Vous êtes sur la page 1sur 17

Rock Mech Rock Eng (2011) 44:513529

DOI 10.1007/s00603-011-0161-6

ORIGINAL PAPER

The Effect of Specimen Size on Strength and Other Properties


in Laboratory Testing of Rock and Rock-Like Cementitious
Brittle Materials
William J. Darlington Pathegama G. Ranjith
S. K. Choi

Received: 24 January 2011 / Accepted: 27 May 2011 / Published online: 17 June 2011
Springer-Verlag 2011

Abstract The effect of specimen size on the measured


unconfined compressive strength and other mechanical
properties has been studied by numerous researchers in the
past, although much of this work has been based on
specimens of non-standard dimensions and shapes, and
over a limited size range. A review of the published literature was completed concentrating on the presentation of
research pertaining to right cylindrical specimens with
height:diameter ratios of 2:1. Additionally, new data has
been presented considering high strength (70 MPa) cement
mortar specimens of various diameters ranging from 63 to
300 mm which were tested to failure. Currently, several
models exist in the published literature that seek to predict
the strengthsize relationship in rock or cementitious
materials. Modelling the reviewed datasets, statistical
analysis was used to help establish which of these models
best represents the empirical evidence. The findings presented here suggest that over the range of specimen sizes
explored, the MFSL (Carpinteri et al. in Mater Struct
28:311317, 1995) model most closely predicts the
strengthsize relationship in rock and cementitious materials, and that a majority of the empirical evidence supports
an asymptotic value in strength at large specimen

W. J. Darlington (&)  P. G. Ranjith


Department of Civil Engineering, Monash University,
Clayton, VIC, Australia
e-mail: William.Darlington@gmail.com
P. G. Ranjith
e-mail: ranjith.pg@monash.edu
S. K. Choi
CSIRO Earth Science and Resource Engineering,
Bayview Avenue, Clayton, VIC, Australia
e-mail: Xavier.Choi@csiro.au

diameters. Furthermore, the MFSL relationship is not only


able to model monotonically decreasing strengthsize
relationships but is also equally applicable to monotonically increasing relationships, which although shown to be
rare do for example exist in rocks with fractal distributions
of hard particles.
Keywords Specimen size  Compressive strength 
Rock  Rock-like materials  Scale effect  Size effect

1 Introduction
For most engineering design, material properties such as
strength, Youngs modulus and Poissons ratio are of
critical importance. Establishing these parameters often
proves problematic when considering materials such as
concrete, and more acutely rock, where the size of the
engineering structure far exceeds the size of any laboratory
test specimen.
A great deal of research has been focused on upscaling
laboratory measured strength parameters to field problems.
This has been done in a number of ways. Empirical studies
have concentrated on finding size-dependent relations
between laboratory measured properties using small size
laboratory specimens (e.g. 30150 mm diameter cylinders). To establish the properties of specimens larger than
this, and thus make sizestrength comparisons with laboratory measured values, in situ testing or back analysis of
large structures is often used.
The results of these are generally well known and
accepted. The most commonly cited; the general work of
Hoek and Brown (1980), Brace (1981) and the original
work of Weibull (1951) form either the basis or what is
directly used in many rock mechanics designs today.

123

514

Figure 1 shows the data compiled, and corresponding


relationship established by Hoek and Brown (1980)
between specimen diameter and the strength of intact rock.
This relation is represented by Eq. 1 where rcd is the
uniaxial compressive strength of a sample with diameter, d;
and rc50 is the uniaxial compressive strength of a 50 mm
diameter sample.
 0:18
50
rcd rc50
1
d
A similar relation has also been proposed by Cunha
(1990):
 0:22
50
rcd rc50
2
d
It is generally accepted that there is a significant
reduction in strength with increasing specimen size.
Equations 1 and 2 are the current benchmarks when
quantifying this phenomenon in rock. Hoek (2000)
suggests this reduction in strength is due to the increased
probability that failure of rock grains will occur as the
specimen size increases. Rock strength will reach an
asymptotic minimum value at a certain specimen size that
will depend on the type and condition of the rock. Hoek
(2000) goes further, hypothesising that the strength of a
rock mass will reach a constant minimum value if the size
of the rock blocks are considerably smaller than the rock
mass under consideration. However, importantly the power
law proposed by Hoek and Brown (1980) includes a
horizontal asymptote of zero as diameter tends to infinity.

Fig. 1 Influence of specimen size on the strength of intact rock


(Hoek 2000)

123

W. J. Darlington et al.

It is important to note that the literature discussed and all


results presented in this paper are exclusively based on
cylindrical specimens with a height to diameter ratio of 2:1.
Although data exists on specimens of other ratios and
shapes, the results of these are outside the scope of this
paper. From a general perspective, with most, if not all,
strengths obtained for both rock and concrete design
obtained from specimens with height to diameter ratios of
2:1, and this being a widely accepted standard, a study
using samples with these geometric properties seems most
useful.
1.1 Cementitious Materials
Carpinteri et al. (1999) noted a significant lack of research
regarding the compressive strengthsize relationship of
laboratory size concrete specimens and explored the
applicability of the size effect law (SEL) (Eq. 3) proposed
by Bazant (1984) and the multifractal scaling law (MFSL)
(Eq. 4) proposed by Carpinteri et al. (1995) through comparisons with a variety of published data (Fig. 2). In relation to Eqs. 3 and 4, rN is the normal strength; d is the
specimen diameter; ft is a strength parameter; b and k0 are
two empirical constants; dmax is the maximum aggregate
size; fc and lch are constants that represent the nominal
compressive strength of an infinitely large specimen and an
internal material length, respectively; these are determined
by non-linear least squares fitting:
bft
rN p
3
1 d=k0 dmax 
r
lch
rN fc 1
4
d
Considering concrete, Carpinteri et al. (1999) noted
Eqs. 3 and 4 as the only published relations to describe the
strengthsize relationship. Furthermore, they showed that
SEL and MFSL generate opposite predictions. The SEL
predicts infinitely large specimens to have zero strength
similar to Eqs. 1 and 2, while the MFSL predicts large
specimens to have a finite asymptotic minimum strength.
Considering the MFSL further, it was found to underestimate large specimen strength by 10%, while the strength of
small specimens (where d 10dmax ) are overestimated by
10%.
Symons (1970) also assesses the scale effect in cement
stabilised materials. Here specimens were made using three
grades of aggregate (well-graded sand, crushed rock and
gravel-sand-clay) and varying cement content percentage.
Four specimens of each aggregate grade at each percentage
of cement were tested. Several height to depth ratios were
assessed in a range of cylindrical specimen diameters and
square prism sizes; the average results obtained for the
cylindrical specimens with a height to diameter ratio of 2:1

The Effect of Specimen Size on Strength and Other Properties

515

(a)

(b)
Fig. 3 Normalised UCS of well-graded sand-cement mortar correlated with specimen diameter (after Symons 1970). The size effect
relations of Hoek and Brown (1980) and Carpinteri et al. (1995) are
also plotted (the MFSL has only been fitted to the 14% well-graded
sand data in order to preserve clarity of the figure)

(c)

Fig. 4 Normalised UCS of cemented crushed rock correlated with


specimen diameter (after Symons 1970). The size effect relations of
Hoek and Brown (1980) and Carpinteri et al. (1995) are also plotted
(the MFSL has only been fitted to the 4% crushed rock data)
Fig. 2 Plots showing the data of Blanks and McNamara (1935)
for concrete of different water to cement ratios a w/c = 0.53;
b w/c = 0.55; and c w/c = 0.54), and the strengthsize relations of
Bazant (1984) (SEL), Carpinteri et al. (1995) (MFSL), and Hoek and
Brown (1980) (after Carpinteri et al. 1999)

1.04

4.1

1.02

1
3.9

are presented in Figs. 3 and 4. There appears to be no


clear or significant scale relationship for the well-graded
sand-cement mortar specimens, while the cemented crushed rock specimens show a decrease in strength with size
modelled accurately by both the Hoek and Brown and
MFSL relationships. Figure 5 shows the results produced
by Hoskins and Horino (1969) for Plaster of Paris. Considering only their data for specimens with diameters [50
mm the MFSL relationship models the experimental data
well, while the relationship of Hoek and Brown (1980)
significantly underestimates the strength of larger diameter specimens.

0.98

3.8

0.96

3.7

0.94
0.92

3.6

0.9
3.5

0.88
0.86

3.4
0

50

100

150

Fig. 5 UCS of Plaster of Paris in correlation with specimen diameter


(error bars indicate standard error of the mean) (after Hoskins and
Horino 1969). The size effect relations of Hoek and Brown (1980)
and Carpinteri et al. (1995) are also plotted for comparison. The
MFSL line has been fitted only to data [50 mm

123

516

Fig. 6 UCS of jointed yellow limestone correlated with specimen


diameter and line of best fit in the form of Eq. 1 with a power of 1.6
(instead of 0.18). The size effect relation of Carpinteri et al. (1995) is
also plotted for comparison. The UCS of a 50 mm specimen was
obtained for the purpose of normalisation using Eq. 1 (after Natau
et al. 1983)

Fig. 7 UCS of Hawkesbury Sandstone correlated with specimen


diameter (error bars indicate 1.0 standard deviation from the mean)
(after Pells 2004). The size effect relations of Hoek and Brown
(1980), and Carpinteri et al. (1995) are also plotted for each sandstone
outcrop. WPO West Pymble Outcrop, GQ Gosford Quarry

1.2 Sedimentary Rocks


Natau et al. (1983) demonstrated a trend that followed the
same form as Hoek and Browns (1980) model (Eq. 1), but
with a significantly different power (Fig. 6). However, it is
important to note that Fig. 6 depicts the results for jointed
yellow limestone with an unspecified joint spacing and
load orientation which may influence the magnitude of the
decrease in strength with specimen size.
Pells (2004) summarised the effect of specimen size on
the strength of Hawkesbury Sandstone (Fig. 7) by testing
specimens ranging from 18 to 144 mm in diameter. Using
Eq. 1, Pells (2004) expected the 150 mm specimens to
have a strength of around 85% of that of the 50 mm

123

W. J. Darlington et al.

Fig. 8 UCS of limestone in correlation with specimen diameter


(error bars indicate minimum and maximum values in data set) (after
Thuro et al. 2001). The size effect relations of Hoek and Brown
(1980) and Carpinteri et al. (1995) are also plotted for comparison

specimens. However, this was not found to be the case, and


no clear scale effect was seen in this rock. The conclusions
of this research do point out that almost all the rock types
used in the derivation of Hoek and Browns (1980) Eq. 1
were igneous or crystalline and therefore would inherently
contain micro-cracks. Following the theory of Weibull
(1951) these micro-cracks are the cause of a strength scale
effect. Although most sedimentary rocks do exhibit some
scale effect, Hawkesbury Sandstone appears to be an
exception rather than the rule. Figure 7 displays a monotonically increasing MFSL relationship for the WPO data
due to the increasing nature of the original dataset and the
fact that the MFSL is fitted to each dataset using non-linear
least squares fitting to find the correct values of the fitting
constants. It is hypothesised that the monotonically
increasing relationship seen here is due to the fractal distribution of weathering effects in this moderately weathered Sandstone. In modelling rock behaviour where there is
a monotonically increasing strengthsize relationship the
MFSL has the advantage of being able to model both
increasing and decreasing trends.
Thuro et al. (2001) tested two rock types under uniaxial
compression: a coarse-grained two-mica granite; and a
fine- to medium-grained clastic limestone. These results are
shown in Figs. 8 and 12, respectively. No dramatic scale
effect for either UCS or Youngs modulus is shown and the
MFSL relationship plotted in Fig. 8 monotonically
increases due to the increasing nature of the original dataset
being modelled. As the authors point out, this contradicts
what had been previously published by Hoek and Brown
(1980) (Fig. 1), and Hawkins (1998) (Fig. 11).
Hawkins (1998) critically reviews the commonly
accepted relation proposed by Hoek and Brown (1980)
(Eq. 1). In doing so, he presents new data pertaining to the
strengthsize relationship of a range of sedimentary rocks.
Hawkins results are plotted in Fig. 11. They show that
over a sample of specimen diameters, peak strength will be

The Effect of Specimen Size on Strength and Other Properties

Fig. 9 UCS of Longmont Sandstone in correlation with specimen


diameter (error bars indicate standard error of the mean) (after
Hoskins and Horino 1969). The size effect relations of Hoek and
Brown (1980) and Carpinteri et al. (1995) are also plotted for
comparison. The MFSL line has been fitted only to data [50 mm

Fig. 10 UCS of Kansas Limestone in correlation with specimen


diameter (error bars indicate standard error of the mean) (after
Hoskins and Horino 1969). The size effect relations of Hoek and
Brown (1980) and Carpinteri et al. (1995) are also plotted for
comparison. The MFSL line has been fitted only to data [50 mm

seen at a specimen diameter of 4060 mm, and that at


diameters smaller or larger than this the strength of the
specimen will decrease. If only considering strength predictions for specimens with a diameter [54 mm, the Hoek
and Brown (1980) relation seems to provide a reasonable
prediction for Purbeck Limestone and Hollington Sandstone. This cannot be said for the other rock types presented in Figs. 9, 10, 11.
1.3 Igneous Rocks
The recent work presented in Thuro et al. (2001) largely
contradicts the Weibull (1951) expectation of a strength
scale effect in materials containing micro-cracks i.e.
igneous or crystalline rock types. The experimental data
presented by Thuro et al. (2001) shows no such relationship

517

Fig. 11 UCS of various sedimentary rocks in correlation with


specimen diameter (after Hawkins 1998). The size effect relations
of Hoek and Brown (1980) and Carpinteri et al. (1995) are also
plotted (the MFSL has only been fitted to the Pennant Sandstone data,
and only to data [54 mm)

Fig. 12 UCS of granite in correlation with specimen diameter (error


bars indicate minimum and maximum values in data set) (after Thuro
et al. 2001). The size effect relations of Hoek and Brown (1980) and
Carpinteri et al. (1995) are also plotted for comparison

(Fig. 12). If proved by further studies, the implications of


these results are immense and far reaching; forcing a
reassessment of a large amount of the empirical theory that
is currently relied upon in rock mechanics design.
Aside from sedimentary rocks and plaster of Paris,
Hoskins and Horino (1969) present results for salida
granite which show an obvious size effect (Fig. 13).
Figure 14 shows the results of research carried out by
Jackson and Lau (1990). Their work on Lac du Bonnet
granite shows a decrease in strength with increasing
specimen size, although it is not as significant as what is
predicted by Eq. 1.
Yuki et al. (1995) studied the anisotropic behaviour of
Ohya Stone (welded tuff) over varying diameters from 30
to 150 mm. They showed that no strength decrease was
evident when specimens were loaded either parallel, or

123

518

Fig. 13 UCS of Salida granite in correlation with specimen diameter


(error bars indicate standard error of the mean) (after Hoskins and
Horino 1969). The size effect relations of Hoek and Brown (1980)
and Carpinteri et al. (1995) are also plotted for comparison

Fig. 14 UCS of Lac du Bonnet granite in correlation with specimen


diameter (error bars indicate 1.0 standard deviation from the mean)
(after Jackson and Lau 1990). The size effect relations of Hoek and
Brown (1980) and Carpinteri et al. (1995) are also plotted for
comparison

Fig. 15 UCS of Ohya Stone (welded tuff) (loaded horizontally to the


depositional surface) in correlation with specimen diameter (after
Yuki et al. 1995)

normal to the stratification, moreover, a slight increase in


strength was seen (Figs. 15, 16). This finding is extremely
pertinent and remains the only study to this date and to the
authors knowledge to consider an anisotropic strength

123

W. J. Darlington et al.

Fig. 16 UCS of Ohya Stone (welded tuff) (loaded vertically to the


depositional surface) in correlation with specimen diameter (after
Yuki et al. 1995)

Fig. 17 UCS of Saajome andesite in correlation with specimen


diameter (after Nishimatsu et al. 1969)

size effect in specimens with height to diameter ratios of


2:1. The unusual monotonically increasing nature of the
MFSL seen fitted to these increasing data sets is due to the
non-linear least squares fitting requirement inherent in the
MFSL function. Yuki et al. (1995) concludes that the
increase in UCS with increasing specimen diameter is
related to the percentage of pumice fragments contained in
the specimen. As the specimen size increases, so does the
percentage of pumice.
Nishimatsu et al. (1969) conducted sizestrength studies
on several igneous rocks (Figs. 17, 18, 19, 20, 21). The
saajome andesite shows a strong size effect and excellent
correlation with both the MFSL and Eq. 1. Nishimatsu
et al. (1969) tested specimens between 13 and 70 mm, but
aside from the aforementioned saajome andesite which has
a good spread of specimen diameters within this range, the
other rocks tested in the study fail to include data from the
3070 mm range. This makes it difficult to compare these
datasets to those of other researchers as it is impossible to
say if the low strength values seen at the 70 mm diameter
are preceded by a general decline, as would be modelled by
the MFSL or Eq. 1, with a peak strength at a diameter

The Effect of Specimen Size on Strength and Other Properties

519

Fig. 18 UCS of Ogino tuff in correlation with specimen diameter


(after Nishimatsu et al. 1969)

Fig. 21 UCS of Aoishi sandy tuff in correlation with specimen


diameter (after Nishimatsu et al. 1969)

Fig. 19 UCS of Inada granite in correlation with specimen diameter


(after Nishimatsu et al. 1969)

Fig. 22 UCS of Carthage marble in correlation with specimen


diameter (error bars indicate standard error of the mean) (after Hoskins
and Horino 1969). The size effect relations of Hoek and Brown (1980)
and Carpinteri et al. (1995) are also plotted for comparison

Fig. 20 UCS of Shinkomatsu andesite in correlation with specimen


diameter (after Nishimatsu et al. 1969)

produced by Hoskins and Horino (1969) (Fig. 22) and


relates to Carthage marble. Obviously it is not possible to
make any broad conclusions as to the existence and qualities of a sizestrength relationship in metamorphic rocks
due to a distinct lack of published data. However, the
sample supports a decrease in strength of approximately
7% when moving from a 50 to 127 mm specimen diameter.
Here again the power relation of Hoek and Brown (1980)
overpredicts this decrease in strength significantly. Clearly
more research is required into the strengthsize effect in
metamorphic rock types in order to reach sound and significant conclusions.
1.5 Youngs Modulus and Poissons Ratio

within this untested range (similar to what has been seen in


some sedimentary rocks) or if the low strength is an outlier
(although this seems unlikely).
1.4 Metamorphic Rocks
The only example of strengthsize effect testing regarding
metamorphic rocks reported in the published literature was

Aside from the works of Thuro et al. (2001) (Fig. 23),


Jackson and Lau (1990) (Fig. 24), Yuki et al. (1995)
(Figs. 25, 26) and the new results presented in this paper,
few have explored the existence of a size effect in Youngs
modulus or Poissons ratio. When considering the Youngs
modulus results of Thuro et al. (2001) a significant size
effect is not obvious. While Jackson and Lau (1990) show a

123

520

Fig. 23 Youngs modulus of limestone in correlation with specimen


diameter (error bars indicate minimum and maximum values in data
set) (after Thuro et al. 2001)

Fig. 24 Youngs modulus of Lac du Bonnet granite in correlation


with specimen diameter (error bars indicate 1.0 standard deviation
from the mean) (after Jackson and Lau 1990)

W. J. Darlington et al.

Fig. 26 Youngs modulus of Ohya Stone (welded tuff) (loaded


vertically to the depositional surface) in correlation with specimen
diameter (error bars indicate 1.0 standard deviation from the mean)
(after Yuki et al. 1995)

Fig. 27 Poissons ratio of Lac du Bonnet granite in correlation with


specimen diameter (error bars indicate 1.0 standard deviation from
the mean) (after Jackson and Lau 1990)

mentioned previously Yuki et al. believes this to be related


to the percentage of pumice fragments contained in the
specimen.
In the only published dataset pertaining to Poissons
ratio, Jackson and Lau (1990) observed a decrease in
Poissons ratio for Lac du Bonnet with increasing specimen
size (Fig. 27). Interestingly, a decrease in strength and
Youngs modulus with increasing size was also seen for the
same rock.
Fig. 25 Youngs modulus of Ohya Stone (welded tuff) (loaded
horizontally to the depositional surface) in correlation with specimen
diameter (error bars indicate 1.0 standard deviation from the mean)
(after Yuki et al. 1995)

significant decrease (*10%) in Youngs modulus with an


increase in specimen diameter (from 45 to 300 mm).
Conversely, Yuki et al. (1995) shows an increase in
Youngs modulus with specimen size, of approximately
27% for Ohya Stone (welded tuff) loaded horizontally to
the depositional surface and only 4% for Ohya Stone loaded vertically to the depositional surface. However, as

123

2 Experimental Method
In order to explore the effect of specimen size, high
strength cylindrical mortar specimens with a range of
diameters were manufactured. The diameters of the specimens are listed in Table 1. The 63.5 and 83.5 mm diameters were chosen in line with standard diamond drill rock
core sizes; HQ and PQ, respectively, while 150 mm
specimens are sometimes used in the initial stages of deep
drilling or in weak, weathered or fractured rock, and are
also commonly used during concrete testing. The 300 mm

The Effect of Specimen Size on Strength and Other Properties

521
100

Table 1 Cylindrical sample dimensions


Number of specimens

Diameter (mm)

Length (mm)

10

63.5

127

10

83.5

167

150

300

200

400

300

600

90

Percentage Passing

80
70
60
50
40
30
20
10

Table 2 Mix design

Sand
(kg)

Cement
(kg)

Water
(kg)

Plasticiser
(Glenium 27) (ml)

85

34

13.6

150

diameter specimens were chosen as the largest diameter


that could be tested at the current testing facility; with the
required failure load approaching the maximum load
capacity of the loading frame. Each cylindrical specimen
tested had lengths twice their diameter.
The mortar mix was designed to produce samples which
closely resemble sandstone in terms of peak strength,
Youngs modulus and Poissons ratio. As a high cement
water ratio was used, a plasticiser (Glenium 27) was added
to the cement mortar mix to increase workability and
prevent segregation. Additionally, the Glenium 27 aided by
increasing early and ultimate compressive strength and
Youngs modulus, while also decreasing any shrinkage.
These characteristics were desirable as it was found that
small amounts of shrinkage could cause precision specimen end preparation to differentially shrink to unacceptable levels of flatness and perpendicularity if not tested
soon after being prepared.
The cement mortar was mixed in an 85-l pan mixer
using the proportions outlined in Table 2. Initially half of
the total amount of sand required was placed in the mixer
and the total amount of cement was added followed by the
balance of the sand. The dry mix was then mixed for 60 s.
Half the volume of water was then added and it was mixed
for a further 60 s during which time the plasticiser was
poured in, followed by the remaining volume of water. The
mix was then rested for 120 s before it was finally mixed
for another 120 s.
The particle size distribution of the sand used in the
mortar is shown in Fig. 28. It is to be noted that the particle
size is not typical of natural sandstonethe effect of sand
particle size distribution will be examined in future studies.
After being left to cure for approximately 24 h the
specimens were removed from their individual moulds,
relocated to a misted curing room and left to cure for
28 days. Due to the large size of some of the specimens it
was not possible to use curing tanks.

10

100

1000

10000

Particle Size (m)

Fig. 28 Particle size distribution of sand used in cement mortar mix


(following methodology outlined in ASTM D422-63 (ASTM 2007)

Once the specimens had cured for 28 days, they were


prepared in accordance with ASTM D4543-01 (ASTM
2001). End flatness of the 63.5 and 83.5 mm diameter
specimens was achieved by machining the specimen ends
in a specially constructed v-block clamp fixed to a high
quality tool and cutter grinder. A diamond cup wheel with
a 126-grit size suitable for roughing and finishing precision
tools was used to grind the end surfaces. Once the specimen was clamped in the v-block it was not removed until
both ends had been ground. This was achieved by rotating
and relocating the entire v-block clamping system on the
tool and cutter grinder feeder table while the specimen was
secured. These techniques and procedures ensured that the
ends were parallel. The ends of the larger specimens were
trimmed parallel using a 900 mm diameter diamond saw
and large v-block clamping system.
The tests were conducted in accordance with ASTM
D7012-04 (ASTM 2004). A strain-controlled Amsler
loading frame with a 5,000-kN capacity was used during all
testing. Two Schaevitz LVDTs were used to measure the
axial displacement to a reported accuracy of 0.375% of the
full range. To calculate the applied strain, the respective
displacements were averaged.
In addition to this measure of axial strain four 68 mm
long Kyowa strain gauges were attached to the specimen;
two axially and two laterally around the specimen midheight at 908 intervals such that the axial strain gauges
were separated by 180. To avoid strain readings resulting
from artificial strength variations caused by the gauge
adhesive (where strain measurements may have been more
representative of the adhesive rather than the mortar under
examination or defects and voids in the mortar surface) a
two-stage application process was used. After accurate
centres and alignment lines were marked on the specimen,
a layer of epoxy resin was applied to the mortar surface
over these markings. The resin filled any surface irregularities and pore spaces. After the epoxy had cured it was

123

522

W. J. Darlington et al.

sanded back to the original surface level of the mortar,


leaving in situ any resin that had filled irregularities and
pore voids of the mortar along the gauging area. These
areas were then cleaned and neutralised prior to the
application of the strain gauges using the recommended
low temperature curing adhesive. A dataTaker DT8000
data acquisition system recorded the strain, displacement
and force readings at a rate of 20 Hz.
Published data pertaining to laboratory testing of cylinders of rock and rock-like brittle materials with height
to diameter ratios of 2:1 were collected from the published literature. Statistical analysis was carried out on the
new experimental data presented in this paper as well as
the data collected from the literature (as described earlier). The analysis was completed using five different
relational forms comprising a linear relationship between
UCS and diameter, a linear relationship between UCS and
1/diameter, a power relationship in the form of Hoek and
Brown (1980) relationship (i.e. a power relationship
between UCS and 1/diameter), the MFSL relationship
(Carpinteri et al. 1995) and an exponential relationship
between UCS and diameter. This comprehensive series of
statistical analysis had two aims. Firstly, to assess the
applicability of the currently accepted Hoek and Brown
(1980) relationship and the similar Cunha (1990) relationships for rock strength variation with specimen size.
Secondly, to explore the goodness of fit of other relational
forms.
Using simple transformations, all the commonly used
theoretical and empirical relationships along with other
mathematical forms to describe the strengthsize effect
could be reduced to linear equations enabling the use of
linear regression analysis methods. In addition to calculating the coefficients of the different relational forms, R2
values, p values, 95% confidence intervals and residual
plots were generated to assess the goodness of fit of the
models listed hereunder.
To test the fit of Hoek and Brown (1980), the following
transformation was used to enable the use of simple linear
regression analysis:
 k
50
rN rc50
d
 k !
 
 k

50
1
Y lnrcd ln rc50
ln 50 rc50 k ln
d
d
a bX
5
After completing the linear regression, the original
coefficients of the power curve were found using:
rc50

ea
; and k b
50k

123

Similarly the MFSL was transformed as follows:


r
lch
rN fc 1
d


lch
2
2
a bX
Y rN fc 1
d

After completing the linear regression, the coefficients


of the original equation were found using:
p
fc a
lch

b
fc2

The fit of an exponential model was also assessed using


the following transformation:
rN Aekd


Y lnrN ln Aekd ln A kd a bX

After completing the linear regression, the coefficients


of the original equation were found using:
A ea
kb
A simple linear model of the following form was also
assessed:
rN A kd
8
In addition to the simple linear model, the following
linear relationship between UCS and 1/diameter was also
tested:
1
9
rN A k :
d
3 Results and Discussion
3.1 StrengthSize Effect
The average 7-day strength of the mortar was found to be
41 MPa. To establish the consistency between the different
batches of cement mortar used in making the range of
various sized specimens, three 100 mm diameter reference
samples were taken from each mortar batch. The
mechanical properties of these cylinders were found to be
very consistent with an average UCS of 66.1 MPa
(SD = 3.8 MPa) and Youngs modulus of 23.4 GPa
(SD = 0.4 GPa).
Considering the specimens made to explore the size
effect; the average UCS and Youngs modulus across all of
the specimens were found to be 66.0 MPa (SD = 2.8 MPa)
and 23.7 GPa (SD = 1.5 GPa), respectively. The stress
strain results of a typical cylinder are presented in Fig. 29.
Comparing these standard deviations with the results of the

The Effect of Specimen Size on Strength and Other Properties

523

70

rcd rc63

Axial Stress (MPa)

60
50
40
30
20
10
0
-1500 -1000

-500

500

1000

1500

2000

2500

3000

3500

MicroStrain

Fig. 29 Typical stress strain relation for a 83.5 mm diameter cement


mortar cylinder

100 mm batch reference samples the figures are very


similar, suggesting that the variance seen in the strength
size specimens can almost entirely be explained by the
variation in mortar batch properties.
Figures 30, 31 and 32 show the findings of the current
study. The relationship between UCS and test specimen size
(Fig. 30) shows a decrease in the strength of a 63.5 mm
specimen when compared to that of a 300 mm specimen of
approximately 10%. The relations proposed by Hoek and
Brown (1980) (Eq. 1), Cunha (1990) (Eq. 2), and the MFSL
of Carpinteri et al. (1995) (Eq. 4) have been plotted
alongside the experimental data. The first two of these
relations fit the data only when small specimen diameters
are considered (63.5100 mm diameter), but grossly underpredict the UCS of 150, 200, and 300 mm diameter
specimens. The MFSL line of best fit shows an exceptional
fit across the range of specimen diameters tested.
For the current dataset, the decrease in strength with size
is not as strong as is predicted by Eqs. 1 and 2 (also plotted
in Fig. 30). A line of best fit has also been plotted in the
form of Eq. 10 resulting in a fit of reasonable quality
(R2 = 0.26). Here a power of 0.08 has been used instead of
0.16, as suggested originally by Hoek and Brown (1980).
Fig. 30 Experimental data
showing the relationship
between specimen diameter and
average UCS for high strength
cement mortar specimens. The
size effect relations of Hoek and
Brown (1980), Cunha (1990),
and a line of best fit in the form
of Carpinteri et al.s (1995)
MFSL (Eq. 3) are also plotted
for comparison

63
d

0:08
10

During the statistical analysis process, 95% confidence


intervals were estimated for each of the models under
investigation. In order to assess the applicability of the
models represented by Eqs. 1 and 2, the relevant
confidence intervals for cemented materials have been
listed in Table 3. The powers suggested by Hoek and
Brown (1980) and Cunha (1990) of 0.18 and 0.22,
respectively, clearly fall outside the 95% confidence
intervals of half the listed datasets suggesting these
powers are inappropriate when considering cement
stabilised materials. It can also be seen that most of these
results are not statistically significant, with the 95%
confidence intervals encompassing zero.
Some of the datasets, namely those of Hoskins and
Horino (1969) and Hawkins (1998) include data pertaining
to relatively small diameter specimens (i.e. \50 mm).
Considering these data sets, peak strength is seen in specimens ranging from 38 to 76 mm in diameter. Specimens
with diameters smaller or larger than these show lower
strengths, contradicting Weibulls (1951) theory. This has
consequences when trying to fit any of the relational
models commonly attributed to sizestrength effects in
rock [i.e. exponential, power, MFSL, linear (UCS vs.
1/diameter), etc.]. Fitting these models to complete data
sets (i.e. sets that include data pertaining to diameters less
than the peak strength diameter (generally around 50 mm)
can cause extremely poor statistical fits and large diameter
specimen strength predictions (which are of most interest
in terms of large scale design). This phenomenon is only
seen in the sedimentary rocks analysed. However, the
strength of some of the granite samples appears to
asymptote when the specimen diameter is \50 mm.
Assessing the 95% confidence intervals for Eqs. 1 and 2
for sedimentary rocks (Table 4), it can be seen that the
powers suggested by Hoek and Brown (1980) and Cunha
(1990) of 0.18 and 0.22, respectively, are included in many

90
85

1.20

80
1.10
75
1.00

70
65

0.90

60
0.80

55
50
0

50

100

150

200

250

300

350

400

0.70
450

123

524

W. J. Darlington et al.

Fig. 31 Relationship between specimen diameter and average


Youngs modulus (error bars indicate 1.0 standard deviation from
the mean)

Fig. 32 Relationship between specimen diameter and Poissons ratio

of the confidence intervals, especially those that only


include specimens of diameters larger than 54 mm. It can
also be seen that in datasets that include a maximum

strength at diameters larger than the smallest diameters


tested, the model becomes statistically significant when
these small diameter data points are omitted.
With regards to Table 5 and igneous rock types, it can
be seen that the powers suggested by Hoek and Brown
(1980) and Cunha (1990) of 0.18 and 0.22, respectively are
included in a majority of the confidence intervals, but zero
is also included in all but one of the intervals which suggests that this power relationship is not statistically significant at a 5% level.
The original empirical sizestrength relationship proposed by Hoek and Brown (1980) was based on data from
several sources. Each source employed varying test procedures and experimented using specimens that did not
necessarily conform to a single standard. For example, the
test results of Pratt et al. (1972) relating to quartz diorite
pertained to a mix of triangular and circular prisms each
with a height to depth ratio of 1.5, although consideration
of the varied cross-sections was not made. The results of
Mogi (1961) were also used by Hoek and Brown (1980) to
establish the size strength relation apparent in marble. The
specimens used in the aforementioned research were rectangular prisms.
Thuro et al. (2001), amongst others, have discussed
testing specimens of non-standard diameter to height ratios
and/or geometric volumes. Variations in specimen geometry can induce significant differences in stress distributions within a test specimen and lead to variation in
measured peak strength. As with in situ testing, although
these tests have a sound purpose and can provide valuable
information under specific circumstances, the appropriateness of using the results of these various methods in one

Table 3 95% confidence intervals for the power relationship in the form of Hoek and Brown (1980) and Cunha (1990) considering cement
stabilised materials
Reference

Material tested

Lower 95%

Upper 95%

Blanks and McNamara (1935)

Concrete (w/c = 0.53, Agre. max = 38.1 mm)

-0.005

0.123

Blanks and McNamara (1935)

Concrete (w/c = 0.55, Agre. max = 19.05 mm)

0.074

0.125

Blanks and McNamara (1935)

Concrete (w/c = 0.54, Agre. max = 9.52 mm)

0.058

0.136

Symons (1970)

Cement mortar (WGS 2%)

-4.353

4.273

Symons (1970)

Cement mortar (WGS 6%)

-1.780

1.727

Symons (1970)

Cement mortar (WGS 10%)

-0.927

0.765

Symons (1970)

Cement mortar (WGS 14%)

-0.824

0.882

Symons (1970)

Concrete (CR 2%)

-0.425

0.724

Symons (1970)
Symons (1970)

Concrete (CR 4%)


Concrete (CR 6%)

-0.036
-0.320

0.362
0.695

Symons (1970)

Concrete (CR 8%)

-0.421

0.555

New data presented in this paper

Cement mortar

0.017

0.100

Hoskins and Horino (1969)

Plaster of Paris

-0.078

0.060

Hoskins and Horino (1969)a

Plaster of Paris

0.039

0.080

Restricted data set:

123

includes only specimens with diameter [50 mm

The Effect of Specimen Size on Strength and Other Properties


Table 4 95% confidence
intervals for the power
relationship in the form of Hoek
and Brown (1980) and Cunha
(1990) considering cement
sedimentary rocks

Restricted data sets: includes


only specimens with diameter
[54 mm, b [38 mm, c [50 mm
Table 5 95% confidence
intervals for the power
relationship in the form of Hoek
and Brown (1980) and Cunha
(1990) considering cement
Igneous rocks

A single metamorphic rock


has also been included as only
one example exists in the
literature

Restricted data set: b includes


only specimens with diameter
[50 mm

Reference

525

Material tested

Lower 95%

Upper 95%

Hawkins (1998)

Pilton Sandstone

-0.138

0.127

Hawkins (1998)

Clifton Down Limestone

-0.318

0.241

Hawkins (1998)

Purbeck Limestone

-0.437

0.134

Hawkins (1998)

Pennant Sandstone

-0.340

0.122

Hawkins (1998)

Bath Stone

-0.313

0.119

Hawkins (1998)

Burrington Oolite

-0.311

0.139

Hawkins (1998)

Hollington Sandstone

-0.335

-0.047

Hoskins and Horino (1969)

Kansas Limestone

-0.068

0.046

Hoskins and Horino (1969)

Longmont Sandstone

-0.047

0.013

Natau et al. (1983)

Limestone (yellow)

0.951

1.714

Pells (2004)

Hawkesbury Sandstone (West Pymble Outcrop)

-0.244

0.039

Thuro et al. (2001)

Limestone

-0.329

0.220

Hawkins (1998)a

Pilton Sandstone

0.139

0.421

Hawkins (1998)b

Clifton Down Limestone

0.249

0.493

Hawkins (1998)b
Hawkins (1998)a

Purbeck Limestone
Pennant Sandstone

0.177
0.185

0.444
0.547

Hawkins (1998)a

Bath Stone

0.132

0.727

Hawkins (1998)a

Burrington Oolite

0.274

0.461

Hawkins (1998)a

Hollington Sandstone

0.053

0.253

Hoskins and Horino (1969)c

Kansas Limestone

-0.061

0.119

Hoskins and Horino (1969)c

Longmont Sandstone

-0.021

0.024

Reference

Material tested

Lower 95%

Upper 95%

Hoskins and Horino (1969)

Salida granite

-0.380

0.567

Jackson and Lau (1990)

Lac du Bonnet granite

-0.003

0.157

Thuro et al. (2001)

Granite

-0.251

0.314

Yuki et al. (1995)

Ohya Stone (welded tuff) loaded horizontally


to the depositional surface

-0.308

0.133

Yuki et al. (1995)

Ohya Stone (welded tuff) loaded vertically


to the depositional surface

-0.134

-0.012

Nishimatsu et al. (1969)

Saajome andesite

-0.010

0.282

Nishimatsu et al. (1969)


Nishimatsu et al. (1969)

Ogino tuff
Inada granite

-0.088
-0.098

0.242
0.354

Nishimatsu et al. (1969)

Shinkomatsu andesite

-0.163

0.262

Nishimatsu et al. (1969)

Aoishi sandy tuff

-0.128

0.134

Hoskins and Horino (1969)

Carthage marblea

-0.027

0.080

Hoskins and Horino (1969)b

Carthage marblea

-0.042

0.165

cohort to generate a general strengthsize relation needs


further investigation.
In order to assess the fit of the different models to the
new and published datasets R2 and residual plots enabled
a simple method applicable to the given relational forms.
Table 6 shows the R2 values for the five different models
applied to the various datasets pertaining to cemented
materials. It should be noted that R2 is not an absolute
measure of a models fit and that regression residual plots
were also analysed to ensure their randomness, but for the
purpose of discussion in this context R2 provides a simple

measure with which to compare the models fit to the


data. Another issue associated with the regression analysis
carried out on the published datasets is that they are
generally diameter grouped mean results, not entire sets
of experimental data. The issue of using mean grouped
data is highlighted by Rey et al. (2001) who show it will
artificially increase R2 values resulting from any regression analysis. For this reason the above tables are misleading as it is not possible to estimate the artificial
increase in R2 values without access to the original
complete datasets.

123

526

W. J. Darlington et al.

Table 6 R2 values and comparative model ranking for various relational forms obtained from regression analysis of publish data of cement
stabilised materials
Reference

R2 (rank)

Material tested

Linear

Linear
(1/diam)

Power
(1/diam)

MFSL

Expon.

Blanks and McNamara (1935)

Concrete (w/c = 0.53, Agre. max = 38.1 mm)

0.518 (3)

0.286 (4)

0.525 (2)

0.273 (5)

0.533 (1)

Blanks and McNamara (1935)

Concrete (w/c = 0.55, Agre. max = 19.05 mm)

0.808 (5)

0.872 (3)

0.954 (1)

0.884 (2)

0.825 (4)

Blanks and McNamara (1935)

Concrete (w/c = 0.54, Agre. max = 9.52 mm)

0.738 (5)

0.962 (2)

0.923 (3)

0.968 (1)

0.761 (4)

Symons (1970)

Cement mortar (WGS 2%)

0.001 (5)

0.063 (2)

0.014 (3)

0.065 (1)

0.001 (4)

Symons (1970)

Cement mortar (WGS 6%)

0.001 (5)

0.099 (1)

0.036 (3)

0.097 (2)

0.002 (4)

Symons (1970)

Cement mortar (WGS 10%)

0.433 (5)

0.708 (1)

0.596 (3)

0.695 (2)

0.447 (4)

Symons (1970)

Cement mortar (WGS 14%)

0.274 (2)

0.071 (4)

0.161 (3)

0.065 (5)

0.285 (1)

Symons (1970)

Concrete (CR 2%)

0.985 (1)

0.847 (5)

0.916 (3)

0.863 (4)

0.979 (2)

Symons (1970)

Concrete (CR 4%)

0.994 (2)

0.960 (5)

0.991 (3)

0.970 (4)

0.997 (1)

Symons (1970)

Concrete (CR 6%)

0.866 (5)

0.991 (1)

0.957 (3)

0.989 (2)

0.876 (4)

Symons (1970)

Concrete (CR 8%)

0.621 (4)

0.861 (2)

0.755 (3)

0.864 (1)

0.616 (5)

New data presented in this paper

Cement mortar

0.215 (5)

0.294 (1)

0.258 (3)

0.293 (2)

0.216 (4)

Hoskins and Horino (1969)

Plaster of Paris

0.004 (4)

0.148 (1)

0.029 (3)

0.137 (2)

0.002 (5)

Hoskins and Horino (1969)a

Plaster of Paris

0.932 (5)

0.944 (3)

0.967 (1)

0.947 (2)

0.934 (4)

Restricted data set:

includes only specimens with diameter [50 mm

Analysis of Table 6, which compares the models fit to


cementitious datasets shows that of all the models, the
power relationship consistently had the third highest R2
value, while the simple linear (UCS vs. 1/diameter) and the
theoretically derived MFSL relationship generated the first
and second highest. However, for these three models the R2
values were generally very similar, and in some cases the
exponential and a simple linear form fit the data most
closely. The same can also be said when considering the
published data sets featuring sedimentary rocks (Table 7),
although here there is stronger, more consistent support for
a linear (UCS vs. 1/diameter) or MFSL relationship fitting
the experimental data most closely.
Aside from evaluating the model fit solely on the
grounds of statistical parameters it is also important to
consider the intrinsic properties of a given relational form
when assessing the validity of its application to the size
strength relationship of rocks and other brittle materials.
The power and exponential relationships have horizontal
asymptotes of zero, so as the specimen diameter is
increased the predicted strength of the rock tends to zero.
Similarly unrealistic, a negative linear relationship will
predict strength values less than zero. Obviously the
strength of an infinitely large rock will never be zero (or
negative). Conversely both the theoretically derived MFSL
and the empirical linear (1/diameter) relation incorporate a
horizontal asymptote. As the rock diameter tends to infinity, both relations predict a rock strength which tends to a
constant. In light of the relative success of the MFSL
model to predict the strength of a majority of rocks, its

123

theoretical basis, and its close relation to what is logically


expected in a physical respect, it appears to be the best
model to use when modelling the decrease in strength of a
rock or cemented specimen with an increase in diameter.
Table 8 shows that the scale effect in igneous rocks is
best modelled by exponential or linear models, however,
given the aforementioned practical considerations and the
nature of these models to predict zero or negative strength
values at large specimen sizes, it seems unlikely they are
the most appropriate predictive tools. Figures 12, 13, 14,
15, 16, 17, 18, 19, 20, 21 and 22 show the MFSL producing
an acceptable fit when plotted alongside the experimental
datasets pertaining to igneous rocks.
Aside from the majority of experimental results that
indicate a strengthsize effect (with specimen strength
decreasing with size), there are some exceptions to the size
law including the samples tested by Symons (1970) (14%
well-graded sand-cement mortar), Pells (2004) (Hawkesbury Sandstone), Thuro et al. (2001) (limestone and granite), Hoskins and Horino (1969) (Longmont Sandstone),
and Yuki et al. (1995) (welded tuff) that show an increase
in strength with specimen size or no significant strength
size relationship. Yuki et al. (1995) justifies this monotonically increasing relationship due to the fractal distribution of hard particles within the rock. This is an
important point and highlights the possible role of micro
structure and rock fabric in any size effect relationship. The
lack of a clearly consistent size effect makes application of
any derived relationship or model difficult and warrants
further study to quantify which materials demonstrate a

The Effect of Specimen Size on Strength and Other Properties

527

Table 7 R2 values and comparative model ranking for various relational forms obtained from regression analysis of publish data of sedimentary
rocks
Reference

R2 (rank)

Material tested

Linear

Linear
(1/diam)

Power
(1/diam)

MFSL

Expon.

Hawkins (1998)

Pilton Sandstone

0.073 (3)

0.105 (1)

0.002 (5)

0.100 (2)

0.072 (4)

Hawkins (1998)

Clifton Down Limestone

0.051 (3)

0.117 (1)

0.018 (5)

0.061 (2)

0.020 (4)

Hawkins (1998)

Purbeck Limestone

0.009 (5)

0.320 (1)

0.220 (2)

0.188 (3)

0.045 (4)

Hawkins (1998)

Pennant Sandstone

0.003 (5)

0.373 (1)

0.182 (3)

0.285 (2)

0.016 (4)

Hawkins (1998)

Bath Stone

0.009 (5)

0.246 (1)

0.167 (3)

0.176 (2)

0.026 (4)

Hawkins (1998)

Burrington Oolite

0.000 (5)

0.314 (1)

0.128 (3)

0.240 (2)

0.003 (4)

Hawkins (1998)

Hollington Sandstone

0.301 (5)

0.794 (1)

0.638 (3)

0.723 (2)

0.327 (4)

Hoskins and Horino (1969)

Kansas Limestone

0.001 (5)

0.200 (1)

0.069 (3)

0.195 (2)

0.001 (4)

Hoskins and Horino (1969)

Longmont Sandstone

0.573 (5)

0.881 (1)

0.749 (3)

0.881 (2)

0.573 (4)

Natau et al. (1983)

Limestone (yellow)

0.307 (5)

0.625 (2)

0.638 (1)

0.525 (4)

0.540 (3)

Pells (2004)

Hawkesbury Sandstone (West Pymble Outcrop)

0.923 (1)

0.571 (4)

0.830 (3)

0.561 (5)

0.913 (2)

Thuro et al. (2001)

Limestone

0.053 (5)

0.088 (1)

0.071 (3)

0.086 (2)

0.055 (4)

Hawkins (1998)a

Pilton Sandstone

0.959 (2)

0.894 (5)

0.930 (3)

0.914 (4)

0.960 (1)

Hawkins (1998)
Hawkins (1998)b

Clifton Down Limestone


Purbeck Limestone

0.795 (5)
0.746 (5)

0.974 (1)
0.959 (1)

0.947 (3)
0.912 (3)

0.952 (2)
0.934 (2)

0.858 (4)
0.802 (4)

Hawkins (1998)a

Pennant Sandstone

0.888 (5)

0.940 (2)

0.932 (3)

0.943 (1)

0.914 (4)

Hawkins (1998)a

Bath Stone

0.758 (5)

0.911 (1)

0.876 (3)

0.892 (2)

0.799 (4)

Hawkins (1998)a

Burrington Oolite

0.920 (5)

0.988 (2)

0.981 (3)

0.989 (1)

0.941 (4)

Hawkins (1998)a

Hollington Sandstone

0.888 (3)

0.878 (5)

0.888 (2)

0.892 (1)

0.887 (4)

Kansas Limestone

0.376 (2)

0.139 (3)

0.257 (5)

0.136 (4)

0.381 (1)

Longmont Sandstone

0.250 (5)

0.429 (2)

0.339 (3)

0.429 (1)

0.250 (4)

Hoskins and Horino (1969)

Hoskins and Horino (1969)c


a

Restricted data sets: includes only specimens with diameter [54 mm, [38 mm, [50 mm

Table 8 R2 values and comparative model ranking for various relational forms obtained from regression analysis of publish data of igneous
rocks
Reference

Material tested

Hoskins and Horino (1969)

Salida granite

R2 (rank)
Linear

Linear
(1/diam)

Power
(1/diam)

MFSL

Expon.

0.954 (1)

0.772 (5)

0.863 (3)

0.782 (4)

0.948 (2)

Jackson and Lau (1990)

Lac du Bonnet granite

0.474 (3)

0.379 (4)

0.478 (1)

0.377 (5)

0.474 (2)

Thuro et al. (2001)

Granite

0.103 (2)

0.097 (4)

0.104 (1)

0.095 (5)

0.102 (3)

Yuki et al. (1995)

Ohya Stone (welded tuff) loaded horizontally


to the depositional surface

0.731 (1)

0.433 (5)

0.594 (3)

0.452 (4)

0.708 (2)

Yuki et al. (1995)

Ohya Stone (welded tuff) loaded vertically


to the depositional surface

0.963 (1)

0.830 (4)

0.930 (3)

0.815 (5)

0.963 (2)

Nishimatsu et al. (1969)


Nishimatsu et al. (1969)

Saajome andesite
Ogino tuff

0.801 (2)
0.396 (2)

0.628 (4)
0.062 (4)

0.745 (3)
0.224 (3)

0.618 (5)
0.048 (5)

0.813 (1)
0.427 (1)

Nishimatsu et al. (1969)

Inada granite

0.662 (2)

0.272 (4)

0.520 (3)

0.243 (5)

0.700 (1)

Nishimatsu et al. (1969)

Shinkomatsu andesite

0.329 (2)

0.013 (4)

0.154 (3)

0.008 (5)

0.354 (1)

Nishimatsu et al. (1969)

Aoishi sandy tuff

0.001 (4)

0.009 (1)

0.002 (3)

0.008 (2)

0.000 (5)

Hoskins and Horino (1969)

Carthage marblea

0.615 (2)

0.280 (4)

0.451 (3)

0.279 (5)

0.617 (1)

Hoskins and Horino (1969)b

Carthage marblea

0.783 (1)

0.745 (5)

0.767 (3)

0.751 (4)

0.783 (2)

A single metamorphic rock has also been included as only one example exists in the literature

Restricted data set:

includes only specimens with diameter [50 mm

123

528

size effect whether it be a positive or negative trend and


which do not.
3.2 Youngs Modulus and Poissons Ratio-Size Effect
When considering Fig. 31 it can be concluded that for the
material tested in this study there is a slight increase in
Youngs modulus (approximately 8%), although considering the standard deviations of the data this trend is difficult
to support statistically. Poissons ratio shows large variability with little evidence of a strong size effect, although
in this contribution only very few data were obtained
(Fig. 32). More generally there seems to be a general lack
of data pertaining to this area of research, and in light of the
inconsistencies seen between the published datasets it is
not possible to make a sound conclusion regarding the size
effect in Youngs modulus or more acutely Poissons ratio.

4 Conclusion
Assessing the available Youngs modulus and Poissons
ratio data it cannot be confirmed whether a size effect
exists. Given the limited quantity of data on these properties relationship with specimen size no conclusion can
be considered statistically significant and further research
is warranted in this area.
Considering the new and previously published data
presented in this paper on the strengthsize effect, it is
clear that large variations in the magnitude of any relationship exist. Possible reasons for this variation must be
established. It is hypothesised that no consistent result can
be seen in the published experimental data due to one or
more of the following issues: (a) the testing method/
apparatus used; (b) specimen preparation; and/or (c) the
type of material under examination (not-excluding the
possibility of anisotropic or load orientation strengthsize
relationships).
The particulars of any testing method and apparatus will
inevitably generate some variation in results. It is possible
that the variations seen in the published data are due to a
high sensitivity to the testing methodologies. Specifically,
high sensitivities may be associated with boundary conditions including platen friction, the effect of capping
materials (if used) and specimen end preparation (including
flatness, perpendicularity and smoothness). Other experimental peculiarities such as test rig stiffness, load rate, etc.
may also be a cause for this variability between different
researchers results. It is hypothesised that these factors
have a strong influence in causing some of the high variability, and low strengths seen in specimens \50 mm in
diameter, where stress concentrations due to poor end
preparation will lead to a comparatively dramatic effect on

123

W. J. Darlington et al.

specimen strength (when compared to the effect a stress


concentration may have on a specimen larger than 50 mm
in diameter).
It appears that any scale relationship is highly material
dependent. It seems probable that the nature of any
strengthsize effect is determined to some degree by the
materials general structural classification (igneous, sedimentary or otherwise). This paper presents results showing
high variations of the strengthsize effect even within single
material classifications. Igneous rocks appear to fit all of the
tested sizestrength relations relatively poorly when compared to sedimentary rocks fitted to the same models.
Generally speaking, a majority of the results analysed
are modelled best by a linear (1/diameter) relationship or
the MFSL. Both models incorporate a horizontal asymptote
dictating large scale rock strength. Furthermore, it is
impossible to model the scalestrength effect of all rocks
using a power law model with a single fixed general power.
More research is clearly needed to fully understand this
phenomenon. Such research should aim at generating
consistency across new datasets and aim to generate consistency within a dataset, using a single rock type or
material over a complete scale range (i.e. testing
20300 mm specimens) while maintaining consistent
boundary conditions. The results of specimens at, or
exceeding, 300 mm are of crucial importance given a
specific and distinct lack of these data in publications to
date and their practical relevance to large scale design,
improving statistical relevancy and confirming the existence of a horizontal asymptote and the threshold diameter
at which asymptotic strength is met.
The approach used in this paper is based mainly on
statistical analysis of existing data. The results have indicated that the scale relationship depends strongly on the
type (or structural classification) of the rock. It may be
worthwhile to study the geometric and physical properties
of the heterogeneities that are generally observed in each
type of rock; how evolution of the failure process may be
influenced by the nature of the heterogeneity and the
applied load; how other factors such as specimen size and
boundary conditions may affect the measured peak
strength. This may help to explain the rock type-dependent
scale relationship based on a mechanistic approach.
Acknowledgments The authors would like to extend their thanks
and appreciation to the undergraduate student Sajmir Bella for his
assistance with the laboratory work associated with this project.

References
ASTM (2001) Standard practices for preparing rock core specimens
and determining dimensional and shape tolerances, Designation:
D 454301. Annual Book of ASTM Standards. ASTM International, West Conshohocken

The Effect of Specimen Size on Strength and Other Properties


ASTM (2004) Standard test method for compressive strength and
elastic moduli of intact rock core specimens under varying states
of stress and temperatures, Designation: D 7012-04. Annual
Book of ASTM Standards. ASTM International, West
Conshohocken
ASTM (2007) Standard test method for particle-size analysis of soils,
Designation: D422-63. Annual Book of ASTM Standards.
ASTM International, West Conshohocken
Bazant ZP (1984) Size effect in blunt fracture: concrete rock metal.
J Eng Mech (ASCE) 110:518535
Blanks RF, Mcnamara CC (1935) Mass concrete tests in large
cylinders. J Am Concr Inst 31:280303
Brace WF (1981) The effect of size on mechanical properties of rock.
Geophys Res Lett 8:651652
Carpinteri A, Chiaia B, Ferro G (1995) Size effects on nominal tensile
strength of concrete structures: multifractality of material
ligaments and dimensional transition from order to disorder.
Mater Struct 28:311317
Carpinteri A, Ferro G, Monetto I (1999) Scale effects in uniaxially
compressed concrete specimens. Mag Concr Res 51:217225
Cunha AP (1990) Scale effects in rock mass. In: Cunha AP (ed)
Proceeding of International Workshop on Scale Effects in Rock
Masses, Balkema
Hawkins AB (1998) Aspects of rock strength. Bull Eng Geol Environ
57:1730
Hoek E (2000) Rock engineering course notes by Evert Hoek.
http://www.rocscience.com/education/hoeks_corner
Hoek E, Brown ET (1980) Underground Excavations in Rock.
Institute of Mining and Metallurgy, London
Hoskins JR, Horino FG (1969) Influence of spherical head size and
specimen diameters on the uniaxial compressive strength of
rocks. US Department of the Interior, Bureau of Mines,
Washington

529
Jackson R, Lau JSO (1990) The effect of specimen size on the
laboratory mechanical properties of Lac du Bonnet grey
gragranite. In: Cunha P (ed) Scale effects in rock masses.
Balkema, Rotterdam
Mogi K (1961) The influence of the dimensions of specimens on the
fracture strength of rocks. Bull Earthq Res Inst 40:175185
Natau O, Frolich BO, Mutschler TO (1983) Recent development of
the large scale triaxial test. In: ISRM Congress, Melbourne, pp
A65A74
Nishimatsu Y, Yamaguchi U, Motosugi K, Morita M (1969) The size
effect and experimental error of the strength of rocks. J Min
Mater Process Inst Jpn 18:10191025
Pells PJN (2004) On the absence of size effects for substance strength
of Hawkesbury Sandstone. Aust Geomech 39:7983
Pratt HR, Black AD, Brown WS, Brace WF (1972) The effect of
specimen size on the mechanical properties of unjointed diorite.
Int J Rock Mech Min Sci 9:513529
Rey CCM-VD, Galindo MP, Velarde MAA (2001) Effects of using
mean scores in regression models: an example from environmental psychology. Qual Quant 35:191202
Symons IF (1970) The effect of size and shape of specimen upon the
unconfined compressive strength of cement-stabilized materials.
Mag Concr Res 22:4550
Thuro K, Plinninger RJ, Zah S, Schutz S (2001) Scale effects in rock
strength properties. Part1: Unconfined compressive test and
Brazilian test. In: Rock mechanicsa challenge for society, June
3-7 2001. ISRM, Espoo, pp 169174
Weibull W (1951) A statistical function of wide applicability. J Appl
Mech 18:293297
Yuki N, Aoto S, Yoshinaka R, Yoshihiro O, Terada M (1995) The
scale and creep effect on the strength of welded tuff. In:
Yoshinaka R, Kikuchi K (eds) International Workshop on Rock
Foundation. Balkema, Tokyo

123

Vous aimerez peut-être aussi