You are on page 1of 10

1 INTRODUCTION

1.1 Back to Basics


The basic procedures for rock testing have been
around for decades. Since the mass marketing of
computers in the late 1970s, analytical methods such
as numerical modelling have been growing in
complexity and designers have been advancing the
theories and methods of calculations now possible
with powerful computers. Great effort has been
placed on improving measurement methods such as
through stress cells, extensometers, LiDAR, borehole
cameras, seismic systems, etc. The Rocscience
(https://www.rocscience.com) programs are an
excellent example of computer programs commonly
used to manipulate the basic rock data retrieved for
mine design. This paper supports these initiatives, but
serves to remind that basic experimentation is still the
root of rock engineering design. The lab experiments
for obtaining basic rock properties are well written in
literature, but questions by clients highlight that there
is a gap in knowledge between those requesting the
information and those performing the testing. This
paper helps to guide those requiring rock properties
in how to communicate with their rock mechanics lab
and possibly, their design specialists.
1.2 The Specialists
Lab testing of rock is generally requested directly by
a mining operation, or indirectly by a consultant such
as a numerical modeller that needs rock properties for
calibration purposes. Regardless, it is generally mine
personnel that assemble the samples for the facility
that will perform the lab testing. This paper is aimed
at helping those middle people.
It is very important that industry accepted testing
protocols are explicitly followed by experienced tech-
nicians/technologists. If there is one thing for certain
about lab testing of rock, it is that rock is never pre-
dictable. Proper protocols must be adhered to so that
the variability in results can at least be related to the
rock itself and not to non-standard methodologies
may have affected the results. Goodman (1989) is an
excellent reference that provides background infor-
mation on the fundamentals of lab testing for rock
mechanics.
1.3 Choosing Testing Methodologies
The most common procedures for consideration in
North America are the American Society for Testing
and Materials International (ASTM) and the Interna-
tional Society for Rock Mechanics (ISRM).
The ASTM standards are developed by consensus
and define the way that tests are performed, and de-
scribe the precision of results. These are developed to
represent the ASTM members, producers, users, con-
sumers, government and academia (ASTM, 2014).
The ISRM standards (ISRM, 2007) are similarly de-
veloped by consensus. However, these are recom-
mended procedures and not standards (ISRM, 2014).
There can be slight differences in methodologies be-
tween these organizations which is why it is so im-
portant to reference the actual one used. For instance,
ASTM and ISRM test specimens for UCS have
Lab Testing for Empirical Design
D. Beneteau, D. Milne and Z. Szczepanik
University of Saskatchewan







ABSTRACT: Ideally data for rock mass characterization is obtained through measurement. In all but weak
rock masses and high stress conditions, joint strength and spacing may be more important to the response of the
rock mass to mining than basic intact properties of the rock. However, laboratory testing should be performed
in all cases as it serves as one element of an overall rock mass description. This paper is a practical guide for
mining personnel for requesting laboratory investigations of rock. It addresses questions often received by in-
dustry regarding typical testing options such as the number of tests required, recognizing that properties of
intact rock are often highly variable. It also provides suggestions regarding sample selection. A variety of both
destructive and non-destructive test options are considered as rock testing programs should concentrate on the
properties which most influence the site specific requirements of any given operation.
height-to-diameter ratios of 2.0 to 2.5, and 2.5 to 3.0,
respectively. As both ASTM and ISRM procedures
are well accepted, it is most important to be consistent
over time in choosing procedures so that data can be
properly compared.
For any tests being performed, it is advised to fol-
low an existing method, if available. It is then possi-
ble to back compare results to published information.
However, there are cases when new procedures must
be devised. If new testing is successful, it is beneficial
to industry to allow for publication of the work so that
duplication of effort is avoided. In publishing that
procedure, be very specific to suppliers of compo-
nents and dimensions of equipment as this can save
the people trying to reproduce the work hours of frus-
tration. Most recently, we spent hours trying to find a
supplier of a 50 mm diamond saw blade in order to
conduct fracture toughness testing.
1.4 Selecting the Testing Facility
Different consulting companies and universities pro-
vide rock testing services in Canada. It is common for
perspective clients to request a lab that is ISO certi-
fied as this provides confidence in quality assurance.
ISO certification is a very onerous process for labor-
atories, especially university laboratories where there
may be insufficient staff and budget to perform the
work required of certification. The reality is that not
all labs with acceptable test equipment and qualified
staff are certified.
Furthermore, clients may question relying on test-
ing done by graduate students. Students can learn
from witnessing/performing testing with various rock
types rather than relying on data provided by external
sources. Testing results, however, can suffer if expe-
rienced personnel are not conducting the work. The
following are typical questions that can be asked by
the client to gain confidence in the quality of testing
results by any lab:
1. Is the work being performed or directly super-
vised by an experienced technician/technolo-
gist specializing in rock mechanics?
2. Is the lab adhering to ASTM/ISRM rock stand-
ards, and if so, which ones?
3. What type of equipment is used to ensure the
ends of samples are properly prepared?
4. Can the lab provide calibration curves for
equipment?
5. Will the lab allow the client to watch testing
being done?
Of those questions, ensuring the work is performed
by, or under the direct supervision of a qualified per-
sonnel is extremely important. Due to variability in
rock types and the complexity of equipment, it can
take many years to master rock mechanics testing.
Even after many years of experience there are often
new surprises to deal with. It may be something as
seemingly insignificant as a small change in chemis-
try that will affect the transmission of waves through
a sample.
1.5 Variability between Laboratories
It is reasonable, though not desired, to expect differ-
ent results between labs. One way to remove this un-
certainty is to make use of only one lab, but this may
not be in the clients best interests. Therefore, it is im-
portant to be aware that variability can arise even
while adhering to the standards.
Pincus (1993) reported on an interlaboratory test-
ing program using 8 different laboratories. Testing
was performed on granite, sandstone, limestone and
marble. For the four different ASTM procedures that
were used (unconfined compressive strength (UCS),
velocities, tensile strength, and elastic modulus from
UCS) , it was shown that samples sent to these differ-
ent laboratories showed variable results. There are
many different reasons for variability between labor-
atories including:
1. End effects. In a paper by Szczepanik et al.
(2007), end preparation alone caused variabil-
ity in strength by as much as 50%. The stand-
ards only specify that ends are to be parallel,
with a maximum degree of roughness (surface
flatness). They do not state the actual equip-
ment to be used to prepare the ends. Many op-
tions exist from grinding to polishing and
these all meet end preparation requirements of
perpendicularity to sides and end surface flat-
ness. Testing has shown significant variability
in strength, even when following the stand-
ards.
2. The samples themselves. Two samples ob-
tained side-by-side from a core box can vary
significantly in strength just due to microfrac-
tures and rock fabric.
3. Loading rates. The ASTM and ISRM proce-
dures only state that the UCS samples must be
broken in a certain time frame. For example,
in 2 to 15 minutes for ASTM. Testing, such as
that done by Peng (1975) has shown that rock
will break at a higher compressive strength
when loaded at higher strain rates. Even an ex-
perienced technician can be surprised at how
weak/strong different core samples are so
even setting a loading rate relies on experi-
ence.
1.6 Rock or Soils Laboratory
In the context of this document, a rock mechanics la-
boratory is equipped to test solid rock specimens. One
notable difference between these two materials is that
the strength of soils is generally in KPas while that
of rock is in MPas. While it is very important that the
natural moisture content of soils be preserved for test-
ing as effective stress is highly dependent upon the
water content, most rock mechanics tests for mining
are done once rock is air dry.
Difficulty arises for highly weathered rock for
which strength properties are required. For those
rocks, strength may be highly dependent upon the
moisture content so the client must preserve the core.
End preparation for UCS testing may not be feasible
in some cases as the samples will crumble during
sawing. Or if cutting is successful, it may not be pos-
sible to grind the ends to be perfectly parallel or apply
an end cap. With these considerations in mind, some
deviations from proper ASTM/ISRM procedures may
occur merely due to the nature of the samples. All de-
viations should be noted with results.
1.7 Program Design
Clients do occasionally ask for recommendations on
test program design. They may have been advised by
their numerical modelling consultant to provide rock
properties and would like guidance on what they
should do. The main objective of the testing is to get
the properties required by those specialists so they
should be included in the discussion. But rock me-
chanics testing is a specialty too and there may be
times when all parties should be brought into the dis-
cussion.
Each testing program will have its own require-
ments. We typically see programs that include uniax-
ial, tensile strength and occasionally triaxial testing.
On occasion, clients may request shear testing. Non-
destructive rock characterization tests are also rou-
tinely performed such as density and velocity.
1.8 Number of Tests
It is very common for clients to ask for three repeti-
tions for each rock type/test type. This could have
evolved from concrete testing, in which it is common
to test 3 cylinders at each cure time. ASTM and
ISRM actually recommend many more tests, typically
as many as 10 samples for each rock/test type. The
reality is that for routine rock testing, clients typically
do not follow the recommendations of the standards.
The quantity of testing is actually dictated by the
clients budget and design requirements. A lab should
respect the request of the client as their duty is not to
analyze the data but to provide quality results. Also,
it may not be physically possible to get a large num-
ber of samples due to the availability of the rock being
tested. Samples may be from a narrow seam or the
RQD may be so poor that only a few intact samples
can actually be selected. Gill et al. (2005) emphasize
that the precision index, testing procedures, test se-
lection, sampling locations, specimen preparation are
up to the engineers judgement and those authors do
provide guidance on sampling theories from the per-
spective of the requirements of the standards. It is un-
realistic in rock testing to have a constraint on the
number of rock samples. However, the paper by Ruf-
folo and Shakoor (2009) does support that 10 core
samples is the minimum number required for estimat-
ing the mean unconfined compressive strength.
Regardless of the number, we recommend that the
clients aim to get the greatest visual homogeneity be-
tween samples of a given group. Clients can opt to do
non- destructive testing of those samples as a further
indication of homogeneity, as described in Section 3.
According to Hoek in Practical Rock Engineering
(2007), 10 percent of a well-balanced rock mechan-
ics program should be allocated to laboratory test-
ing. From the laboratories point of view, we do not
want our clients to be upset when they get incon-
sistent results. However, the client should always
consider the sensitivity of the design to the estimated
lab test data. The unconfined compressive strength of
the rock is a key input in many rock mass classifica-
tion systems, but some of the strength categories have
a range of 100 MPa. For this application, inconsistent
lab results may not be a concern. Triaxial testing for
numerical modelling is more sensitive to sample var-
iability. A failure envelope for Hoek-Brown failure
criteria can be estimated from 3 triaxial tests, how-
ever, the results will be very sensitive to rock varia-
bility.
Rock can have hugely variability properties be-
tween specimens, while adhering to given proce-
dures. In a recent test, we had two samples that were
cut side by side and had a similar appearance. The
UCS results were 27.1 MPa and 92.6 MPa versus an-
other set of side-by-side cores from the same ship-
ment that were 84.5 MPa and 116.8 MPa. We do not
reject results due to scatter. We encourage clients to
look at before/after photographs of samples so that
they can get clues as to why there may be such devi-
ation.
1.9 Core Diameter
Literature in general tends to support a 50mm core
diameter as the standard testing size. As core size is
often dictated by geology and drill availability, the
engineering department may not have any control
over this. It is important to document core diameter
with results. We have tested rock as small as 1 diam-
eter, to several inches in diameter, while adhering to
the general principles of sample preparation of the
specified test procedures.
A testing facility will likely have the capability of
preparing cores from blocks of rock. In this case, the
client should consult with the lab in regards to a prac-
tical size block that the lab can handle.
What is most important in sample dimensions is to
ask your laboratory the capacity of their load frames
and triaxial cells prior to shipping samples. As core
size increases, so does the effective load required for
failure. Especially under confinement in triaxial test-
ing, a very strong rock may not be able to be broken
if the core diameter is too large. Also with respect to
triaxial testing, there will be a limitation for sample
dimensions based on the dimensions of the cell itself.
Sometimes oversized samples can be cored down us-
ing a drill press in the lab, but not always. The client
should always be aware that added sample disturb-
ance will result from re-coring samples to a smaller
size.
1.10 Core Selection
Ideally, all samples for a similar rock type are col-
lected side by side so that the composition and stress
environment of groupings of samples are as close as
possible. But the reality of needing to preserve and/or
assay core, and of sampling from thin seams and low
RQD dictates that core of similar rock type may be
sampled from across many holes. This can be one ex-
planation for variable results. This variability should
not be treated as sampling or testing error, it is provid-
ing further information about the properties of the
rock.
Ideally the core for a given test program should also
be of comparable diameter. But it is also common that
multiple core sizes are drilled in a given mine. At a
minimum, all triaxial testing should be on similar size
core if possible.
It is preferable if the client provides an Excel
spreadsheet in which each piece of core is identified,
along with the test required. The nomenclature used
for naming core can often be unusual and difficult to
decipher when written with markers on samples.
There is less chance for error if this is done in advance
by the client. In addition, a geologist should ensure
that groups of similar rock type are identified
properly.
If the rock is weak and may deteriorate with expo-
sure to air and water evaporation, it should be indi-
vidually encased in wax or sealed in tight plastic bags.
All core should be packed carefully into core boxes,
plastic pails or cardboard boxes in such a way that the
core will not slide. Each core piece should be individ-
ually labeled and listed in a table of samples that is
printed. If the core is silty, the label may wear off with
time so those samples should also be placed in la-
belled plastic bags.
Core lengths should be sent at least 3 times the di-
ameter for each triaxial and UCS test so that the final
specimen is sufficiently long once trimmed. It is ad-
vised to send additional core as pieces do arrive that
have split on healed fractures during shipment. In ad-
dition, sometimes samples breaks on fractures during
sample preparation. Those samples must be discarded
and highlight how sample selection alone biases the
results of laboratory testing.
It is recognized that, especially in weak rock, the
samples selected are the samples that have survived
drilling, transport and possibly storage. The bias to-
wards stronger rock is recognized, and is built into
our empirical design methods.


2 STRENGTH TESTING RESULTS
2.1 Presenting results
In acknowledgement of the variation between
strengths of samples, we recommend a simplified
graphical presentation of results as shown in Figure
1. Simple charts as shown can be created in Excel by
inserting High-Low-Close stock chart. This partic-
ular chart illustrates UCS results for dolomite samples
tested for an undergraduate rock mechanics class.
Over the years, different core boxes of dolomite from
a drill program in Saskatchewan have been used in
teaching. A qualified technologist has prepared sam-
ples and demonstrated UCS testing to groups of 3 or
4 students.
In looking at this chart, it can quickly be observed
that there were 5 samples tested in 2012, 4 in 2013,
and 9 in 2014. The standard deviation was variable
between years, and greatest in 2014. With the 18 sam-
ples combined, the standard deviation was similar to
that of 2013 and overall, the average strength across
the years was consistent.
Depending upon the rock mass classification for
which strength testing is being performed, the
strength categories for the classification system can
be overlain on this chart. Figure 1 illustrates the seven
ISRM (2007) R categories. Each classification sys-
tem provides a different rating for the strength of in-
tact rock material. The Bieniawski (1989) rock mass
rating (RMR) provides a strength rating that is be-
tween 0 and 15 for the allowable 100 points required.
This uses the seven ISRM categories. Another com-
mon system for mining, the Q system (Barton et al.,
1974) indirectly incorporates strength through the
stress reduction factor (SRF).
In addition to plotting results,it is also important to
note the type of failure associated with each speci-
men. This could possibly be used to explain the vari-
ability in some of the results. Typical failures are
identified in Figure 2.
2.2 UCS Testing
Figure 3 shows a graphical representations of the
UCS data by Pincus (1993). Each point represents
five replications of a rock type by each of seven dif-
ferent labs. ASTM procedures for UCS were fol-
lowed.
These five rock types were selected to represent a
range of values of rock properties, and were of uni-
form composition. The granite and marble are mostly
in the ISRM R5 strength category and the Sandsotone
and Limesone are generally R3 (Figure 3). Typical
rock from mining operations may not be as homoge-
neous as these rock types.
These simplistic plots in Figure 3 highlight the uni-
formity of these rocks versus the limestone in Figure
1 through narrower standard deviations. As per the
comment in Section 1.5, they also confirm that varia-
bility exists between laboratories.

Figure 1. This square symbols on this chart represent the mean
strength in MPa of dolomite rock tests, and the vertical lines rep-
resent one standard deviation of mean values determined from 5
to 9 samples per group. ISRM rock classifications overlain on
Figure 1 as >250 MPa (R6:Extremely strong), 100-250 MPa
(R5:Very strong), 50-100 MPa (R4:Strong), 25-50 (R3: Medium
Strong), 5-25 (R2:Weak), 1-5 MPa (R1:Weak) and 0.25 1 MPa
(R0: 0.25-1 MPa).













Figure 2. Types of failures that can be observed in UCS testing
include (a) failure along features (b) shear failure (c) vertical ten-
sile failure (c) complete destruction

2.3 Youngs Modulus
Clients have questioned if they should collect
Youngs modulus from UCS and/or triaxial testing.
The Pincus Round 2 interlaboratory study (1994)
found that E increases with confining pressure for
Barre Granite and Berea Sandstone, but decreases for
Tennessee Marble. Logically, the specimens should
become stiffer with confinement and have an increase
in Youngs modulus. However, the results from the
Tennessee Marble highlight that there is nothing pre-
dictable about lab test results from rock.
As shown in Figure 4, there is variability in
Youngs modulus between labs if collecting data
from UCS testing. This is as expected from the varia-
ble UCS results of Figure 3. From a budgeting point
of view, it is typically less expensive to get moduli
(both Youngs Modulus and Poissons ratio) from
UCS testing. Due to variability with all testing, the
results from UCS testing may be adequate in most in-
stances.
.














































Figure 3. The square symbols on this chart represent the mean
UCS in MPa of 4 rock types in an interlaboratory study (Pincus,
1993), and the vertical lines represent one standard deviation of
mean values determined from 5 samples per group.















































Figure 4. Elastic moduli results. The square symbols represent
the average modulus for a given lab in GPa (Pincus, 1993). The
lines represent one standard deviation.

























































Figure 5. Tensile strength results. The square symbols on this
chart represent the mean tensile strength in MPa of 4 rock types
in an interlaboratory study (Pincus, 1993), and the vertical lines
represent one standard deviation of mean values determined
from 5 samples per group.




2.4 Tensile Strength Testing (Brazilian)
The other most commonly requested test from indus-
try is tensile strength or Brazilian. This is also one of
the least expensive tests to perform. Pincus (1993)
data is plotted in Figure 5. The standard deviation is
up to 1.9 MPa which is less than that for their UCS
testing. The ratio average UCS to average Brazilian
strength for each of these labs is as follows:
14.7 for Tennessee Marble
11.8 for Salem Limestone
17.1 for Barre Granite
15.7 for Berea Sandstone
These results highlight that a rock is weaker in ten-
sion than compression and that the ratio is dependent
on rock type
2.5 Shear Strength
Some projects may require shear test results. Figure 6
presents results from tensile, shear and UCS testing
of potash (Neely, 2014). The results are plotted on a
Mohr-Coulomb diagram to illustrate the comparison
between the three tests. The average Brazilian
strength was 1.75 MPa, shear strength was 7 MPa and
UCS was 25.4 MPa. Standard deviations were 0.2
MPa, 1.8 MPa and 2.4 MPa respectively. For this pot-
ash, the ratio of average UCS to average Brazilian
was 14.5. A trend line was sketched in by eye as an
approximation of a failure envelope.
When asked if UCS and triaxial tests should be in-
cluded on the same plot, all we can suggest is that the
client tests if a failure envelope can be fit to the data.
Unfortunately, there were no triaxial results with this
dataset.


Figure 6. UCS, tensile and shear strength results from Potash
testing.
2.6 Triaxial Data
The purpose of this paper is not to interpret lab data
through failure criteria but instead to give practical
guidance on approaching lab testing. The standard
Mohr-Coulomb failure criterion is used to demon-
strate the types of results that may be obtained with
only 3 samples, as follows:
= c + tan
Where,
= shear stressss acting on failure plane (MPa)
= normal stress acting on failure plane (MPa)
c = cohesive strength of rock (MPa)
= angle of internal friction
There are other failure criteria commonly used in
mining such as the Hoek and Brown failure criteria
(1980), but for simplicity and a lack of space, only the
Mohr-Coulomb was selected.
When it comes to triaxial data, the more samples to
create the curves the better. Figure 7 shows three ex-
amples of triaxial results using only three samples.
The goal was to sketch in a linear envelope to the cir-
cles, by eye. The upper chart can easily be represented
by a linear failure curve and as expected, strength in-
creases with increasing confinement. With the middle
curve, an approximate line of best fit can be estimated
but results are not as ideal. Finally, the bottom curve
presents results that are opposite to that expected. The
rock cores actually gave lower failure strengths as
confinement increased. Variability in sample
strengths was though to be the cause.
When selecting intact rock samples, discontinuities
and bedding can affect material strength and behav-
ior. Three samples may not be sufficient for determin-
ing failure criteria parameters for rock mass charac-
terization. We suggest that a minimum of five
samples may improve the interpretation of results.
The client needs to know the application for this data
to help determine the sensitivity of the design to an
accurate estimate of the lab test values.


3 NON-DESTRUCTIVE TESTING
3.1 The purpose
Certain borehole logging tests can be simulated on
rock core without causing damage to the cores. This
is typically done on prepared UCS and triaxial
cores.The advantage of non-destructive testing is that
it provides additional properties to help identify the
differences between similar looking rock cores. These
non-destructive rock characteristics can be compared
to results from destructive test properties.
Depending on how sampling was done, software is
available to plot the results similar to how data is plot-
ted in borehole logs. In Figure 8, density and porosity
are plotted alongside representative photographs of
the cores. Similar plots could be added to this figure
including those from other destructive tests including
UCS, Youngs modulus, Poissons ratio, tensile
strength, shear strength, etc.












































Figure 7. Examples of triaxial results using three samples. Top
chart shows ideal Mohr Coulomb relation, middle requires some
estimating, and bottom is the reverse of that expected.

3.2 Density
The most basic non-destructive test for characterizing
rock is density. If porosities and/or natural water con-
tent are required, it is important to note the test
method used. The porosity can be done by the water
saturation method, or using porosimeters that inject
gas into the samples. Each method may produce dif-
ferent results so it is important to know how the tests
were performed.
3.3 P-wave and S-wave Velocities
To measure P-wave and S-wave velocity, ultrasonic
pulses are passed through rock core samples. The ve-
locity that those pulses travel in the core depends on
the density and elastic properties of the rock. The na-
ture of the rock specimen determines how well the
technique can be applied. As the sample becomes
more heterogeneous due to fractures, large grain size,
increased porosity, etc., this technique becomes more
difficult to apply due to wave attenuation. However
in general, the technique is fairly reliable and P-wave
velocity at a minimum can usually be determined.
The frequency range of the ultrasonic vibrations
used in laboratory testing is dependent upon rock
type. Typical range is 200 kHz to 1 MHz. Again this
test highlights why it is important to have qualified
personnel running the test. Sensors are selected based
on sample size and sample composition. It is not just
a matter of turning on an instrument and running the
test.



Figure 8. Example of plotting laboratory data and photos on a
depth log for a borehole. In this case, average results were plot-
ted when multiple sub-core results were available at the same
depth (Hawkes et al, 2013).


Sample dimensions are important to note as this de-
termines the distance which the pulses travel as pulse
velocity is calculated by dividing path length by
transit time. Ideally this test is done on saturated sam-
ples as the water content can increase pulse velocity.
By having all samples saturated, at least this variable
is removed. Again, this should be noted in test rec-
ords.
This test is a good example of the need to keep up
with evolving equipment. Recently, our laboratory
purchased a new oscilloscope at a cost of approxi-
mately $8,000. This is just one small piece of equip-
ment in our laboratory. A rock mechanics laboratory
can easily exceed the million dollar range for routine
equipment, and equipment should be updated as com-
puter technology evolves. Not only is it important to
have a skilled technologist who is qualified to do the
testing, but also it is important that they are techni-
cally competent to operate and maintain an inventory
of continually upgraded equipment.
3.4 Resistivity
It is sometimes possible to perform geophysical tests
at the core scale. One such test routinely done at the
University of Saskatchewan is resistivity. Figure 10
shows the absolute resistivity measurements made on
a two sections of a core cut side by side. Sample A
was 25 mm long and Sample B was 77 mm long. Both
samples were oven dried, and then saturated together
in distilled water.
This example highlights the variability in results
between adjacent samples of core. The absolute val-
ues of resistivity of these two sub-segments are very
different with Samples A and B having resistivities of
250,000 m and 350,000 m at 0.1 Hz. Fortunately
when these results are normalized as shown in Figure
11, the results are almost identical.
ASTM/ISRM procedures are not available for geo-
physical properties. When required, the client will
need to find a qualified lab.


4 CONCLUSIONS

Laboratory testing is one component of a program to
characterize a rock mass. This must be done by a spe-
cialized testing facility with proper equipment.
Though the basics behind the testing seem simple, the
actual mechanics of carrying our rock mechanics test-
ing is not. It is important that this testing is done by
qualified personnel, that test procedures are strictly
followed, and that test methods are noted with results.
Rock results are usually highly variable and test
design should aim for a number of samples that will
meet the requirements of the project. There may be
requirements for specialized testing that may not be
available from ASTM/ISRM and these should be
documented and ideally published so that methods
can be improved and standardized with time.
Rock test results at the laboratory scale are af-
fected by discontinuities and grain structure as in the
field, sometimes even when samples appear homog-
enous. A rock will fail the way it wants so a saying
in our lab it that it is what it is. RMR allows for 15
of 100 points to be obtained from laboratory testing.
As much care should be placed into that 15% as is
placed in the remaining 85% collected from field
mapping.






Figure 9. Resistivity results for two adjacent samples. Top chart
shows absolute results and bottom chart shows normalized re-
sults.


5 REFERENCES

Rocscience, 2014. https://www.rocscience.com (Last accessed
April 8, 2014)
ISRM, 2014. International Society for Rock Mechanics,
http://www.isrm.net (Last accessed April 8, 2014)
ASTM International, 2014. http://www.astm.org (Last accessed
April 8, 2014)
Goodman, R.E., 1989. Introduction to Rock Mechanics Second
Edition, John Wiley & Sons, 562 p.
ISRM, 2007. The complete ISRM suggested methods for rock
characterizations, testing and monitoring: 1974-2006. Edi-
tors: R. Ulusay&J.A.Hudson, 628 p.
Pincus, H.J., 1993. Interlaboratory testing program for rock
properties (ITP/RP), Round 1 longitudinal and transverse
pulse velocities, unconfined compressive strength, uniaxial
elastic modulus and splitting tensile strength. Geotechnical
Testing Journal GTJODJ, Vol. 16. No. 1, March 1993, pp.
138-163.
Szczepanik, Z., Milne, D. & Hawkes, C., 2007. The confining
effect of end roughness on unconfined compressive strength.
In proceedings of the 1
st
Canada-US Rock Mechanics Sym-
posium, Vancouver, Canada, pp. 191-198.
Peng, S.S., 1975. A note on the fracture propagation and time-
dependent behavior of rocks in uniaxial tension. Int. J. rock
Mech. Min. Sci. 12, 125-127
Gill, D. E., Corthesy, R.,& Leite, M. H., 2005. Determining the
miniumal number of specimens for laboratory testing of rock
properties. Engineering Geology 78, pp. 29-51.
Ruffolo, R.M. & Shakoor, A., 2009. Variability of unconfined
compressive strength in relation to number of test samples.
Engineering Geology 108 (2009) pp. 16-23.
Hoek, E., 2007. Practical Rock Engineering. Notes downloaded
from https://www.rocscience.com, Last accessed April 8,
2014, 237 p.
Bieniawski, Z.T., 1989. Engineering rock mass classifications,
New York: Wiley
Barton, N.R., Lien, R. and Lunde, J. 1974. Engineering classifi-
cation of rock masses for the sesign of tunnel support. Rock
Mechan. 6(4), pp. 189-239.
Pincus, H.J., 1994, Interlaboratory testing program for rock
properties (ITP/RP), Round 2 confined compression: elas-
tic modulus and ultimate strength, Institute for Standards Re-
search Report PCN: 33-000010-38, 80 p.
Neely, D., 2014, Failure mechanism of resin anchored rebar in
potash, MSc. Thesis Draft, Department of Civil and Geolog-
ical Engineering, University of Saskatchewan.
Hoek, E. and Brown, E.T., 1980. Underground excavations in
rock, the Institution of Mining and Metallurgy, 527 p.
Hawkes, C., Beneteau, D., and Szczepanik, Z., 2013. Geological
and geomechanical controls on hydraulic fracturing in the
Bakken formation Part 2-Geomechanics, PTRC Report No.
L0-UOS-00001-2011, Petrolieum Technology Research
Centre STEPS Research Program, 49 p., December 2013

You might also like