can be related to the actual strength

of the intact rock. Different correlations have been proposed based on different rock
29

types, test methods or procedures, and/or regression analysis to detect the best correlation

between the uniaxial compression strength (UCS) and indirect test results.

The main aim of this project is to develop correlation relations between the direct

compression strength and elastic properties with the indirect tests methods for basalt and

rhyolite rock. The core samples were received from MINE A in Nevada, USA.

3.2. Regression Analysis:

Regression analysis is a statistical technique for estimating the relationships among

variables. It includes many techniques for modeling and analyzing several variables,

when the focus is on the relationship between a dependent variable and one or more

independent variables. In general, regression analysis is a statistical tool for the

investigation of relationships between variables. Usually, the investigator seeks to

ascertain the causal effect of one variable upon another (Kenney & Keeping 1962).

3.2.1. Simple Linear Regression:

Simple linear regression fits a straight line through a set of n points in such a way that it

makes the sum of the squared residuals of the model (that is, vertical distances between

the points of the data set and the fitted line) as small as possible (Kenney & Keeping

1962):

(3.1)

3.2.2. Nonlinear Regression:

Nonlinear regression is a form of regression analysis in which observational data are

modeled by a function which is a nonlinear combination of the model parameters and


30

depends on one or more independent variables (Seber et al. 1989). All the non-linear

relations in this study are developed based on the regression power type:

(3.2)

To perform the regression analyses, test data was plotted in two dimensions as a scatter

plot. This format allows visualization/inspection of the data prior to running a regression

analysis. Different curve fitting relationships, such as linear, exponential, logarithmic,

polynomial, and power, can be used to analyze the relationship between a dependent and

independent variable. The curve fitting relationships produce a coefficient of

determination (R2). The coefficient of determination is the measure of the proportion of

variability on one variable that can be accounted for by variability on the other variable

(Sheskin, 2000).

3.2.3. Proposed Regression Models Performance:

To evaluate the performance of each regression equation, coefficient of correlation (R 2),

variance accounted for (VAF), logarithmic standard deviation (β), root mean square error

(RMSE), and mean absolute error (MAE) were calculated (Tzamos & Sofianos, 2006).

The amount of error around the line should be compared to the total amount of variability

in the distribution of the variable to be predicted to figure out whether it is small or large.

That is just the numerator in the equation for the variance. That value gives the total

amount of error (which will always be larger (or at worst equal to) the amount of error

around the regression line. To compare the two errors, the ratio of the amount of error

around the regression line is taken to the total amount of error in the distribution of the

variable to be predicted. Then, by convention, the value of this ratio is subtracted from 1.

We call that value the proportion of variance accounted for and we often refer to it as R-
31

squared. The logarithmic standard deviation (β) and root mean square error (RMSE) are

frequently used to measure the differences between values predicted by a model or an

estimator and the values actually observed. Mean absolute error (MAE) is a quantity used

to measure how close forecasts or predictions are to the eventual outcomes.

( ( )) (3.2)

√ ∑ [ ( )] (3.3)

∑ | | (3.4)

√ ∑ (3.5)

Where Aimeas is the ith measured element, Aipred is the ith predicted element and n is the

number of data set In the equation 3.2, var stands for variance.

3.3. Correlations between the Indirect and Direct Compression Strength

Test Methods

3.3.1. Ultimate compression strength (UCS) and Young’s modulus (E):

The relationship between ultimate strength and Young‟s modulus was calculated based

on the eight uniaxial compression tests. To evaluate the accuracy of the proposed

equation, Young‟s modulus of the specimen 2U and 7U was predicted and compared with

the actual value. It should be noted that the specimen 7U failed along a weakness plane.

The specimen 7U was from the same core as specimen 5U, However the ultimate

strength of specimen 7U is much less than the ultimate strength of specimen 5U.
32

Fig. 3.1. Specimen 7U (left) and 5U (right) after the failure

70

60
E = 2.17(UCS)0.57
Young’s Modulus (GPa)

50

40

30

20

10

0
0 50 100 150 200 250 300 350 400
Ultimate Strength (MPa)

Fig. 3.2. Correlation between ultimate strength and Young‟s modulus

Table 3.1. Performance of the ultimate strength-Young‟s modulus equation

R2 β VAF MAE RMSE


0.99 0.04 0.99 1.27 1.59

Sample ID UCS (MPa) Young’s modulus Young’s modulus β


(GPa) (GPa)
(actual) (predicted)
2U 213.89 59.47 46.20 0.25
7U 199.99 55.39 44.47 0.22
33

3.3.2. Ultimate compression strength (UCS) and Schmidt rebound hammer (SRH)

number:

The relationship between ultimate strength and SRH was calculated based on seven

uniaxial compression tests and seven averaged SRH numbers. The SRH value for

specimen 11U was not considered in this correlation, since a weakness plane affected the

SRH value. The averaged SRH numbers were calculated based on 20 readings and the

data was reduced based on ASTM D-5873. To evaluate the accuracy of the proposed

equation, Young‟s modulus of the specimens 2U and 7U were predicted and compared

with the actual values.

400

350
Ultimate Strength (MPa)

300 UCS = 0.25SRH1.77


250

200

150

100

50

0
0 10 20 30 40 50 60
SRH

Fig. 3.3. Correlation between ultimate strength and SRH

Table 3.2. Performance of the SRH- ultimate strength equation

R2 β VAF MAE RMSE


0.88 0.31 0.58 36.17 81.03

Sample ID SRH UCS (MPa) UCS (MPa) β


(actual) (predicted)
2U 50 213.89 254.17 0.17
7U 53 199.99 281.78 0.34
34

3.3.3. Ultimate compression strength (UCS) and splitting tensile strength (BRZ):

The relationship between ultimate strength and splitting tensile strength was calculated

based on eight uniaxial compression tests and eight splitting tensile strength values taken

from twenty two Brazilian tests. To evaluate the accuracy of the proposed equation,

ultimate strengths of the specimens 2U and 7U were predicted and compared with the

actual values.

400
350
Ultimate Strength (MPa)

300 UCS = 6.75(BRZ)1.08


250
200
150
100
50
0
0 5 10 15 20 25 30 35
BRZ (MPa)

Fig. 3.4. Correlation between ultimate strength and splitting tensile strength (BRZ)

Table 3.3. Performance of the splitting tensile strength-ultimate strength equation

R2 β VAF MAE RMSE


0.80 0.37 0.74 39.07 54.36

Sample ID BRZ UCS UCS (MPa) β


(MPa) (MPa) (predicted)
(actual)
2U 14.90 213.89 124.83 0.54
7U 22.19 199.99 191.93 0.04
35

3.3.4. Ultimate compression strength (UCS) and Ultrasonic Pulse velocity (P-wave):

The relationship between ultimate strength and ultrasonic pulse velocity was calculated

based on the eight uniaxial compression tests and eight ultrasonic velocity tests. To

evaluate the accuracy of the proposed equation, ultimate strengths of the specimens 2U

and 7U were predicted and compared with the actual values.

7.00

6.00

5.00
Ln(UCS (MPa))

4.00

3.00

2.00 ln(UCS) = 3.94*ln(P-wave) - 28.12


1.00

0.00
8.00 8.10 8.20 8.30 8.40 8.50 8.60 8.70
Ln(P-wave velocity (m/s))

Fig. 3.5. Correlation between ultimate strength and P-wave velocity

Table 3.4. Performance of the P-wave-ultimate strength equation

R2 β VAF MAE RMSE


0.92 0.23 0.84 32.37 42.96

Sample ID P-wave UCS UCS (MPa) β


velocity (m/s) (MPa) (predicted)
(actual)
2U 5220 213.89 272.43 0.24
7U 5089 199.99 246.47 0.21
36

3.3.5. Ultimate Compression Strength (UCS) and Point Load Index (Is(50)):

The relationship between ultimate strength and point load index was calculated based on

seven uniaxial compression tests and seven point load index values from ninety five point

load tests. To evaluate the accuracy of the proposed equation, ultimate strength of the

specimen 2U was predicted and compared with the actual value. There is no point load

result for samples 5U and 7U since the specimens were so strong that the point load

machine could not break them.

250

200 UCS = 90.14(Is(50)) 0.92


UCS (MPa)

150

100

50

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Is(50) (MPa)

Fig. 3.6. Correlation between ultimate strength and point load index

Table 3.5. Performance of the point load index-ultimate strength equation

R2 β VAF MAE RMSE


0.91 0.22 0.84 20.69 28.05

Sample ID Is (50) (MPa) UCS (MPa) UCS (MPa) β


(actual) (predicted)
2U 1.80 213.89 154.8 0.32
37

3.3.6. Young’s Modulus (E) and Point Load Index (Is(50)):

The relationship between Young‟s modulus and point load index was calculated based on

the seven uniaxial compression tests and seven point load index values from ninety five

point load tests. To evaluate the accuracy of the proposed equation, ultimate strength of

the specimen 2U was predicted and compared with the actual value.

4.0
3.8
ln(E) = 0.43(ln(Is(50)))1.06
3.6
3.4
Ln (E) GPa

3.2
3.0
2.8
2.6
2.4
2.2
2.0
5.0 5.5 6.0 6.5 7.0 7.5 8.0 8.5
LN (Is(50)) (kPa)

Fig. 3.7. Correlation between Young‟s modulus and point load index

Table 3.6. Performance of the point load index-Young‟s modulus equation

R2 β VAF MAE RMSE


0.86 0.17 0.79 4.32 5.64

Sample ID Is (50) (MPa) Young’s modulus Young’s modulus β


(GPa) (GPa)
(actual) (predicted)
2U 1.80 59.47 37.98 0.44

3.3.7. Young’s Modulus (E) and Splitting Tensile Strength (BRZ):

The relationship between Young‟s modulus and splitting tensile strength was calculated

based on the eight uniaxial compression tests and eight splitting tensile strength values
38

from twenty two Brazilian tests. To evaluate the accuracy of the proposed equation,

Young‟s modulus of the specimens 2U and 7U were predicted and compared with the

actual values.

70

60

50 E = 6.51(BRZ)0.61

40
E (GPa)

30

20

10

0
0 5 10 15 20 25 30 35
BRZ (MPa)

Fig. 3.8. Correlation between Young‟s modulus and splitting tensile strength

Table 3.7. Performance of the splitting tensile strength-Young‟s modulus equation

R2 β VAF MAE RMSE


0.78 0.22 0.79 5.70 6.86

Sample ID BRZ Young’s modulus Young’s modulus β


(MPa) (GPa) (GPa)
(actual) (predicted)
2U 14.90 59.47 33.82 0.56
7U 22.19 55.39 43.12 0.25

3.3.8. Splitting Tensile Strength (BRZ) and Point Load Index (Is(50)):

The relationship between splitting tensile strength and point load index was calculated

based on the seven splitting tensile strength number and seven point load index values.
39

To evaluate the accuracy of the proposed equation, ultimate strength of the specimens 2U

were predicted and compared with the actual values.

35

30
BRZ = 11(Is (50))
25
BRZ (MPa)

20

15

10

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Is (50) (MPa)

Fig. 3.9. Correlation between splitting tensile strength and point load index

Table 3.8. Performance of the splitting tensile strength-point load index equation

R2 β VAF MAE RMSE


0.81 0.37 0.81 3.67 4.19

Sample ID Is (50) (MPa) BRZ (MPa) BRZ (MPa) β


(actual) (predicted)
2U 1.80 14.9 19.8 0.28

3.3.9. Splitting Tensile Strength (BRZ) and Schmidt Rebound Hammer (SRH)

Number:

The relationship between splitting tensile strength and SRH was calculated based on the

seven splitting tensile strength tests and seven averaged SRH numbers. As mentioned, the

SRH value for specimen 11U was not considered in this correlation, since a weakness

plane affected the SRH value. To evaluate the accuracy of the proposed equation,
40

Young‟s modulus of the specimens 2U and 7U were predicted and compared with the

actual values.

35

30 BRZ = 0.15SRH1.33
25
BRZ (MPa)

20

15

10

0
0 10 20 30 40 50 60
SRH

Fig. 3.10. Correlation between splitting tensile strength and SRH

Table 3.9. Performance of the splitting tensile strength-SRH equation

R2 β VAF MAE RMSE


0.83 0.32 0.71 4.52 6.11

Sample ID SRH BRZ BRZ (MPa) β


(MPa) (predicted)
(actual)
2U 50 14.9 27.27 0.60
7U 53 22.19 29.46 0.28

3.3.10. Young’s Modulus (E) and Schmidt Rebound Hammer (SRH) Number:

The relationship between Young‟s modulus and SRH number was calculated based on

seven uniaxial compression tests and seven SRH numbers. To evaluate the accuracy of

the proposed equation, Young‟s moduli of the specimens 2U and 7U were predicted and

compared with the actual values.


41

70

60
E = 32.90ln(SRH) - 77.53
50

E (GPa)
40

30

20

10

0
0 10 20 30 40 50 60
SRH

Fig. 3.11. Correlation between splitting tensile strength and SRH

Table 3.10. Performance of the splitting tensile strength-SRH equation

R2 β VAF MAE RMSE


0.82 0.14 0.82 4.37 6.12

Sample ID SRH Young’s modulus (GPa) Young’s modulus (GPa) β


(actual) (predicted)
2U 50 59.47 51.17 0.15
7U 53 55.39 53.09 0.04

3.3.11. Young’s Modulus from Uniaxial Compression Test (EUCS) and Young’s

Modulus from Ultrasonic Velocity Test (EP-wave):

The relationship between Young‟s modulus from uniaxial compression test and Young‟s

modulus from ultrasonic velocity test was calculated based on the eight uniaxial

compression tests and eight ultrasonic velocity tests. To evaluate the accuracy of the

proposed equation, the results from the uniaxial compression strength for the specimens

2U and 7U were predicted and compared with the actual values.


42

4.5

4 ln (EUCS) = 1.3*ln(EP-wave) - 0.8

Ln (EUCS) (GPa)
3.5

2.5

2
2 2.5 3 3.5 4
Ln (EP-wave) (GPa)

Fig. 3.12. Correlation between EUCS and EP-wave

Table 3.11. Performance of EUCS-EP-wave equation

R2 β VAF MAE RMSE


0.91 0.14 0.87 4.33 5.45

Sample ID EP-wave EUCS EUCS (GPa) β


(GPa) (GPa) (predicted)
(actual)
2U 213.89 59.47 81.40 0.31
7U 199.99 55.39 66.68 0.18
43

The Effect of Cylindrical Core Specimen Length on the


Schmidt Rebound Hammer (SRH) Number

The Schmidt hammer is used for prediction of uniaxial compression strength and

determination of hardness of rock. Different sampling method, test procedure and

calculation are indicated in ASTM standard and ISRM suggested method for

determination of rock hardness by rebound hammer method. This study aims to

investigate the effect of cylindrical core sample length on the value of Schmidt rebound

hardness (SRH). In order to assess the effect of the length on the SRH value, HQ (63.5

mm diameter) core samples having sample lengths of 3, 5, 6, 8, 10, 12, 15, and 20 cm

from two different rock types have been prepared.

4.1. Introduction:

The Schmidt hammer is a commonly used device for hardness determination and for

prediction of the unconfined compressive strength of concrete and rock. The Schmidt

hammer test is a non-destructive, easy, inexpensive testing method which can be

performed both in laboratory and in the field. The Schmidt hammer test was originally

developed to test the surface rebound hardness of concrete (Schmidt, 1951). The Schmidt

hammer has been used worldwide as an index test for a quick rock strength and

deformability characterization due to its rapidity and easiness in execution, simplicity,

portability, low cost and non-destructiveness (Yilmaz & Sendir 2002). Haase (1962)

proposed a simple method of determining the strength of the surrounding rock and of
44

coal. Hucka (1965) presented Schmidt hammer test as a rapid method of determining the

strength of rock in-situ.

4.2. Principle of Schmidt hammer test:

The Schmidt hammer consists of a spring-loaded piston which is released when the

plunger is pressed against a surface. The impact of the piston onto the plunger transfers

the energy to the material. The extent to which this energy is recovered depends on the

hardness (or impact penetration/damage resistance) of the material, which is expressed as

a percentage of the maximum stretched length of the key spring before the release of the

piston to its length after the rebound (Aydin & Basu, 2005). The distance traveled by the

piston after rebound (expressed as a percentage of the initial extension of the key-spring)

is called the rebound value, which is considered to be an index of surface hardness (Aydin

& Basu, 2005). When the hammer is pressed orthogonally against a surface, the piston is

automatically released onto the plunger.

ASTM (ASTM D-5873, 2005) requires that specimens should be securely clamped to a

steel base and that cylindrical specimens should be placed along a machined slot (with an

arc-shaped cross-section of the same radius) or a V-block. It is essential to ensure that the

hammer axis is perpendicular to the test surface (with 5° of vertical with the bottom of

the piston at right angles to and in firm contact with the surface of the test specimen).
45

Fig. 4.1. Required V-shape steel base for Schmidt rebound hammer test

For tests conducted in situ on a rock mass, the rebound hammer can be used at any

desired orientation provided the plunger strikes perpendicular to the surface tested. The

results are corrected to a horizontal or vertical position using the correction curves

provided by the manufacturer (ASTM D-5873, 2005). It should be noted that in situ

testing may produce a wider scatter due to roughness of natural surfaces, lack of control

for the existence of cracks below the surface, and variations in moisture content (Aydin &

Basu, 2005).

4.3. Hammer type:

The standard L-and N-type Schmidt hammers are built to generate different levels of

impact energy: 0.735 and 2.207 Nm, respectively (Demirdag et al., 2009). ASTM

standard (ASTM D-5873, 2005) does not specify the hammer type; however, the earlier

ISRM suggested method (1978) suggested the use of only the L-type SH. ISRM (2009)

suggested for a given plunger tip diameter and radius of curvature, the impact energy of

the SH determines its range of applicability. The N-type hammer is less sensitive to

surface irregularities, and should be preferred in field applications; while the L-type
46

hammer has greater sensitivity in the lower range and gives better results when testing

weak, porous and weathered rocks (ISRM, 2009).

Potro & Hurlimann (2009) used L-type and N-type Schmidt hammers to estimate rock

strength in the thirty four different locations in the interested area. The rocks from 5

different geotechnical units were tested with both Schmidt hammers, with 20 readings

recorded for each material at each location. An error analysis for both Schmidt hammer

tests showed they both have very similar coefficient of variation distributions. In both

cases, the errors decrease linearly for stronger rocks. A correlation between both Schmidt

hammers was proposed according to available data (RN=1.0642RL+2.5687). The aim of

the study for Schmidt hammer was to determine whether it is valid to use the results from

this test to estimate the uniaxial compressive strength or not (both types). The results

indicated that for both hammers the errors are higher for weaker materials and the R L

values tend to be slightly more precise than the RN values for very weak (σci < 20 MPa)

and for relatively strong and very strong rocks (σci > 90 MPa). This study demonstrated

that RN provides very good linear correlation with unit weight; however, R L is less

sensitive to variations in unit weights.

4.4. Significance and use:

The rebound hardness method provides a means for rapid classification of the hardness of

rock during site characterization for engineering, design, and construction purposes,

geotechnical mapping of large underground openings in rock, or reporting the physical

description of rock core (ASTM D-5873, 2005). The methodology of the Schmidt hammer

test is expected to ensure reliable data acquisition and analysis on site or in the laboratory

(Basu & Aydin, 2006).


47

The hardness values from the laboratory and in-situ testing can be compared. Kahraman

et al. (2002) compared SRH values from core samples and in situ large block. In the

laboratory, core samples or unfractured rock blocks can be tested. In the field, Schmidt

hammer test can be performed on large rock blocks and on fractured rock mass. The

results from tests conducted on fractured rock mass will be different from the other test

results because of the fracturing. In this study, both in situ and laboratory Schmidt

hammer tests were performed on nine different rock types, five of which were igneous,

three of which were metamorphic and one of which was sedimentary. Tests were

performed with N-type hammer having impact energy of 2.207 Nm. Each test was

repeated at least three times on every rock type and the average value was recorded as

rebound number. For the laboratory testing, rock blocks were cored perpendicularly to

any visible bedding plane in the laboratory using an NX size (54 mm) diamond-coring

bit. In order to be able to describe the relationships between in situ and laboratory

Schmidt hammer rebound values of the tested rocks, regression analysis was made. The

authors mentioned that the derived regression equations can practically be used for the

prediction purposes with acceptable accuracy. Since all the testing in this research were

made with the hammer held vertically downward and at right angles to horizontal rock

faces, regression equations derived are valid for this case only (Kahraman et al., 2002).

4.4. Specimen requirements:

The Schmidt hammer is mainly used as index to predict the uniaxial compressive strength

and elastic modulus (Yilmaz & Sendir 2002). Most of the works related to SRH establish

relations between hardness and the other parameters of rocks. Day and Goudie (1977)

showed that the test points should be away from the boundaries to avoid abnormally low
48

values due to strong dissipation of impact energy. ISRM (1978) suggested that “block

specimens should have an edge length of at least 6 cm.” ISRM (2009) stated that length

of cores and surface area of blocks should be large enough to accommodate suggestions

related to dissipation of impact energy in the form of wave scatter or cracking. ASTM

(2005) recommended “drill core specimens shall be NX or larger core of at least 15 cm in

length. Block specimens shall have edge lengths of at least 15 cm.”

Demirdag et al. (2009) studied the optimum sample size based on laboratory and field

investigations. In this study, different approaches have been applied for the Schmidt

Hammer test procedure: ISRM, Hucka, Poole & Farmer, and Fowell & Smith methods.

In order to analyze the effect of sample size on the SRH values, eight different rock types

were studied. Cubic samples with edge dimensions of 6,7,8,10,12, and 15 cm were

sampled from each rock type. By analyzing the Schmidt hardness values of rocks

measured in the laboratory, the optimum edge dimension of cubic samples is found to be

11 cm, based on measurements performed by different methods of Schmidt hardness

tests. Also, in-situ SRH measurements showed that in-situ SRH value is equal to SRH

values obtained from samples with edge dimension larger than 11 cm.
49

Fig. 4.2. Schmidt hammer test requirements based on ASTM standard: hammer axis is
perpendicular to the surface test surface.

The effect of sample size on Schmidt rebound hardness (SRH) value of rocks is

investigated. Core samples having sample lengths of 3, 5, 6, 8, 10, 12, 15, and 20 cm

from three different rocks have been prepared. In this study, the Schmidt hammer tests

are performed based on both ASTM standard and ISRM suggested method using L-type

Schmidt hammer with impact energy of 0.735 Nm (Aydin & Basu, 2005). The

recommended test procedures used in the study are:

ASTM standard given number (2005): 10 reading at different locations separated by at

least the diameter of the piston and only one test may be taken at any one point. Discard

readings differing from the average of ten readings by more than seven units and

determine the average of the remaining readings.

ISRM suggested method (1978): at least 20 rebound values from single impacts

separated by at least a plunger diameter, averaging the upper 50% values and eliminating

any taken from rock that shows signs of cracking.


50

ASTM Standard
70
60
50
40
SRH
SH-1
30
SH-2
20
SH-3
10
0
0 2 4 6 8 10 12 14 16 18 20 22
Length (cm)

Fig. 4.3. Core sample length against SRH value based on ASTM standard procedure

ISRM Suggested Method


70
60
50
40
SRH

SH-1
30
SH-2
20
SH-3
10
0
0 2 4 6 8 10 12 14 16 18 20 22
Length (cm)

Fig. 4.4. Core sample length against SRH value based on ISRM suggested method procedure

The curves demonstrate that the SRH value initially decreases significantly for the length

of 3 cm to 5 cm. However, the SRH values increase with increasing core sample length

from 5 cm to 12 cm and reach a constant value for the core samples with a length greater

than 12 cm.
51

Fig. 4.5. Cylindrical core samples with lengths of 20 cm to 3 cm

Table 4.1. Critical core sample length according to ASTM standard and ISRM suggested method
procedures

Rock type ASTM (cm) ISRM (cm)

SH-1 12 12.2
SH-2 11.5 11
SH-3 13 13

60

50
SRH

40

30
0 2 4 6 8 10 12 14 16 18 20 22
Sample Length (cm) (SH-2 sample)

Fig. 4.6. The Effect of Cylindrical Core Sample Length on Schmidt Hammer Hardness Value
(Method for critical core sample length determination)
52

Numerical Simulation of Block Punch Index (BPI) Test

5.1. Block Punch Index Test:

The UCS is an important input parameter in rock mass classifications. Often, the drilling

process breaks up the weaker core pieces, and they are too thin or fragmented to be used

in uniaxial compression strength test. Another undesirable aspect of the UCS test is the

amount of time and labor necessary for sample preparation (Sulukcu & Ulusay 2001).

Therefore, a test procedure that would use small segments of the core with minimal

sample preparation to determine directly or indirectly the rock strength has always been

attractive. The block punch index (BPI) test, which requires flat disc specimens without

special treatment, has been developed during the last decade.

Fig. 5.1. Block Punch Index Machine (Modified GCTS PLI-100 Model)
53

Four of the earliest references to a punch test were in the articles by Lacharite, Mazanti

and Sowers, Vutukuri et al., and Stacey where the test was used for determining the direct

shear strength of rock specimens using a simple apparatus (Ulusay et al. 2001). These

studies have been extended in The Netherlands by Taselaar (Schrier 1982) and Schrier,

and have led to the development of the BPI test. Schrier obtained high correlations

between UCS, Brazilian tensile strength and BPI values from a limited number of

specimens of sedimentary and metamorphic rocks. But none of these investigators

considered the size effect of the test specimens on BPI and the general usefulness of this

test.

Between 1997 and 1999, Ulusay and Gokceoglu used the BPI test extensively to assess

the size effect, strength anisotropy and its possible use in rock engineering. They

suggested the corrected BPI value by using the size-correction factors, and the strength

anisotropy transformation factor to estimate the strength index in the strongest direction

in conjunction with possible uses of the BPI in rock engineering applications. However,

in these latest works on the BPI test, it was recommended that application of the testing

method should be extended to other rock types; the stress distribution within the rock

discs under BPI test should be considered by means of numerical methods in future

studies; and the rock strength predicted from different index tests including the BPI test

should be experimentally compared to make recommendations regarding the accuracy

and applicability of the results of the BPI test (Sulukcu & Ulusay 2001).

The flat disk shaped specimen is fastened symmetrically in a clamp and the band of rock

between the supports is vertically loaded by a rectangular rigid punch block. Fracturing is

forced to take place along two parallel planes on which the normal stress is considered to
54

be zero while tensile stresses caused by bending are reduced to a minimum. After failure,

the specimen is broken into three parts, the two ends which are fixed in the apparatus and

the band which is punched out. In order to provide a more representative comparison of

the results from different strength and index test, BPI disc specimens cut from the top and

bottom of the core were prepared for UCS, Brazilian, and PLI tests (Ulusay et al. 2001).

Fig. 5.2. Prepared specimens for BPI test

In 2001, Ulusay et al. proposed the draft ISRM suggested method for BPI determination.

In the same year, Sulukcu and Ulusay (2001) presented a study of the BPI test device to

provide new contributions to previous works on size effect in BPI tests using a wide

range of rock types, and also to assess the effectiveness of the test in predicting rock

strength by an experimental way.

The study by Karakul et al. (2011) aimed to investigate the strength anisotropy associated

with discontinuity orientation by performing block punch index (BPI) and uniaxial

compressive strength tests, and to develop some empirical equations for estimating BPI

and UCS in the strongest direction. In the study, for the assessment of the strength
55

anisotropy, a total of 1568 specimens for both UCS and BPI tests were extracted from

rock blocks in varying directions (0o, 30o, 45o, 60o, and 90o) relative to the weakness

planes, using metal wedges manufactured with different surface inclinations. The strength

anisotropy transformation factor proposed in this study is Kσα= σc90/ σcα, where σc90 is the

UCS value at α=90o and σcα is UCS determined at any angle to weakness plane. In terms

of BPI, the following equation was proposed for predicting the BPI value in the strongest

direction (BPIc90): BPIc90=1.47e-0.00496αBPIcα. In order to indirectly estimate the UCS from

BPI, the transformation equation, which was recommended by Ulusay (2001) (σc90=5.1

BPIc90), was modified for assessment of strength anisotropy that can estimate the UCS

from BPI of the specimen at any angle between the loading direction and the weakness

plane: σcα=5.1(1.47e-0.00496αBPIcα)/[1.7exp[(αt-24.7)2/(-1530)]+0.01αt], where αt is the

angle between the loading direction and the weakness plane for the UCS. The finite

difference simulation showed that the numerically estimated BPI values are very close to

experimentally determined values for each angle between the loading direction and

weakness plane involved by the BPI samples.

Mishra and Basu (2012) developed a correlation between BPI and UCS for three rock

types (granite, schist, and sandstone). In the experimental study, correlations of BPI and

point load index (PLI) with UCS were proposed for all three types of rock. The results

demonstrated that: (1) for granite, the UCS is in a better correlation with BPI than PLI;

(2) for schist and sandstone, PLI provides better correlation with UCS rather than PBI.

Also, a new empirical equation was developed by the authors between UCS and BPI

(UCS=4.93BPIc, R2=0.87) and was compared with other equations. The comparison
56

shows the validity of this empirical equation. The same results were given for

correlations of BTS with BPI and PLI.

5.2. Block Punch Index Limitations:

The block punch index (BPI) test is intended as an index test for the strength

classification of rock materials and can be correlated with the UCS (Ulusay et al.

2001). This method is not widely used, and according to available references almost all

the studies have been limited to few authors. There are some shortcomings associated

with this unconventional test; this test can be conducted on very thin specimens only.

Irregular failure (invalid test result) occurrence causes the test to require too many rock

specimens (Yalmiz 2009). Another problem is that there is no ASTM standard available

for this test; also, the ISRM suggested method for BPI is limited to one author and

corrections may not be applicable to various types of rock.

Mishra and Basu (2012) studied the failure patterns in the BPI test when the compression

force is applied to the specimen. When the force gradually increases, the middle part of

the specimen is punched out by the induced double shear failure. In the BPI test, the test

is invalid if parallel vertical fracture planes are either absent or not fully developed.

However, weakness planes are characteristic of anisotropic rocks. The authors believed

that in case of anisotropic rock, the mechanical behavior will be controlled by the planes

under a certain state of stress. So, according to the mechanical behavior of anisotropic

rock, the common shear failure in these rocks may be taken into account as valid tests.
57

5.3. Numerical Simulation of Block Punch Index Test:

In this study, the block punch index (BPI) test was considered as one of the index test

methods for the intact rock strength determination. However, approximately 95% of the

block punch index tests that were performed should be considered as invalid tests. Fig.

5.3 shows the invalid and valid block punch index as it is suggested by ISRM (Ulusay et

al. 2001).

According to Ulusay et al. (2001), the specimen should be broken into three parts: the

two ends which are fixed in the apparatus and the band which is punched out. However,

in many cases the middle part (which is punched out) is broken in many parts. This

happened in many of the BPI tests which were performed on the prepared specimens with

the uniaxial compression strength between 39 MPa to more than 170 MPa.

Fig. 5.3. Valid (left) and invalid (right) block punch index tests
58

Fig. 5.4. Many of the block punch index tests were invalid (according to ISRM suggested
method) since the middle part of the specimen which is under compression load was broken in
many parts.

The problem associated with the middle part of the specimen led to study of the stress

distribution in the block punch index test. The HQ (63.39 mm) core specimen with 10

mm thickness was modeled based on the information from the uniaxial compression

strength and splitting tensile strength tests of specimens 5U, 6U, and W-5U (table 5.1) in

the ANSYS 14. Software.


59

Table 5.1. Isotropic properties of the specimens 5U, 6U, and W-5U for numerical simulation of
BPI test

Sample ID Ultimate compression Ultimate tensile Young’s Poisson’s ratio


strength (MPa) strength (MPa) modulus (GPa)
5U 375.62 29.35 62.23 0.24
6U 170.68 16.92 42.04 0.25
W-5U 39.08 9.37 17.4 0.18

For the block punch index test modeling, top and bottom surfaces of the core sample

were split in three parts to model the fixed supports and loading plate of the machine as

shown in Fig. 5.5. The main reason for splitting the surface, not the body of the sketch, is

to build a model close to the real block punch index test.

Fig. 5.5. Splitting the top and bottom surfaces of the body sketch to model the supports and

loading plate

Since the UNR-ECC Lab has the student license for ANSYS 14., the maximum mesh

nodes that can be utilized is 30000. An attempt was made to refine the mesh at the most

important areas, which are the four splitting lines on the specimen surfaces with a

minimum element size of 1.9 mm. After the refinement, the number of nodes and

elements are equal to 27919 and17375, respectively.


60

Fig. 5.6. Mesh refinement at the two parallel planes at the top and bottom of the specimen

5.3.1. Compression only support at the bottom of the specimen:

Compression only support was applied at the bottom surface of the body. The

compression support was used to fix the body in compression (loading) direction and free

the body in other directions. Maximum principal stress and safety factor- Mohr-Coulomb

were calculated based on the available data from the experiment. The load was increased

to reach the safety factor equal to 1 for Mohr-Coulomb. The results of the modeling for

three core specimen were the same. In this chapter, stress distribution and safety factors

are presented for specimen W-5U.

A stress of 2.5 MPa was applied to the top-middle part of the body. The results show that

the specimen is broken under this load with the safety factor less than one. There are

shear and normal stresses, distributed in the middle part of the body. Particularly when

the specimen is relatively strong, the stress at the middle of the specimen is high, in

respect to stress at two parallel planes. For the specimen W-5U, the stress is distributed

more uniformly in comparison with the other two specimens. As was indicated in the

Table 5.1, specimen W-5U is moderately weak rock.


61

Fig. 5.7. Mohr-Coulomb safety factor at the bottom (left) and top (right) for specimen W-5U at
the 2.5 MPa Pressure with compression support at the bottom of the body

Fig. 5.8. Maximum principal stresses (MPa) at bottom (left) and top (right) for the specimen W-
5U with compression support at the bottom of the body

5.3.2. Fixed support at bottom and top of the specimen:

Fixed support was applied at the bottom and top surfaces of the sample. The fixed

support with the face splitting at the bottom and top of the specimen is the most realistic

model in comparison with experimental block punch index test. The assumption is that

the clamping bars are fixed and there is no bending at the bottom of the specimen.

Maximum principal stress and safety factor- Mohr-Coulomb were calculated based on the

available data from the experiment (Table 5.1). The load was increased to reach the

safety factor of one for either shear or Mohr-Coulomb.


62

Fig. 5.9. Mohr-Coulomb safety factor at the bottom (left) and top (right) for specimen W-5U at
the 2.5 MPa applied stress with fixed supports

Fig. 5.10. Maximum principal stresses (MPa) at bottom (left) and top (right) for the specimen W-
5U with fixed supports

Both models indicate that the Mohr-Coulomb safety factor is less than 1 in 2.5 MPa

pressure. In general, the highest and lowest values for safety factors and stress

distribution are caused by singularity issue. Without considering these extreme values, it

is obvious that the maximum stresses are at the two parallel planes. However, the realistic

model could be a model between these two scenarios. Although the clamping bars can be

adjusted for different diameters, for stronger rock, a very small place is needed for

bending to occur at the bottom-middle part of the specimen. Practically, the load is

increasing using a hydraulic hand pump on a brittle rock. It is going to take a minute to

few minutes until the failure occurs. That means the specimen is under an increasing

compression force and there are increasing forces against the clamping bars at the two
63

part of the specimen, which are fixed on the apparatus. Two extreme cases would be two

scenarios which were defined in this study. Considering the stress distributions in the BPI

specimen, should not the tests which are shown in Fig. 5.1 be considered as valid tests

and not as invalid tests? As far as the two complete shear failures occur at the two

parallel planes, the specimen is broken into three main parts, two ends which are fixed,

and the middle part which is punch out. Secondary failure may occur owing to the fact

that there is tensile stress at the middle part of the specimen caused by bending. In the

splitting tensile strength test, secondary failures occur in the specimen; and the test is not

considered as invalid test.


64

Conclusions

Uniaxial compressive strength of intact rock is important for engineering geology and

geotechnics, because it is major design parameter for mines, tunnels, slopes, and rock

foundations, and it is used as input parameter in many rock mass classification systems.

The difficulties associated with performing direct compression strength tests on rock

leads to development of indirect test methods for the rock strength assessment. Indirect

test methods are widely used because they are simple, more economical, less time

consuming, and easily adaptable to the field. The main aim of this study was to define

correlations between the results of direct and indirect test methods for the core sample

available from Carlin Trend in NV. Eleven correlations between the direct and indirect

compression strength test results were developed, using linear and nonlinear regression

analyses. To evaluate the performance of each regression equation, coefficient of

correlation (R2), variance accounted for (VAF), root mean square error (RMSE) and

mean absolute error (MAE) were calculated. The results show that the uniaxial

compression strength has the best correlation with the splitting tensile strength.

Furthermore, the Poisson‟s ratio has no correlation with any of the direct and indirect test

results.

The validity of the proposed equations is limited by the data range and sample types

which were used to derive the equations. Future work can be focused on increasing the

data of direct and indirect tests for the same types of rock. It is important to consider

specimens which come with RQD and core logging information.

Ten UCS specimens were prepared for the uniaxial compression strength test. The

specimens were divided in three groups: (1) with four strain gages, (2) with two strain
65

gages, and (3) with no strain gage. The results of UCS tests on the prepared specimens

with four bonded rosettes indicate that the size of the rosettes is acceptable to apply to

HQ (63.39 mm) core specimen.

Comparison between the errors of three methods of Young‟s modulus calculation for five

different UCS samples indicate that the “Average Modulus of Linear Portion of Axial

Stress-Strain Curve” is the best method for Young‟s modulus and Poisson‟s ratio

calculations. A correction factor for the deformation recorded with a MTS test machine

is calculated based on the slope of linear curve of all the considered displacements, which

is equal to “J = 2.25”. The core specimens with no rosettes were used for evaluating the

correlation relations performances.

The aim of the study of the cylindrical core specimen length was to assess the effect of

the core sample length, as a dimension factor of rock, on the Schmidt rebound hammer

(SRH) value. The evaluation of the calculated data showed that the SRH value changes

considerably when the length of the specimen is less than 12 cm. For the HQ core

samples, the effect of the specimen length on the SRH values is not significant for the

specimens having a length of more than 12 cm. However, in many cases, a core sample

with length greater than 12 cm is not available. Further studies are needed to investigate

the effect of discontinuities on the SRH values. A reasonable data base is needed to study

to evaluate if the SRH-Length plot acts the same for the length less than 12 cm for

different rock types. If the behavior of the SRH-Length is the same for a variety of rock

types, the SRH15cm can be predicted based on smaller core samples. The curves

demonstrate that the SRH value initially decreases significantly for the lengths 3 cm to 5

cm.
66

The block punch index (BPI) test, which requires flat disc specimens without special

treatment, has been developed during the last decade that can be used as an index test for

the strength classification of rock materials and correlated with the other direct and

indirect test methods. According to the ISRM suggested method for block punch index

test the test should be rejected as invalid if the parallel fracture planes are either absent or

not fully developed, or cross joints develop. The observations indicate that the middle

part of the specimen may be broken into many parts. The numerical simulation

demonstrates that there tensile stress at the bottom-middle part of the specimen,

particularly when the clamping bars cannot hold the specimen fixed (and there is place

for bending). Secondary failures or developed cross joints can be considered as secondary

failures as occur in splitting tensile strength or point load index test. For stronger

specimen secondary failures are likely to occur. As far as the two complete shear failures

occur at the two parallel planes, the specimen is broken into three main parts, two ends

which are fixed, and the middle part which is punch out. Secondary failure may occur

owing to the fact that there is tensile stress at the middle part of the specimen caused by

bending.

One of the interesting subjects on the splitting tensile strength is to study stress

distribution on the standard and flattened specimens. Observations from the experiment

show that the tensile failure does not frequently occur along the loaded diameter. In many

cases, the fractures do not go through the center and separate the disc in two halves.

Flattened specimens may solve this issue, however it would turn the splitting test to a

difficult and time consuming test. Future work can be focused on calculation of tensile

elastic modulus from the splitting test.


67

Recommendations

Indirect test methods are widely used because they are simple, more economical, less

time consuming, and easily adaptable to the field. However, there are a few

considerations for application of indirect test methods for compression strength

determination. The number of tests and the uncertainty associated with validity of the

tests are the most important considerations in the application of the indirect tests.

For an HQ (63.39 mm diameter) specimen, the approximate time for cutting the core

specimen is 10 to 15 minutes. The procedure of cutting the specimens is very time

consuming, especially for the indirect test methods, when multiple tests are needed (such

as block punch index). It is recommended to replace or modify the saw in the UNR-Rock

Mechanics Lab.

Performing a splitting tensile strength test with the MTS machine is time consuming.

First of all, it is not easy to run a test with the MTS machine. Secondly, with the hand

loading device, more tests can be performed on the specimens. It is recommended to

study the importance of the uniform load on the splitting test specimen, as is

recommended by ASTM D-3967. Many of the previous works on the splitting tests were

done using a hand pump loading device. It is also recommended to place a simple hand

pump loading device for performing the splitting tensile strength test.
68

References

Aksoy C.O., Ozacar V., Demirel N., Ozer S.C., and Safak S., (2011), “Determination of instantaneous
breaking rate by geological strength index, block punch index, and power of impact hammer for various
rock mass conditions”, Journal of Tunneling and Underground Space Technology, Vol. 26, pp. 534-540.

Andreev G.E., (1991) “A Review of Brazilian Test for Tensile Strength Determination. Part 1: Calculation
Formula,” Mining Science and Technology, Vol. 13, pp. 445-456.

ASTM, (2005) “Standard Test Method for Determination of Rock Hardness by Rebound Hammer
Method,” ASTM Stand. D5873 – 00(2005).

ASTM (2008) “Standard Method for Determination of The Point Load Strength Index of Rock”, ASTM
Stand. D5731–08(2008).

ASTM, (2010) “Standard Method for Compressive Strength and Elastic Moduli of Intact Rock Core
Specimens under Varying States of Stress and Temperatures,” ASTM Stand. D5873 – 10(2010).

ASTM, (2008) “Standard Test Method for Splitting Tensile Strength of Intact Rock Core Specimens,”
ASTM Stand. D3967 – 08(2008).

ASTM, (2008) “Standard Test Method for Slake Durability of Shales and Similar Weak Rocks,” ASTM
Stand. D4644– 08(2008).

ASTM, (2010) “Standard Test Method for Laboratory Determination of Water (Moisture) Content of Soil
and Rock by Mass,” ASTM Stand. D2216– 10(2010).

ASTM, (2008) “Standard Test Method for Laboratory Determination of Pulse Velocities and Ultrasonic
Elastic Constants of Rock,” ASTM Stand. D2845– 08(2008).

ASTM, (2008) “Standard Test Method for Preparing Rock Core as Cylindrical Test Specimens and
Verifying Conformance to Dimensional and Shape Tolerances” ASTM Stand. D4543– 08(2008).

Aydin A., (2009) “ISRM Suggested Method for Determination of the Schmidt Hammer Rebound Hardness:
Revised Version,” International journal of rock mechanics and mining sciences, Vol. 46, pp. 627-634.

Aydin A., Basu A., (2005) “The Schmidt Hammer in Rock Material Characterization,” Engineering
Geology, Vol. 81, pp. 1-14.

Aydin A., (2009) “ISRM Suggested Method for Determination of the Schmidt Hammer Rebound Hardness:
Revised Version,” International Journal of Rock Mechanics and Mining Sciences, Vol. 46, pp. 627-634.

Barton, N., & Choubey, V. (1977). “The shear strength of rock joints in theory and practice,” Rock
mechanics Vol. 10, pp.1-54.

Basu A. and Kamran M., (2010) “Point Load Test on Schistose Rocks and its Applicability in Predicting
Uniaxial Compressive Strength,” International Journal of Rock Mechanics and Mining Sciences, Vol. 47,
pp. 823–828.

Basu A. and Aydin A., (2006) “Predicting Uniaxial Compressive Strength by Point Load Test: Significance
of Cone Penetration,” Rock Mechanics and Rock Engineering, Vol. 39, pp. 483–490.

Bieniawski Z. T., (1975)“The point-load test in geotechnical practice,” Engineering Geology, vol. 9, pp. 1–
11.

Broch E. and Franklin J. A., (1972) “The point-load strength test,” International Journal of Rock
Mechanics and Mining Sciences & Geomechanics Abstracts, Vol. 9, pp. 669–676.
69

Cobanoglu I. & Celik S.B., (2008), “Estimation of Uniaxial Compressive Strength from Point Load
Strength, Schmidt Hardness and P-wave Velocity,” Journal of Bulletin of Engineering Geology and the
Environment, Vol. 67, pp. 491-498.

Day MJ, Goudie AS, (1977) “Field Assessment of Rock Hardness using the Schmidt Hammer Test,” Br
Geomorph Res Group Tech Bull, Vol. 18, pp. 19-29.

Dan D., Konietzky H., Herbst M., (2013) “Brazilian tensile strength tests on some anisotropic rocks,”
International Journal of Rock Mechanics and Mining Science, Vol. 58, pp. 1-7.

Deere, D. U., & Miller, R. P. (1966). Engineering classification and index properties for intact rock.
ILLINOIS UNIV AT URBANA DEPT OF CIVIL ENGINEERING.

Demirdag S., Yavuz H., and Altindag R., (2009) “The Effect of Sample Size on Schmidt Rebound
Hardness Value of Rocks,” International journal of rock mechanics and mining science, Vol. 46, pp. 725-
730.

Fairhurst, C., (1964) “On the validity of the „Brazilian‟test for brittle materials,” In International Journal of
Rock Mechanics and Mining Sciences & Geomechanics Abstracts Vol. 1, pp. 535-546.

Franklin J. A. & Chandra R., (1972) “The Slake Durability Test,” International Journal of Rock Mechanics
and Mining Science, Vol. 9, pp. 325-341.

Fener, M., Kahraman, S., Bilgil, A., & Gunaydin, O. (2005). A comparative evaluation of indirect methods
to estimate the compressive strength of rocks,‟ Rock Mechanics and Rock Engineering, Vol. 38, pp. 329-
343.

Goodman, R. E. (1989). “Introduction to rock mechanics”. New York: Wiley, p. 562.

Hack. R & Huisman. M., (2002), “Estimating the Intact Rock Strength of a Rock Mass by Simple Means,”
Proceedings of 9th Congress of the International Association for Engineering Geology and the
Environment. Durban, South Africa, 16 - 20 September 2002 - J. L. van Rooy and C. A. Jermy, editors.

Heidari M., Khanlari G. R., Kaveh M., and Kargarian S., (2012) “Predicting the Uniaxial Compressive and
Tensile Strengths of Gypsum Rock by Point Load Testing,” Rock Mechanics and Rock Engineering, Vol.
45, pp. 265–273.

Hoek E., (1977) “Rock Mechanics Laboratory Testing in the Context of a Consulting Engineering
Organization,” International Journal of Rock Mechanics and Mining Sciences & Geomechanics Abstracts,
Vol. 14, pp. 93–101.

Hucka V., (1965) “A Rapid Method of Determining the Strength of Rock In-Situ,” International Journal of
Rock Mechanics and Mining Sciences, Vol. 2, pp. 127-134.

ISRM (International Society of Rock Mechanics), (1973) “Suggested Methods for Determining the Point
Load Strength Index”, ISRM Commission on Laboratory Tests, pp. 8–12.

Kahraman S., Fener M., Gunaydin O., (2002), “Predicting the Schmidt Hammer Values of In–Situ Intact
Rock from Core Sample Values,” International Journal of Rock Mechanics and Mining Sciences, Vol. 39,
pp. 385-399.

Kahraman S., Gunaydin O., and Fener M., (2005), “The Effect of Porosity on the Relation between
Uniaxial Compressive Strength and Point Load Index,” International Journal of Rock Mechanics and
Mining Sciences, Vol. 42, pp. 584-589.

Karakul H., Ulusay R., and Isik N.S., (2011), “Empirical Models and Numerical Analysis for Assessing
Strength Anisotropy based on Block Punch Index and Uniaxial Compression Strength Tests,” International
Journal of Rock Mechanics and Mining Sciences, Vol. 47, pp. 657-665.
70

Kenney, J. F. and Keeping, E. S. (1962) "Linear Regression and Correlation." Ch. 15 in Mathematics of
Statistics, Pt. 1, 3rd ed. Princeton, NJ: Van Nostrand, pp. 252-285.

Miller, R.P., (1965) “Engineering Classification and Index Properties for Intact Rock,” Ph.D. Thesis,
University of Illinois.

Mishra D. A. and Basu A., (2012) “Use of the Block Punch Test to Predict the Compressive and Tensile
Strengths of Rocks,” International Journal of Rock Mechanics and Mining Sciences, Vol. 51, pp. 119–127.

Monjezi M., Rezaei M., Yazdian A., (2010), “Prediction of Back-Break in Open-Pit Blasting using Fuzzy
Set Theory”, Expert Systems with Applications Journal, vol. 37, pp. 2637-2643.

Quac Dan D., Konietzky H., Herbst M., (2013), “Brazilian Tensile Strength Tests on Some Anisotropic
Rocks,” International Journal of Rock Mechanics and Mining Sciences, Vol. 57, pp. 1-7.

Potro R. and Hurlimann M., (2009), “A Comparison of Different Indirect Techniques to Evaluate Volcanic
Intact Rock Strength”, Rock Mechanics and Rock Engineering, Vol. 42, pp. 931-938.

Sabatakakis N., Koukis G., Tsiambaos G., Papanakli S., (2008) “Index Properties and Strength Variation
Controlled by Microstructure for Sedimentary Rocks,” Engineering Geology , Vol. 97, pp. 80-90.

Schrier, J. S. (1988) “The block punch index test,” Bulletin of the International Association of Engineering
Geology-Bulletin de l'Association Internationale de Géologie de l'Ingénieur, Vol. 38, pp. 121-126.

Seber G.A.F and Wild C.J.,(1989) ”Nonlinear Regression”. New York: John Wiley and Sons.

Sharma P. K., Khandelwal M., Singh T. N., (2011), “A Correlation between Schmidt Hammer Rebound
Numbers with Impact Strength Index, Slake Durability Index and P-wave Velocity,” International Journal
of Earth Sciences, Vol. 100, pp. 189-195.

Sharma P.K. & Singh T.N., (2008), “A Correlation between P-wave Velocity, Impact Strength Index, Slake
Durability Index, and Uniaxial Compressive Strength,” Journal Bulletin of Engineering Geology and the
Environment, Vol. 67, pp.17-22.

Sheskin, D.J. (2000). "Handbook of parametric and nonparametric statistical procedures, 2nd edition",
Chapman & Hall, New York.

Smith H. J., (1997) “The Point Load Test for Weak Rock in Dredging Applications,” International Journal
of Rock Mechanics and Mining Sciences, vol. 34, pp. 295.e1–295.e13.

Sulukcu S. and Ulusay R., (2001) “Evaluation of the Block Punch Index Test with Particular Reference to
the Size Effect, Failure Mechanism and its Effectiveness in Predicting Rock Strength,” International
Journal of Rock Mechanics and Mining Sciences, Vol. 38, pp. 1091–1111.

Tzamos S. and Sofianos, A. I., (2006), “Extending the Q System‟s Prediction of support in tunnel
employing Fuzzy logic and extra parameters”, International Journal of Rock Mechanics and Mining
Sciences, Vol. 43, pp. 938-949.

Ulusay, R., Gokceoglu, C., & Sulukcu, S., (2001) “Draft ISRM suggested method for determining block
punch strength index (BPI),” International Journal of Rock Mechanics and Mining Sciences, Vol. 38, pp.
1113-1120.

Wijk G., (1978) “Some New Theoretical Sspects of Indirect Measurements of the Tensile Strength of
Rocks,” International Journal of Rock Mechanics and Mining Sciences & Geomechanics Abstracts, Vol.
15, pp. 149–160.

Yilmaz I., (2009) “A new testing method for indirect determination of the unconfined compressive strength
of rocks,” International Journal of Rock Mechanics and Mining Sciences, Vol. 46, pp. 1349-1357.
71

Yilmaz I, Sendir H, (2002) “Correlation of Schmidt Hardness with Unconfined Compressive Strength and
Young‟s Modulus in Gypsum from Sivas (Turkey),” Engineering Geology, Vol. 97, pp. 80-90.

Yu Y., Zhang J., Zhang J., (2009), “A Modified Brazilian Disc Tension Test”, International Journal of
Rock Mechanics and Mining Sciences, Vol. 46, pp. 421-425.
72

Appendix-1: Specimen Preparation Report

“Rock is a complex engineering material that can vary greatly as a function of lithology,

stress history, weathering, moisture content and chemistry, and other natural geologic

processes. As such, it is not always possible to obtain or prepare rock core specimens

that satisfy the desirable tolerances given in this practice. Most commonly, this situation

presents itself with weaker, more porous, and poorly cemented rock types and rock types

containing significant or weak (or both) structural features. For these and other rock

types which are difficult to prepare, all reasonable efforts shall be made to prepare a

specimen in accordance with this practice and for the intended test procedure. ASTM D-

4543 specifies procedures for laboratory rock core test specimen preparation of rock

core from drill core and block samples for strength and deformation testing and for

determining the conformance of the test specimen dimensions with tolerances established

by this practice (2010).” There are preparation considerations in the ASTM D-4543

about the unit and the specimen requirements for “compressive strength test” which are

comprehensively described in ASTM D-4543:

1. The dimensional values stated in either inch-pound units or SI units are to be

regarded as standard.

2. Test specimens shall be right circular cylinders within the tolerances specified

herein.

3. The specimen shall have a length-to-diameter ratio (L/D) of 2.0 to 2.5 and a

diameter of not less than 1-7⁄8 in. (47 mm).

4. The larger the internal friction angle of a specimen the more desirable it will be to

have larger L/D ratios so that the specimen can potentially develop a true shear
73

plane that does not pass through either end of the specimen or is not altered by the

specimen size.

5. The cylindrical surfaces of the specimen shall be generally smooth and free of

abrupt irregularities, with all the elements straight to within 0.020 in. (0.50 mm)

over the full length of the specimen.

6. The ends of the specimen shall be cut parallel to each other and at right angles to

the longitudinal axis. The end surfaces shall be surface ground or lapped flat to a

tolerance not to exceed 0.001 in. (25 μm).

7. The ends of the specimen shall not depart from perpendicularity to the axis of the

specimen by more than 0.25o.

8. The parallelism tolerance is the maximum angular difference between the

opposing best-fit straight line on each specimen end.

All the specimens for the uniaxial compression strength tests were prepared based on the

ASTM standard “ASTM D-4543” and “ASTM D-7012”. All the specimens met the

ASTM D-4543 standard requirement, except sample W-2U which did not meet the side

smoothness requirement. For each of the samples, all the information about the specimen

(including diameter, length, weight, unit weight, density, moisture content, and slake

durability index) have been shown in the preparation form. The preparation form also

includes specimen dimension checks.

In the specimen preparation form, the diameter and length calculation were not included.

The diameter was calculated as an “average of two approximately perpendicular

diameters, at top, middle, and bottom of the specimen.” The length was calculated as an

“average of the length of the specimen along three lines at 120o from each other.”
74

Sample 2U

Specimen ID 2U
Source location Mine “A”
Rock type Basalt
Equipment used for sample preparation Drill, Saw, Grinder
Storage condition Room
Moisture condition Oven dried
Specimen length (inch) 4.74
Specimen Diameter (inch) 2.495
Specimen Weight (lbm) 2.054
Unit Weight (lb/ft3) 153.127
Density (g/cm3) 2.453
Moisture content (%) 0.32
Slake durability index (%) 99.06

“T” and “B” stands for top and bottom of the specimen, respectively. There are three

mark points at 120o around periphery of top or bottom.

Checks on the ASTM-D4543 Dimensional Requirements:

1. Side Smoothness and Straightness

Maximum and minimum along each line, and the difference between the maximum and

minimum

Δ0= 0.018 inch Δ120= 0.008 inch Δ240= 0.013 inch

Maximum Difference (DM) = 0.018 inch

Is DM≤0.02 inch? YES


75

2. End Flatness and Parallelism

Readings along two perpendicular diameters on the both top and bottom of the specimen,

at 1/8 inch intervals were taken (see the plots).

ΔT1= 0.003 inch ΔT2= 0.003 inch

ΔB1= 0.003 inch ΔB2= 0.002 inch

Maximum Difference (DM) = 0.003 inch

3. Perpendicularity

Is Δmax/D ≤ 0.0043? YES

1
mils

0
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
-1

-2
End 1 (TOP), Diameter 1

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES

3
2
1
mils

0
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
-1
-2
-3
End 1 (TOP), Diameter 2
76

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES

2.5
2
1.5
1
mils

0.5
0
-2 -1.5 -1 -0.5 -0.5 0 0.5 1 1.5 2
-1
-1.5
End 2 (BOTTOM), Diameter 1

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES

1
0.5
0
-2 -1.5 -1 -0.5 -0.5 0 0.5 1 1.5 2
mils

-1
-1.5
-2
-2.5
End 2 (BOTTOM), Diameter 2

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES
77

Sample 5U

Specimen ID 5U
Source location Mine “A”
Rock type Rhyolite
Equipment used for sample preparation Drill, Saw, Grinder
Storage condition Room
Moisture condition Oven dried
Specimen length (inch) 5.615
Specimen Diameter (inch) 2.496
Specimen Weight (lbm) 2.515
Unit Weight (lb/ft3) 158.210
Density (g/cm3) 2.534
Moisture content (%) 0.17
Slake durability index (%) 97.39

“T” and “B” stands for top and bottom of the specimen, respectively. There are three

mark points at 120o around periphery of top or bottom.

Checks on the ASTM-D4543 Dimensional Requirements:

1. Side Smoothness and Straightness

Maximum and minimum along each line, and the difference between the maximum and

minimum

Δ0= 0.008 inch Δ120= 0.009 inch Δ240= 0.009 inch

Maximum Difference (DM) = 0.009 inch

Is DM≤0.02 inch? YES


78

2. End Flatness and Parallelism

Readings along two perpendicular diameters on the both top and bottom of the specimen,

at 1/8 inch intervals were taken (see the plots).

ΔT1= 0.004 inch ΔT2= 0.003 inch

ΔB1= 0.003 inch ΔB2= 0.003 inch

Maximum Difference (DM) = 0.003 inch

3. Perpendicularity

Is Δmax/D ≤ 0.0043? YES

3
2
1
mils

0
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
-1
-2
-3
End 1 (TOP), Diameter 1

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES
79

4
3
2

mils
1
0
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
-1
End 1 (TOP), Diameter 2

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES

1.5
1
0.5
mils

0
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
-0.5
-1
-1.5
End 2 (BOTTOM), Diameter 1

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES

2.5
2
1.5
mils

1
0.5
0
-2 -1.5 -1 -0.5 -0.5 0 0.5 1 1.5 2
End 2 (BOTTOM), Diameter 2

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES
80

Sample 6U

Specimen ID 6U
Source location Mine “A”
Rock type Basalt
Equipment used for sample preparation Drill, Saw, Grinder
Storage condition Room
Moisture condition Oven dried
Specimen length (inch) 4.671
Specimen Diameter (inch) 2.500
Specimen Weight (lbm) 1.946
Unit Weight (lb/ft3) 146.626
Density (g/cm3) 2.349
Moisture content (%) 0.79
Slake durability index (%) 99.10

“T” and “B” stands for top and bottom of the specimen, respectively. There are three

mark points at 120o around periphery of top or bottom.

Checks on the ASTM-D4543 Dimensional Requirements:

1. Side Smoothness and Straightness

Maximum and minimum along each line, and the difference between the maximum and

minimum

Δ0= 0.002 inch Δ120= 0.004 inch Δ240= 0.007

inch

Maximum Difference (DM) = 0.007 inch

Is DM≤0.02 inch? YES


81

2. End Flatness and Parallelism

Readings along two perpendicular diameters on the both top and bottom of the specimen,

at 1/8 inch intervals were taken (see the plots).

ΔT1= 0.003 inch ΔT2= 0.003 inch

ΔB1= 0.004 inch ΔB2= 0.002 inch

Maximum Difference (DM) = 0.003 inch

3. Perpendicularity

Is Δmax/D ≤ 0.0043? YES

2
1
0
mils

-2 -1.5 -1 -0.5 0 0.5 1 1.5 2


-1
-2
-3
End 1 (TOP), Diameter 1

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES
82

2
1
0
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2

mils
-1
-2
-3
-4
End 1 (TOP), Diameter 2

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES

3
2
1
mils

0
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
-1
-2
End 2 (BOTTOM), Diameter 1

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES

0.5
0
-2 -1.5 -1 -0.5 -0.5 0 0.5 1 1.5 2
mils

-1
-1.5
-2
-2.5
End 2 (BOTTOM), Diameter 2

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES
83

Sample 7U

Specimen ID 7U
Source location Mine “A”
Rock type Rhyolite
Equipment used for sample preparation Drill, Saw, Grinder
Storage condition Room
Moisture condition Oven dried
Specimen length (inch) 5.137
Specimen Diameter (inch) 2.499
Specimen Weight (lbm) 2.249
Unit Weight (lb/ft3) 154.22
Density (g/cm3) 2.47
Moisture content (%) 0.34
Slake durability index (%) 97.39

“T” and “B” stands for top and bottom of the specimen, respectively. There are three

mark points at 120o around periphery of top or bottom.

Checks on the ASTM-D4543 Dimensional Requirements:

1. Side Smoothness and Straightness

Maximum and minimum along each line, and the difference between the maximum and

minimum

Δ0= 0.008 inch Δ120= 0.006 inch Δ240= 0.008

inch

Maximum Difference (DM) = 0.008 inch

Is DM≤0.02 inch? YES


84

2. End Flatness and Parallelism

Readings along two perpendicular diameters on the both top and bottom of the specimen,

at 1/8 inch intervals were taken (see the plots).

ΔT1= 0.001 inch ΔT2= 0.002 inch

ΔB1= 0.001 inch ΔB2= 0.002 inch

Maximum Difference (DM) = 0.002 inch

3. Perpendicularity

Is Δmax/D ≤ 0.0043? YES

0.4
0.2
0
-2 -1.5 -1 -0.5 -0.2 0 0.5 1 1.5 2
mils

-0.4
-0.6
-0.8
-1
-1.2
End 1 (TOP), Diameter 1

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES
85

mils
0
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
-1

-2
End 2 (BOTTOM), Diameter 2

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES

2
1.5
1
0.5
mils

0
-2 -1.5 -1 -0.5 -0.5 0 0.5 1 1.5 2
-1
-1.5
End 2 (BOTTOM), Diameter 2

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES

1.5

1
mils

0.5

0
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
-0.5
End 2 (BOTTOM), Diameter 2

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES
86

Sample 8U

Specimen ID 8U
Source location Mine “A”
Rock type Rhyolite
Equipment used for sample preparation Drill, Saw, Grinder
Storage condition Room
Moisture condition Oven dried
Specimen length (inch) 5.307
Specimen Diameter (inch) 2.499
Specimen Weight (lbm) 2.259
Unit Weight (lb/ft3) 149.94
Density (g/cm3) 2.40
Moisture content (%) 0.49
Slake durability index (%) 99.51

“T” and “B” stands for top and bottom of the specimen, respectively. There are three

mark points at 120o around periphery of top or bottom.

Checks on the ASTM-D4543 Dimensional Requirements:

1. Side Smoothness and Straightness

Maximum and minimum along each line, and the difference between the maximum and

minimum

Δ0= 0.004 inch Δ120= 0.004 inch Δ240= 0.008

inch

Maximum Difference (DM) = 0.008 inch

Is DM≤0.02 inch? YES


87

2. End Flatness and Parallelism

Readings along two perpendicular diameters on the both top and bottom of the specimen,

at 1/8 inch intervals were taken (see the plots).

ΔT1= 0.003 inch ΔT2= 0.002 inch

ΔB1= 0.003 inch ΔB2= 0.002 inch

Maximum Difference (DM) = 0.003 inch

3. Perpendicularity

Is Δmax/D ≤ 0.0043? YES

1.5
1
0.5
0
mils

-2 -1.5 -1 -0.5 -0.5 0 0.5 1 1.5 2


-1
-1.5
-2
-2.5
End 1 (TOP), Diameter 1

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES
88

2
1.5
1
0.5

mils
0
-2 -1.5 -1 -0.5 -0.5 0 0.5 1 1.5 2
-1
-1.5
End 2 (BOTTOM), Diameter 2

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES

2
mils

0
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
-1
End 2 (BOTTOM), Diameter 1

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES

2.5
2
1.5
mils

1
0.5
0
-2 -1.5 -1 -0.5 -0.5 0 0.5 1 1.5 2
End 2 (BOTTOM), Diameter 2

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES
89

Sample 11U

Specimen ID 11U
Source location Mine “A”
Rock type Rhyolite
Equipment used for sample preparation Drill, Saw, Grinder
Storage condition Room
Moisture condition Oven dried
Specimen length (inch) 6.04
Specimen Diameter (inch) 2.496
Specimen Weight (lbm) 2.259
Unit Weight (lb/ft3) 140.63
Density (g/cm3) 2.25
Moisture content (%) 1.67
Slake durability index (%) 98.37

“T” and “B” stands for top and bottom of the specimen, respectively. There are three

mark points at 120o around periphery of top or bottom.

Checks on the ASTM-D4543 Dimensional Requirements:

1. Side Smoothness and Straightness

Maximum and minimum along each line, and the difference between the maximum and

minimum

Δ0= 0.011 inch Δ120= 0.008 inch Δ240= 0.006

inch

Maximum Difference (DM) = 0.008 inch

Is DM≤0.02 inch? YES


90

2. End Flatness and Parallelism

Readings along two perpendicular diameters on the both top and bottom of the specimen,

at 1/8 inch intervals were taken (see the plots).

ΔT1= 0.002 inch ΔT2= 0.003 inch

ΔB1= 0.002 inch ΔB2= 0.004 inch

Maximum Difference (DM) = 0.003 inch

3. Perpendicularity

Is Δmax/D ≤ 0.0043? YES

1.2
1
0.8
0.6
mils

0.4
0.2
0
-2 -1.5 -1 -0.5 -0.2 0 0.5 1 1.5 2
-0.4
End 1 (TOP), Diameter 1

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES
91

1.5
1
0.5

mils
0
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
-0.5
-1
-1.5
End 1 (TOP), Diameter 2

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES

1.5
1
0.5
mils

0
-2 -1.5 -1 -0.5 -0.5 0 0.5 1 1.5 2

-1
-1.5
End 2 (BOTTOM), Diameter 1

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES

1
mils

0
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
-1

-2
End 2 (BOTTOM), Diameter 2

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES
92

Sample W-1U

Specimen ID W-1U
Source location Mine “A”
Rock type Rhyolite
Equipment used for sample preparation Drill, Saw, Grinder
Storage condition Room
Moisture condition Oven dried
Specimen length (inch) 5.075
Specimen Diameter (inch) 2.496
Specimen Weight (lbm) 2.286
Unit Weight (lb/ft3) 159.09
Density (g/cm3) 2.55
Moisture content (%) 0.67
Slake durability index (%) 99.3

“T” and “B” stands for top and bottom of the specimen, respectively. There are three

mark points at 120o around periphery of top or bottom.

Checks on the ASTM-D4543 Dimensional Requirements:

1. Side Smoothness and Straightness

Maximum and minimum along each line, and the difference between the maximum and

minimum

Δ0= 0.019 inch Δ120= 0.012 inch Δ240= 0.011

inch

Maximum Difference (DM) = 0.008 inch

Is DM≤0.02 inch? YES


93

2. End Flatness and Parallelism

Readings along two perpendicular diameters on the both top and bottom of the specimen,

at 1/8 inch intervals were taken (see the plots).

ΔT1= 0.00 inch ΔT2= 0.00 inch

ΔB1= 0.01 inch ΔB2= 0.008 inch

Maximum Difference (DM) = 0.008 inch

3. Perpendicularity

Is Δmax/D ≤ 0.0043? YES

1
0.8
0.6
mils

0.4
0.2
0
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
End 1 (TOP), Diameter 1

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES
94

1
0.8
0.6

mils
0.4
0.2
0
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
End 1 (TOP), Diameter 2

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES

6
4
2
mils

0
-2 -1.5 -1 -0.5 -2 0 0.5 1 1.5 2

-4
-6
End 2 (BOTTOM), Diameter 1

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES

6
4
2
mils

0
-2 -1.5 -1 -0.5 -2 0 0.5 1 1.5 2
-4
-6
End 2 (BOTTOM), Diameter 2

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES
95

Sample W-2U

Specimen ID W-2U
Source location Mine “A”
Rock type Basalt
Equipment used for sample preparation Drill, Saw, Grinder
Storage condition Room
Moisture condition Oven dried
Specimen length (inch) 5.562
Specimen Diameter (inch) 2.481
Specimen Weight (lbm) 2.040
Unit Weight (lb/ft3) 131.12
Density (g/cm3) 2.10
Moisture content (%) 1.91
Slake durability index (%) 99.39

“T” and “B” stands for top and bottom of the specimen, respectively. There are three

mark points at 120o around periphery of top or bottom.

Checks on the ASTM-D4543 Dimensional Requirements:

1. Side Smoothness and Straightness

Maximum and minimum along each line, and the difference between the maximum and

minimum

Δ0= 0.027 inch Δ120= 0.039 inch Δ240= 0.022

inch

Maximum Difference (DM) = 0.039 inch

Is DM≤0.02 inch? NO
96

2. End Flatness and Parallelism

Readings along two perpendicular diameters on the both top and bottom of the specimen,

at 1/8 inch intervals were taken (see the plots).

ΔT1= 0.001 inch ΔT2= 0.003 inch

ΔB1= 0.001 inch ΔB2= 0.001 inch

Maximum Difference (DM) = 0.008 inch

3. Perpendicularity

Is Δmax/D ≤ 0.0043? YES

1.2
1
0.8
0.6
mils

0.4
0.2
0
-2 -1.5 -1 -0.5 -0.2 0 0.5 1 1.5 2
-0.4
End 1 (TOP), Diameter 1

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES
97

2
1
0

mils
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
-1
-2
-3
End 1 (TOP), Diameter 2

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES

1.5

1
mils

0.5

0
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
-0.5
End 2 (BOTTOM), Diameter 1

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES

1.5

1
mils

0.5

0
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
-0.5
End 2 (BOTTOM), Diameter 2

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES
98

Sample W-5U

Specimen ID W-5U
Source location Mine “A”
Rock type Basalt
Equipment used for sample preparation Drill, Saw, Grinder
Storage condition Room
Moisture condition Oven dried
Specimen length (inch) 5.214
Specimen Diameter (inch) 2.498
Specimen Weight (lbm) 1.910
Unit Weight (lb/ft3) 129.18
Density (g/cm3) 2.07
Moisture content (%) 4.34
Slake durability index (%) 94.10

“T” and “B” stands for top and bottom of the specimen, respectively. There are three

mark points at 120o around periphery of top or bottom.

Checks on the ASTM-D4543 Dimensional Requirements:

1. Side Smoothness and Straightness

Maximum and minimum along each line, and the difference between the maximum and

minimum

Δ0= 0.011 inch Δ120= 0.009 inch Δ240= 0.011

inch

Maximum Difference (DM) = 0.011 inch

Is DM≤0.02 inch? YES


99

2. End Flatness and Parallelism

Readings along two perpendicular diameters on the both top and bottom of the specimen,

at 1/8 inch intervals were taken (see the plots).

ΔT1= 0.003 inch ΔT2= 0.002 inch

ΔB1= 0.005 inch ΔB2= 0.001 inch

Maximum Difference (DM) = 0.008 inch

3. Perpendicularity

Is Δmax/D ≤ 0.0043? YES

2.5
2
1.5
1
mils

0.5
0
-2 -1.5 -1 -0.5 -0.5 0 0.5 1 1.5 2
-1
-1.5
End 1 (TOP), Diameter 1

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES
100

0.5
0
-2 -1.5 -1 -0.5 -0.5 0 0.5 1 1.5 2

mils
-1
-1.5
-2
-2.5
End 1 (TOP), Diameter 2

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES

2.5
2
1.5
1
mils

0.5
0
-2 -1.5 -1 -0.5 -0.5 0 0.5 1 1.5 2
-1
End 2 (BOTTOM), Diameter 1

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES

0.5
0
-2 -1.5 -1 -0.5 -0.5 0 0.5 1 1.5 2
mils

-1
-1.5
-2
-2.5
End 2 (BOTTOM), Diameter 2

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES
101

Sample W-6U

Specimen ID W-6U
Source location Mine “A”
Rock type Basalt
Equipment used for sample preparation Drill, Saw, Grinder
Storage condition Room
Moisture condition Oven dried
Specimen length (inch) 5.37
Specimen Diameter (inch) 2.482
Specimen Weight (lbm) 1.843
Unit Weight (lb/ft3) 122.58
Density (g/cm3) 1.96
Moisture content (%) 4.14
Slake durability index (%) 90.73

“T” and “B” stands for top and bottom of the specimen, respectively. There are three

mark points at 120o around periphery of top or bottom.

Checks on the ASTM-D4543 Dimensional Requirements:

1. Side Smoothness and Straightness

Maximum and minimum along each line, and the difference between the maximum and

minimum

Δ0= 0.013 inch Δ120= 0.009 inch Δ240= 0.007

inch

Maximum Difference (DM) = 0.013 inch

Is DM≤0.02 inch? YES


102

2. End Flatness and Parallelism

Readings along two perpendicular diameters on the both top and bottom of the specimen,

at 1/8 inch intervals were taken (see the plots).

ΔT1= 0.001 inch ΔT2= 0.001 inch

ΔB1= 0.001 inch ΔB2= 0.002 inch

Maximum Difference (DM) = 0.002 inch

3. Perpendicularity

Is Δmax/D ≤ 0.0043? YES

1.2
1
0.8
0.6
mils

0.4
0.2
0
-2 -1.5 -1 -0.5 -0.2 0 0.5 1 1.5 2
End 1 (TOP), Diameter 1

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES
103

1.5
1
0.5

mils
0
-2 -1.5 -1 -0.5 -0.5 0 0.5 1 1.5 2

-1
-1.5
End 1 (TOP), Diameter 2

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES

1.5

1
mils

0.5

0
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
-0.5
End 2 (BOTTOM), Diameter 1

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES

1.5
1
0.5
mils

0
-2 -1.5 -1 -0.5 -0.5 0 0.5 1 1.5 2
-1
-1.5
End 2 (BOTTOM), Diameter 2

Is the maximum difference between the smooth curve fit and strain line best fit≤0.001

inch? YES
104

Appendix-2: Strain Gages Results

In the design of rock structures such as dam foundations or vertical mineshafts, the

deformation of the rock is of prime importance since excessive deformation may result in

failure of the dam wall in severe misalignment of the shaft steelwork. On the other hand,

the stability of rock structures such as tunnels, open pits or underground excavations is

dependent upon the strength and failure characteristics of the rock. The deformation

characteristics of rock most conveniently studied on cylindrical specimens, subject it to

uniaxial compression. Strain gages, bonded to the rock surface are the most reliable

measuring devices is and the convenience of using gages would normally outweigh the

relatively high cost of the gages. when that deformation characteristics of rocks which

occur in structures subjected to low levels of stress are being studied, great care must be

taken to ensure that the load and strain and measured with sufficient accuracy. Strain

gages are sensors that are used to measure compressive or tensile strain in a part due to an

applied load.

These sensors can be integrated into other types of sensors, such as load cells. Strain

gages are resistive sensors, which mean the output of the sensor is based on the change of

resistance of the gage, which is essentially a variable resistor. A strain gage is essentially

a single wire that has been wound back and forth to form a rectangular matrix. Modern

strain gages are made using semiconductor fabrication methods to ensure quality and

repeatability. When a strain gage is pulled or compressed, each part of the rectangular

matrix extends or compresses, changing the overall length of the wire. As the length

changes, the resistance of the wire changes, and the resistance of the wire is proportional

to the strain induced on the gage.


105

If electrical resistance strain gages are used, the length of the gages over which axial and

circumferential strains are determined should be at least ten grain diameters in magnitude

and the gages should not encroach within D/2 of the specimen ends, where D is the

diameter of the specimen.

When selecting a strain gage for a measurement application, the characteristics of the

selected sensor must match the characteristics of the system or object being measured.

Strain gages are available in many different sizes and configurations, so understanding

the variables will result in proper gage selection.

Alloy

The alloy is the type of material used for the strain gage wire. A variety of alloys are

available for strain gages, depending on the specific application. Constantan is an all-

purpose alloy, which is applicable for most general use applications. Annealed constantan

is useful for high-strain applications. Isoelastic alloys are appropriate for dynamic strain

applications or fatigue cycling. The Karma alloy is most applicable for long-term

monitoring, as it maintains performance well over time.

Gage Factor

The gage factor is the sensitivity of the strain gage, or the relationship between resistance

and strain. The gage factor depends on the specific strain gage wire material. The

majority of strain gages have a gage factor around 2.


106

Gage Length

Gage length is the length of the gage matrix in the direction of strain. Larger gage lengths

are more resistant to localized temperature and strain effects, and are easier to handle due

to their larger size. Smaller gage lengths are appropriate for confined spaces or to

measure very localized strains. Gages with large gage lengths are often less expensive

than gages with small gage lengths.

Resistance

Resistance refers to the electrical resistance of the gage wire, measured in ohms.

Common strain gage resistances are 350 ohms and 120 ohms. If a half-bridge or quarter-

bridge circuit is used, the other arms of the bridge must be filled with dummy or active

gages with the same resistance.

Self-Temperature Compensation (S-T-C)

S-T-C is the ability of the gage to resist environmental thermal changes. The S-T-C

number is related to the coefficient of thermal expansion of the strain gage materials.

Rosette:

A rosette is a configuration of two or more individual gages in one package. Rosettes can

include two gages to measure strain in two perpendicular axes. In this project because of

financial and technical considerations, larger strain gages could not be available to bond

to HQ rock samples. The CEA-13125WT-350 rosettes were used to measure the core

sample deformations under a uniaxial compression load.


107

Table A-2-1: Rosette CEA-13-125WT-350 properties


GAGE PATTERN GAGE DESIGNATION RES. IN OHMS
ES: Each CP: Complete Inch
section pattern
S: Section M: Matrix Millimeter
125WT
Two-element 90o tee stacked rosette. Exposed
solder tab area o.1 × 0.07 in (2.5 × 1.8 mm).
Maximum operating temperature + 150 F (+65o
c)
GAGE OVERALL GRID OVERALWIDTH
LENGTH LENGTH WIDTH CEA-xx- 350±5%
0.125 ES 0.325 CP 0.18 ES 0.325 CP 125WT-350
3.18 ES 8.2 CP 4.57 ES 8.26 CP
MATRIX 0.42L × 0.42 W 10.7L × 10.7 W
SIZE

The operating equation for strain gages is the definition of gage factor (GF), which is

how the gage‟s resistance changes as a function of its change in length as it is strained.

R=ρL/A Equ. A-2.1

In which, R is the resistance, ρ is the resistivity (as a material property of the wire which

is used in the strain gages), L is the length of the wire, and A is the cross sectional area of

the wire. To find out how the resistance changes when the length of the wire changes,

mercury wire is considered as a liquid material which is incompressible (no changes in

the volume). The reason that the mercury is used to carry out the gage factor expression

is the A (area of the wire) is not constant and by changing the length of the wire:

Equ. A-2.2
108

Equ. A-2.3

Equ. A-2.4



Equ. A-2.5

From the above calculation, the definition of the gage factor (GF) is as follow:

GF = (ΔR/R0) / (ΔL/L0) Equ. A-2.6

R0 is the initial, unstrained resistance, ΔR is the change in resistance resulting from the

application of the load, and ΔL/L0 (ε) is the strain. The definition of the gage factor may

be re-arranged as:

ε = ((Rstrained-R0)/R0)/GF Equ. A-1.3

For the gages used in this project, the GF=2.12. The unstrained initial resistance (R0) for

each gage was observed from the first few data point recorded in the data file. The

calculations were carried out to 6 decimal places based on the measured resistance

values.

Once an appropriate value for R0 has been determined for each gage, the strain can be

readily calculated from the equation A 1-7, as a function of the applied load. Usually, the

strain values are averaged to reduce the effects due to non-parallel ends and so for. The

procedures of the calculation of the averaged strains are explained in the appendix 1.

Please refer to ASTM D7012 “Standard Test Method for Compressive Strength and

Elastic Moduli of Intact Rock Core Specimen under Varying States of Stress and

Temperatures”, section11 and 12, for guidelines on calculating the elastic moduli. The
109

procedure of the Poisson‟s ration and elastic moduli calculations is explained in the

appendix 1.

As it has been shown in the table A-2-1 and from the observations of the core samples,

the rosette which was used in this study is small. Therefore, it is important to ensure that

the load and strain and measure with sufficient accuracy. Three core samples with larger

grain size were considered to evaluate the accuracy of the rosettes. Four rosettes were

bonded to each of the samples: two rosettes next to each other on both sides of the core

samples. The procedure of strain gage rosettes installation is as follow:

1. Choose approximate location for the gages, and use the center finder to draw

marks along the diameter of the core, at the edges.

2. Clean off broad area with alcohol and Kim-wipes, allow drying.

3. Apply a thin coating of Loctite adhesive to an area larger than the gage. Wipe off

excess Loctite with manila card, allow drying for 30 minutes.

4. Sand each “patch” area smooth, and make sure it is large enough for the gage.

5. Wipe the sanded area with alcohol, allow drying.

6. Use the combination square to draw an axial line with segment along one side of

the patch. This is used for alignment of the gage, or its midpoint, is located.

7. Use a piece of scotch tape about 2 ½” long. Position this with sticky side up on

the work table.

8. Use tweezers to position the strain gage on the scotch tape. The copper soldering

tabs should be in contact with the tape.


110

9. Paint a very thin Layer of Loctite on the gage. All the gage should be covered

with Loctite, but with as thin a layer a possible. It is fine to get some of the

Loctite on the tape.

10. Lift the tape off the work table, and very carefully position the gage on the

specimen where the lines have been drawn. Roll the gage on, on edge first, and

pressing any air bubbles out. Press very firmly, and rub the tape onto the

specimen.

11. Allow the Loctite to cure for at least 30 minutes. Then very carefully, peel the

tape off, starting at one end, pulling the tape back at a sharp angle.

A total of 10 UCS core samples were tested that were divided to three main groups:

1- Three core samples with four rosettes (8 strain gages) to assess if the size of the

strain gages is acceptable to install on the HQ core samples.

2- Five core samples with two rosettes to develop correlation relations of the

mechanical properties of the rock with indirect test methods.

3- Two core sample without any strain gages.

1-Study the applicability of CEA-13-125WT-350 rosettes to the HQ core


samples:

Fig. A-2-1: Two rosettes were installed next to each other on both side of the specimens to
evalaute the accuaracy of strain gages.
111

Specimen 5U:

Information about the specimen 5U is available in the appendix 1 “SPECIMEN

PREPRATION REPORT”. To evaluate the performance of the strain gages, four rosettes

were bonded to this specimen, and named rosette 1 to rosette 4. For each two rosettes,

uniaxial compression test was performed using MTS machine load control configuration,

to the displacement of less than 1% of the length of the specimen. Four uniaxial

compression tests were run, three of which with the maximum load of 65 kips. The

specimen was under load of zero kips, one kips, five kips, and the load was increased to

65 kips with increase of 5 kips in each step. The computer records the strain gages

resistance each five seconds. Five resistance numbers were recorded and averaged as

“strain gages resistance” of the each step. The load was decreased under the same pattern

from 65 kips to 5 kips, 1 kips, and zero force to see the strain gages behavior. In the last

run (rosette 2 and 4), the uniaxial compression test were performed using MTS machine

displacement control configuration, with displacement rate of 0.16 mm/min to break the

sample.

The results of the strain gages performance tests on the specimen 5U are as follow. In the

following curves, “R-i” stands for stress and axial strain of the rosette “i”; “R-i-j” shows

two connected rosettes to the computer; and, “Circ Strain” stands for circumferential

strain.
112

Fig. A-2-2: Two rosettes were bonded next to each other to compare the results of the gages.

Fig. A-2-3: Strain gages data is recorded each five seconds using the computer.
113

120

100

R-1 (R-1-2)
80
R-2 (R-1-2)
Stress (MPa)

R-1 (R-1-3)
60
R-3 (R-1-3)
R-4 (R-3-4)
40 R-3 (R-3-4)
R-4 (R-2-4) Final
20 R-2 (R-2-4) Final

0
-0.0025 -0.002 -0.0015 -0.001 -0.0005 0
Strain

Fig. A-2-4: Stress-Strain curve based on the data recorded for a total of four runs

0.0007

0.0006

0.0005 R-1 (R-1-2)


R-2 (R-1-2)
Circ Strain

0.0004
R-1 (R-1-3)
R-3 (R-1-3)
0.0003
R-4 (R-3-4)
0.0002 R-3 (R-3-4)
R-4 (R-2-4) Final
0.0001 R-2 (R-2-4) Final

0
-0.0025 -0.002 -0.0015 -0.001 -0.0005 0
Axial Strain
114

Fig. A-2-5: Axial strain- circumferential strain curve based on the data recorded for a total of four

runs

100 0.0003
80 0.00025
Stress (MPa)

Circ Strain
0.0002
60
R-1 0.00015
R-1 40 0.0001
20 R-2
R-2 0.00005
0 0
-0.002 -0.0015 -0.001 -0.0005 0 -0.002 -0.0015 -0.001 -0.0005 0
Strain Axial Strain

Fig. A-2-6: Data recorded for Rosettes 1-2

Table A-2-2: Results from the rosettes 1-2 (specimen 5U)

Rosette Number Poisson‟s Average Young‟s Modulus


Ratio (GPa)
1 0.19 66.01
2 0.19 65.88
100 0.0005

80 0.0004
Stress (MPa)

Circ Strain

60 0.0003

R-1 40 0.0002
R-1
R-3 20 R-3 0.0001

0 0
-0.002 -0.0015 -0.001 -0.0005 0 -0.002 -0.0015 -0.001 -0.0005 0
Strain Axial Strain

Fig. A-2-7: Data recorded for Rosettes 1-3

Table A-2-3: Results from the rosettes 1-3 (specimen 5U)

Rosette Number Poisson‟s Average Young‟s Modulus


Ratio (GPa)
1 0.18 65.55
3 0.16 71.00
115

100 0.0003
80 0.00025

Stree (MPa)

Circ Strain
60 0.0002
R-4 0.00015
40 R-4
0.0001
R-3 20 R-3 0.00005
0
0
-0.0015 -0.001 -0.0005 0
-0.0015 -0.001 -0.0005 0
Strain Axial Strain

Fig. A-2-8: Data recorded for Rosettes 3-4

Table A-2-4: Results from the rosettes 3-4 (specimen 5U)

Rosette Number Poisson‟s Average Young‟s Modulus


Ratio (GPa)
3 0.22 66.9
4 0.16 70.88

100 0.0007
0.0006
80 R-4
Stress (MPa)

0.0005
Circ Strain

60 R-2 0.0004
40 0.0003
R-4
0.0002
R-2 20 0.0001
0 0
-0.0025 -0.002 -0.0015 -0.001 -0.0005 0 -0.0025 -0.002 -0.0015 -0.001 -0.0005 0
Strain Axial Strain

Fig. A-2-10: Data recorded for Rosettes 2-4

Table A-2-5: Results from the rosettes 2-4 (specimen 5U)

Rosette Number Poisson‟s Average Young‟s Modulus


Ratio (GPa)
1 0.22 65.44
2 0.18 62.90
116

Specimen W-1U:

Information about the specimen W-1U is available in the appendix ()“SPECIMEN

PREPRATION REPORT”. To evaluate the performance of the strain gages, four rosettes

were bonded to this specimen, and named rosette 1 to rosette 4. For each two rosettes,

uniaxial compression test was performed using MTS machine load control configuration,

to the displacement of less than 1% of the length of the specimen. Four uniaxial

compression tests were run, three of which with the maximum load of 50 kips. The

specimen was under load of zero kips, one kips, five kips, and the load was increased to

50 kips with increase of 5 kips in each step. The computer records the strain gages

resistance each five seconds. Five resistance numbers were recorded and averaged as

“strain gages resistance” of the each step. The load was decreased under the same pattern

from 65 kips to 5 kips, 1 kips, and zero force to see the strain gages behavior. In the last

run (rosette 2 and 4), the uniaxial compression test were performed using MTS machine

displacement control configuration, with displacement rate of 0.16 mm/min to break the

sample.

The results of the strain gages performance tests on the specimen W-1U are as follow. In

the following curves, “R-i” stands for stress and axial strain of the rosette “i”; “R-i-j”

shows two connected rosettes to the computer; and, “Circ Strain” stands for

circumferential strain.
117

Fig. A-2-11: Specimen W-1U with four rosettes to evaluate the accuracy of the gages

60

50
R-1 (R-1-2)
40 R-2 (R-1-2)
Stress (MPa)

R1 (R-1-3)
30
R-3 (R-1-3)
20 R-4 (R-3-4)
R-3 (R-3-4)
10 R-4 (R-2-4) Final
R-2 (R-2-4) Final
0
-0.002 -0.0015 -0.001 -0.0005 0 0.0005 0.001
Strain

Fig.A-2-12: Stress-Strain curve based on the data recorded for a total of four runs
118

0.00025

0.0002

0.00015
R-1 (R-1-2)

0.0001 R-2 (R-1-2)


R-1 (R-1-3)
Circ Strain

0.00005 R-3 (R-1-3)


R-4 (R-3-4)
0 R-3 (R-3-4)
-0.002 -0.0015 -0.001 -0.0005 0 0.0005 0.001
R-4 (R-2-4) Final
-0.00005 R-2 (R-2-4)

-0.0001

-0.00015
Axial Strain

Fig. A-2-13: Axial strain- circumferential strain curve based on the data recorded for a
total of four runs

80 0.00025
0.0002
60
Stress (MPa)

0.00015
Circ Strain

R-1 (R-1-2) 0.0001


40
R-1 (R-1-2) 0.00005
R-2 (R-1-2)
20 0
R-2 (R-1-2) -0.0015 -0.001 -0.0005
-0.00005 0 0.0005
0
-0.0001
-0.002 -0.001 0 0.001
Strain Axial Strain

Fig. A-2-14: Data recorded for Rosettes 1-2

Table A-2-6: Results from the rosettes 1-2 (specimen W-1U)

Rosette Number Poisson‟s Average Young‟s Modulus


Ratio (GPa)
1 0.18 45.68
2 0.15 43.15
119

80 0.0002
R-1 (R-1-3) 0.00015
60

Circ Strain
Stress (MPa) 0.0001 R-1 (R-1-3)
R-3 (R-1-3) 0.00005
40
R-3 (R-1-3)
0
20 -0.001 -0.0005
-0.00005 0 0.0005 0.001
-0.0001
0
-0.001 -0.0005 0 0.0005 0.001 -0.00015
Strain Axial Strain

Fig. A-2-15: Data recorded for Rosettes 1-3

Table A-2-7: Results from the rosettes 1-3 (specimen W-1U)

Rosette Number Poisson‟s Average Young‟s Modulus


Ratio (Gpa)
1 0.18 48.31
3 0.14 67.39

80 0.00025
70
0.0002
60
Stress (MPa)

50 0.00015
Circ Strain

40
R-4 (R-3-4) 0.0001
R-4 (R-3-4) 30
20 R-3 (R-3-4) 0.00005
R-3 (R-3-4) 10
0 0
-0.0015 -0.001 -0.0005 0 -0.0015 -0.001 -0.0005 0
-0.00005
Strain Axial Strain

Fig. A-2-16: Data recorded for Rosettes 3-4

Table A-2-8: Results from the rosettes 3-4 (specimen W-1U)

Rosette Number Poisson‟s Average Young‟s Modulus


Ratio (GPa)
4 0.17 60.81
3 0.14 66.54
120

60
0.00025
50
0.0002
Stress (MPa) 40

Circ Strain
0.00015
30 R-4 (R-2-4) 0.0001
R-4 (R-2- 20
4) 10 R-2 (R-2-4) 0.00005
0
0
-0.002 -0.0015 -0.001 -0.0005 0 0.0005
-0.002 -0.0015 -0.001 -0.0005 0 0.0005 -0.00005
Strain Axial Strain

Fig. A-2-17: Data recorded for Rosettes 2-4

Table A-2-9: Results from the rosettes 2-4 (specimen W-1U)

Rosette Number Poisson‟s Average Young‟s Modulus


Ratio (Gpa)
4 0.17 60.37
2 0.15 47.29

Specimen 11U:

Information about the specimen 11U is available in the appendix ()“SPECIMEN

PREPRATION REPORT”. To evaluate the performance of the strain gages, four rosettes

were bonded to this specimen, and named rosette 1 to rosette 4. For each two rosettes,

uniaxial compression test was performed using MTS machine load control configuration,

to the displacement of less than 1% of the length of the specimen. Two uniaxial

compression tests were run, three of which with the maximum load of 65 kips. The

specimen was under load of zero kips, one kips, five kips, and the load was increased to

65 kips with increase of 5 kips in each step. The computer records the strain gages

resistance each five seconds. Five resistance numbers were recorded and averaged as

“strain gages resistance” of the each step. The load was decreased under the same pattern

from 65 kips to 5 kips, 1 kips, and zero force to see the strain gages behavior. The
121

specimen 11U was broken in the second run. There was an obvious weakness plane

which is shown in the figure A-2-18. The sample was broken through this weakness

plane. The results of the test show nonlinear behavior of the specimen. However, indirect

tests which were performed on this rock from same core indicate the rock was fairly hard.

Consequently, because of existence of an obvious weakness plane and failure which

occurred, the results are not reliable for the further conclusion.

The results of the strain gages performance tests on the specimen 11U are as follow. In

the following curves, “R-i” stands for stress and axial strain of the rosette “i”; “R-i-j”

shows two connected rosettes to the computer; and, “Circ Strain” stands for

circumferential strain.

Fig. A-2-18: Specimen 11U; there is no acceptable results from the specimen because of a
weakness plane in the specimen.
122

100

80

60
Stress (MPa)

R-1 (R-1-2)
40 R-2 (R-1-2)
R-1 (R-1-3)
20 R-3 (R-1-3)

0
-0.005 -0.004 -0.003 -0.002 -0.001 0 0.001

-20
Strain

Fig. A-2-19: Stress-Strain curve based on the data recorded for a total of four runs

0.0012

0.001

0.0008

R-1 (R-1-2)
Circ Strain

0.0006
R-2 (R-1-2)
0.0004 R-1 (R-1-3)
R-3 (R-1-3)
0.0002

0
-0.005 -0.004 -0.003 -0.002 -0.001 0 0.001
-0.0002
Axial Strain

Fig. A-2-20: Axial strain- circumferential strain curve based on the data recorded for a total of
four runs
123

100 0.0012
0.001
Stress (MPa) 80

Circ Strain
0.0008
60
0.0006
R-1 (R-1-2) 40 R-1 (R-1-2) 0.0004
20 0.0002
R-2 (R-1-2) R-2 (R-1-2)
0 0
-0.005 -0.004 -0.003 -0.002 -0.001 0 -0.005 -0.004 -0.003 -0.002 -0.001 0
Strain Axial Strain

Fig. A-2-21: Data recorded for Rosettes 1-2

Table A-2-10: Results from the rosettes 1-2 (specimen W-1U)

Rosette Number Poisson‟s Average Young‟s Modulus


Ratio (Gpa)
1 0.18 21.20
2 0.26 26.11

100 0.0012
80 0.001
Stress (MPa)

0.0008
Circ Strain

60
R-1 (R-1-3) 0.0006
40 0.0004
R-1 (R-1-3)
20 R-3 (R-1-3) 0.0002
R-3 (R-1-3)
0
0 -0.005 -0.004 -0.003 -0.002 -0.001
-0.005 -0.004 -0.003 -0.002 -0.001 0 0.001 -0.0002 0 0.001
Strain Axial Strain
Fig. A-2-22: Data recorded for Rosettes 1-3

Table A-2-11: Results from the rosettes 2-4 (specimen W-1U)

Rosette Number Poisson‟s Average Young‟s Modulus


Ratio (Gpa)
1 0.20 24.25
3 0.36 33.12
124

2-Calculation of the mechanical properties of intact rock samples with two


rosettes

To calculate the mechanical properties of the intact rock sample (E and ν), the slope of

axial and lateral curves are measured based on the methods which are suggested by

ASTM standard (ASTM-D7012-10). Two rosettes were bonded to each sample with 180o

degree angle to each other. To compare the results, the slopes of axial and lateral curves

for each rosette are calculated. In the next step, the results from rosette 1 and 2 are

averaged, which is shown as “Average” in the following tables.

To find the errors of the strain gages and the calculations, “Strain Average” is calculated

for each rosette as the average of the strain results from the gages. This can be used as a

bench mark for the error calculation (mean of the values).

For each of the samples, table “sample ID-1” shows the results of Young‟s modulus and

Poisson‟s ratio according to the different calculations for the Young‟s Modulus. Table

“Sample ID-2” shows the “logarithmic standard deviation” (equation 1-1) of the each

method using the “Average Strain” as a mean (actual) value.

The Young‟s modulus and Poisson‟s ratio are calculated using three methods, as follows:

 Tangent modulus, at a stress level at 50% of the maximum strength,

 Secant modulus, from 5% of maximum stress to 95% maximum strength,

 Average modulus, slope of the straight line portion of the stress-strain curve (with

the r = 0.95).

Comparison between the errors of each method for five different samples indicates that

the “Average Modulus of Linear Portion of Axial Stress-Strain Curve” is the best method
125

for Young‟s modulus and Poisson‟s ratio calculation. It is important to note that the linear

portion of the stress-strain curve has been identified using an almost complicated and

time consuming method. For each rosette, a very short and perfectly linear portion of the

stress-strain curve has been chosen. Using error calculation, the number of the data has

been increased until to get the r = 0.95.

Although the average modulus calculation is the best method with the least error, the

errors of the all three methods are acceptable and the results are reasonably good. The

error calculation for other rock type, strain gages, or other compression test machines

may be different.

Sample 6U

Table 6U-1: Results from the rosettes based on three Young‟s modulus calculation methods.
Young‟s Modulus Calculation Method Young‟s Modulus Slope of Lateral Poisson‟s
(GPa) Curve (GPa) Ratio
Rosette 1 36.11 156.68 0.23
Tangent modulus Rosette 2 48.39 192.59 0.25
Average 42.25 174.63 0.24
Strain Average 41.36 172.79 0.24
Rosette 1 36.04 130.90 0.28
Secant Modulus Rosette 2 47.41 153.02 0.31
Average 41.73 141.96 0.29
Strain Average 40.95 141.10 0.29
Rosette 1 37.41 152.07 0.25
Average Modulus Rosette 2 48.66 198.63 0.24
Average 43.04 175.35 0.25
Strain Average 42.30 172.26 0.25

Table 6U-2: Error calculation for three Young‟s modulus calculation methods.
Young‟s Modulus Logarithmic Standard Deviation (β)
Calculation Method Young‟s Modulus Slope of Lateral Curve
(GPa) (GPa) Poisson‟s Ratio
Tangent Modulus 0.120 0.085 0.036
Secant Modulus 0.113 0.064 0.049
Average Modulus 0.108 0.110 0.002
126

180
160
140
120
Stress (MPa) 100
80
R-1-Axial
60 R-1-Circ
40 R-2-Circ
20 R-2-Axial
0
-0.002 -0.001 0 0.001 0.002 0.003 0.004 0.005 0.006
Strain
Fig. 6U: Stress-Strain curve for rosette 1 (R-1) and rosette 2 (R-2)

Sample 8U
Table 8U-1: Results from the rosettes based on three Young‟s modulus calculation methods.
Young‟s Modulus Calculation Young‟s Modulus Slope of Lateral Poisson‟s
Method (GPa) Curve (GPa) Ratio
Rosette 1 42.91 200.78 0.21
Tangent modulus Rosette 2 54.40 385.45 0.14
Average 48.66 293.12 0.17
Strain Average 47.98 264.03 0.18
Rosette 1 40.37 177.06 0.23
Secant Modulus Rosette 2 54.89 315.79 0.17
Average 47.63 246.43 0.19
Strain Average 46.52 226.90 0.21
Rosette 1 40.99 188.40 0.22
Average Modulus Rosette 2 56.61 368.89 0.15
Average 48.80 278.64 0.18
Strain Average 47.55 249.41 0.19

Table 8U-2: Error calculation for three Young‟s modulus calculation methods.
Young‟s Modulus Logarithmic Standard Deviation (β)
Calculation Method Young‟s Modulus Slope of Lateral Curve
(GPa) (GPa) Poisson‟s Ratio
Tangent Modulus 0.097 0.276 0.181
Secant Modulus 0.126 0.243 0.118
Average Modulus 0.133 0.285 0.155
127

250

200

Stress (MPa)
150

100 R-1-Axial
R-1-Circ
50 R-2-Circ
R-2-Axial
0
-0.002 -0.001 0 0.001 0.002 0.003 0.004 0.005 0.006
Strain

Fig. 8U: Stress-Strain curve for rosette 1 (R-1) and rosette 2 (R-2)

Sample W-2U
Table W-2U-1: Results from the rosettes based on three Young‟s modulus calculation methods.
Young‟s Modulus Young‟s Modulus Slope of Lateral Poisson‟s
Calculation Method (GPa) Curve (GPa) Ratio
Rosette 1 23.16 109.93 0.21
Tangent modulus Rosette 2 30.60 144.87 0.21
Average 26.88 127.40 0.21
Strain Average 26.36 125.01 0.21
Rosette 1 19.94 82.80 0.24
Secant Modulus Rosette 2 29.01 121.44 0.24
Average 24.48 102.12 0.24
Strain Average 23.64 98.47 0.24
Rosette 1 25.23 125.52 0.20
Average Modulus Rosette 2 32.67 170.21 0.19
Average 28.95 147.87 0.20
Strain Average 28.47 144.49 0.20

Table W-2U-2: Error calculation for three Young‟s modulus calculation methods.
Young‟s Modulus Logarithmic Standard Deviation (β)
Calculation Method Young‟s Modulus Slope of Lateral Curve Poisson‟s
(GPa) (GPa) Ratio
Tangent Modulus 0.115 0.113 0.001
Secant Modulus 0.155 0.158 0.003
Average Modulus 0.106 0.125 0.019
128

120

100

80
Stress (MPa)
60
R-1-Axial
40 R-1-Circ
R-2-Axial
20
R-2-Circ

0
-0.002 -0.001 0 0.001 0.002 0.003 0.004 0.005 0.006
Strain

Fig. W-2U: Stress-Strain curve for rosette 1 (R-1) and rosette 2 (R-2)

Sample W-5U

Table W-5U-1: Results from the rosettes based on three Young‟s modulus calculation methods.
Young‟s Modulus Young‟s Modulus Slope of Lateral Poisson‟s
Calculation Method (GPa) Curve (GPa) Ratio
Rosette 1 17.14 86.38 0.20
Tangent modulus Rosette 2 18.63 115.10 0.16
Average 17.88 100.74 0.18
Strain
Average 17.85 98.69 0.18
Rosette 1 16.70 78.15 0.21
Secant Modulus Rosette 2 17.39 100.41 0.17
Average 17.04 89.28 0.19
Strain
Average 17.04 87.89 0.19
Rosette 1 16.93 83.81 0.20
Average Modulus Rosette 2 17.87 105.26 0.17
Average 17.40 94.53 0.18
Strain
Average 17.39 93.32 0.19

Table W-5U-2: Error calculation for three Young‟s modulus calculation methods.
Young‟s Modulus Logarithmic Standard Deviation (β)
Calculation Method Young‟s Modulus Slope of Lateral Curve
(GPa) (GPa) Poisson‟s Ratio
Tangent Modulus 0.034 0.118 0.084
Secant Modulus 0.016 0.103 0.091
Average Modulus 0.022 0.093 0.072
129

45
40
35

Stress (MPa) 30
25
20
R-1-Axial
15 R-1-Circ
10 R-2-Circ
5 R-2-Axial
0
-0.001 -0.0005 0 0.0005 0.001 0.0015 0.002 0.0025
Strain
Fig. W-5U: Stress-Strain curve for rosette 1 (R-1) and rosette 2 (R-2)

Sample W-6U

Table W-6U-1: Results from the rosettes based on three Young‟s modulus calculation methods.
Young‟s Modulus Young‟s Modulus Slope of Lateral Poisson‟s
Calculation Method (GPa) Curve (GPa) Ratio
Rosette 1 16.99 79.89 0.21
Tangent modulus Rosette 2 17.66 102.54 0.17
Average 17.33 91.22 0.19
Strain Average 17.32 89.81 0.19
Rosette 1 18.87 106.54 0.18
Secant Modulus Rosette 2 19.27 129.81 0.15
Average 19.07 118.18 0.16
Strain Average 19.07 117.03 0.16
Rosette 1 17.36 87.50 0.20
Average Modulus Rosette 2 18.87 116.59 0.16
Average 18.12 102.04 0.18
Strain Average 18.08 99.97 0.18

Table W-6U-2: Error calculation for three Young‟s modulus calculation methods.
Young‟s Modulus Logarithmic Standard Deviation (β)
Calculation Method Young‟s Modulus Slope of Lateral Curve
(GPa) (GPa) Poisson‟s Ratio
Tangent Modulus 0.034 0.132 0.102
Secant Modulus 0.113 0.304 0.158
Average Modulus 0.053 0.143 0.093
130

3-Comparison between the recorded displacement from MTS machine and


Strain-Gages:
Comparison between the recorded displacement/strain from the MTS machine and the

strain gages shows that the displacement recorded with the MTS machine is much larger

than the actual displacement from the strain gages. In some cases, the measured strain

with the MTS machine is three times more than the measured strain from the strain gages.

One of the causes of this difference is the deflections of the steel plates above and below

the core samples (Fig A-2-23). Another cause is the deformation of the load cell.

Nevertheless, it is important to find out if there is any relation between these two

displacements since, in this project, two of the prepared core samples were not tested

with the strain gages. Moreover, in the future projects, the strain gages may not be pasted

on the core sample because of any sort of consideration. However, the best alternative to

calculate the actual strain is to use the strain gage for all the UCS tests.

Fig. A-2-23: Steel platens above and below the core sample (MTS machine)
131

To find out if there is any (mathematical) relation between the recorded displacement on

the MTS machine and strain gages, the maximum, minimum, and average recorded strain

from the gages are plotted against the maximum, minimum, and average recorded strain

with the MTS machine. Since MTS machine and machine for the strain gages do not

record the data at a same time, it was not possible to plot all the available data. It might

be important to eliminate the time which may cause some sort of uncertainty. Therefore,

the calculations are based on the minimum data possible since the comparison is between

two different recording devices.

Table A-2-13 shows the slope of plotted curves for the strain gages strain against

recorded strain from the MTS machine. A correction factor for the deformation recorded

with MTS is calculated based on average of the slopes of the each figure (for each core

sample) and the slope of linear curve of all the considered displacement which is equal to

2.25 (Fig. A-2-25).

Fig. A-2-24: Displacement data was recorded with both MTS machine and computer.
132

Table A-2-12: Maximum, minimum, and average strain values recorded with MTS machine and
strain gages.
Sample ID Strain MTS machine Strain Gage
Secant Modulus 0.010752146 0.00425293
6U Tangent -8.92167E-06 -1.51776E-06
Modulus
Average 0.005819994 0.001445321
Modulus
Secant Modulus 0.01114397 0.005099604
8U Tangent
Modulus 1.275E-05 -1.07565E-06
Average
Modulus 0.005907099 0.001618576
Secant Modulus 0.006218328 0.002258719
W-5U Tangent
Modulus 0.000123126 -1.41488E-06
Average
Modulus 0.003545903 0.000667455
Secant Modulus 0.005800093 0.002258699
W-6U Tangent
Modulus -2.325E-05 -1.43534E-06
Average
Modulus 0.002983407 0.000667434
Table A-2-13: The slope of the strain gages strain-MTS machine strain curves
Sample ID Slope
6U 2.4242
8U 2.0745
W-5U 2.5122
W-6U 2.4302
Average=2.36
133

0.014

0.012

0.01
MTS Strain
0.008

0.006

0.004
MTS Strain = 2.25 (strain Gauges Strain) + 0.0009
0.002 R² = 0.94

0
-0.001 0 0.001 0.002 0.003 0.004 0.005 0.006
-0.002
strain gages strain

Fig A-2-25: Calculated J correction factor based on the Table 1 data

To evaluate which of the above slopes (2.36 as average of the slope of each curve for

each of the samples and 2.25 as slope of curve for all the considered data), the Young‟s

modulus of above samples were calculated using both numbers. The results demonstrate

that 2.25 is the better correction factor for the recorded strain with the MTS machine.

Moreover, this comparison confirms that “Average Modulus of Linear Portion of Axial

Stress-Strain Curve” method is the best method to find out the Young‟s modulus and

Poisson‟s ratio of the intact rock (Table A-2-14).


134

Table A-2-14: Comparison between the calculated Young‟s modulus from MTS machine and
strain gages using correction factor J=2.25.
Sample Calculation MTS Machine Strain Gage Young’s β
ID Method Young’s modulus modulus (GPa)
(GPa)
Maximum 43.3 41.8 0.04
6U Minimum 32.7 42.3 0.26
Average 43.9 43.1 0.02
Maximum 50.2 47.6 0.05
8U Minimum 47.5 48.7 0.02
Average 50.1 48.9 0.04
Maximum 17.4 17 0.02
W-5U Minimum 21.3 17.9 0.18
Average 19.2 17.4 0.1
Maximum 13.5 17.3 0.25
W-6U Minimum 14.6 19.1 0.26
Average 16.7 18.1 0.09

To validate the proposed correction factor “J”, aluminum cylindrical sample was tested

under compression with the MTS machine. The maximum force on the aluminum core

was 95 kips (less than 1 percent of strain). Table 4 shows the results of the aluminum

UCS test. The result indicates that the correction factor is fairly acceptable to find the

accurate deformation from the MTS machine.

Table 4. Validation of correction factor “J” based on Aluminum UCA test


Young‟s Modulus of Calculated Young‟s β
Aluminum Modulus
69 GPa 61.3 GPa 0.07

You might also like