You are on page 1of 15

University of kirkuk – College of engineering

Civil department – Third stage A

Triaxial shear behavior of a cement-


treated sand and gravel mixture

Supervised by: Dr.Aram Mohammed


Prepared by: Mahmood Younis Mohammed (A)
Abstract:
A number of parameters, e.g. cement content, cement type, relative density, and grain
size distribution, can influence the mechanical behaviors of cemented soils. In the present
study, a series of conventional triaxial compression tests were conducted on a cemented
poorly graded sand gravel mixture containing 30% gravel and 70% sand in both
consolidated drained and untrained conditions. Portland cement used as the cementing
agent was added to the soil at 0%, 1%, 2%, and 3% (dry weight) of sand gravel mixture.
Samples were prepared at 70% relative density and tested at confining pressures of 50
kPa, 100 kPa, and 150 kPa. Comparison of the results with other studies on well graded
gravely sands indicated more dilation or negative pore pressure in poorly graded samples.
Untrained failure envelopes determined using zero Skempton’s pore pressure coefficient
ðA ¼ 0Þ criterion were consistent with the drained ones. Energy absorption potential
was higher in drained condition than untrained condition, suggesting that more energy
was required to induce deformation in cemented soil under drained state. Energy
absorption increased with increase in cement content under both drained and untrained
conditions. _ 2014 Institute of Rock and Soil Mechanics, Chinese Academy of Sciences.
Production and hosting by Elsevier B.V. All rights reserved.

1. Introduction
Cementation occurs due to the various geological processes that create bonds between
soil particles like aging, chemical reactions, carbonates, silicates, iron oxides, and natural
cementing agents. Due to the difficulties of in situ sampling, the mechanical
characteristics of cemented soils are usually studied using artificial samples prepared in
laboratory and cured by different cementing agents. The mechanical behavior of
cemented soils is influenced by a number of parameters including cement content, cement
type, density, confining stress, grain size, and stress-strain history (Saxena and Lastrico,
1978; Clough et al., 1979, 1981, 1989; Acar and El Tahir, 1986; Leroueil and Vaughan,
1990; Airey, 1993; Coop and Atkinson, 1993; Das et al., 1995; Malandraki and Toll,
1996; Cuccovillo and Coop, 1997, 1999; Huang and Airey, 1998; Consoli et al., 2000,
2007, 2009, 2010, 2011; Schnaid et al., 2001; Ismail et al., 2002; Rotta et al., 2003; Lee
et al., 2010; Park, 2010; Baxter et al., 2011; Hamidi and Hooresfand, 2013; Shahnazari
and Rezvani, 2013). According to the previous studies, the cementation can increase
brittleness, shear strength, and dilative behavior of sands. However, it should be noted
that most of the previous studies have focused on the mechanical behavior of cemented
fine sands, rather than the mechanical behavior of coarse grained gravels or gravely
sands. In the last decade, a number of studies have been performed to investigate the
mechanical behavior of cemented gravely sands or sandy gravels (Haeri et al., 2005a,b,
2006), in which a series of triaxial compression tests were performed on a representative
gradation of the Tehran coarse grained alluvium. Cemented samples were prepared with
different cementing agents including lime, Portland cement, and gypsum. They
concluded that the strain associated with the peak deviatory stress decreases as the
cementation increases. Also it was indicated that the maximum rate of dilation and
negative pore water pressure occur after the maximum shear strength is obtained. Review
of the literature shows that there is rarely particular study on investigation of the
mechanical behavior of cemented poorly graded sand gravel mixtures. Indeed, all
previous studies on the behavior of cemented gravely sands concern fine sands or well
graded gravely sands as the base soil. Therefore, the objective of present research is to
investigate the mechanical behavior of a cement-treated poorly graded sand gravel
mixture. In this regard, a number of new features of the mechanical behavior of cemented
soils are reported.

2. Testing program
Thirty groups of conventional triaxial compression tests were performed, among which
24 groups are considered in this study. Six groups of additional tests (25% of the total)
were performed to check repeatability of the experiments and results. Also, 12 groups of
unconfined compression tests were conducted and reported. Cement content and
confining pressure were considered as the variables of testing program, and the triaxial
compression tests were performed in the consolidated drained and untrained conditions.

2.1. Soil and cementing agent


Clean and uniform quartz beach sand with sub-round to sub angular particles from the
shores of the Caspian sea (specifically Babolsar, Iran) was first sieved using a #30 sieve
and then was mixed with 30% uni-sized (9.5e12.5 mm) gravel grains. The mixed soil can
be named as SP in unified soil classification system and was used as the base material.
Gradation curves and physical properties of the base material are shown in Fig. 1 and
Table 1, respectively. In Table 1, Gs is the specific gravity; D10 is the effective diameter;
CU and CC are the uniformity and curvature coefficients, respectively; and gd, min and
gd, max are the minimum and maximum unit weights, respectively. All physical
characteristics were determined according to the ASTM (1998) standard methods.
Portland cement (Type II) with a setting time lasting for about 4 h was used as the
cementing agent. It was first sieved using a #100 sieve and then added to the base soil.

2.2. Sample preparation and test procedure


The undercompaction method was used for sample preparation proposed by Ladd (1978).
Required weight of the soil was mixed with desired cement content and about 7%
distilled water. Samples were prepared using a split mold, 100mmin diameter and 200mm
in height, and were compacted in eight layers. Each layer was poured into the mold and
compacted using metal hammer until the desired height was reached. Cemented samples
were stored at a (25 3) C humid room with >90% relative
humidity for 24 h. After that, samples were extracted from the mold and were kept for 6 d
at humid room. On the 7th day, the diameter, height and weight of the samples were
measured. For unconfined compression tests, samples were prepared with curing times of
7 d and 28 d. The variables considered in sampling process are listed in Table 2. A
computer-controlled triaxial cell was used to test the samples at the confining pressures
(CPs) of 50 kPa,100 kPa, and 150 kPa. The outer surface of samples was soft enough to
minimize the effect of membrane penetration. As a result, flexible membranes do not
affect pore pressure generation in saturated condition. Membranes with average thickness
of 1 mm were used and corrections such as

membrane thickness and cross-sectional area were considered according to Bishop and
Henkel (1969). All samples were fully saturated in two stages prior to shearing. At the
first stage, de-aired water was flushed from the bottom of sample under a very low
pressure difference of 10 kPa for 24 h. After that both cell and back pressure were
ramped simultaneously to 310 kPa and 300 kPa for complete saturation at the second
stage. Saturation procedure was considered to be completed until Skempton’s B value of
0.9 was reached. The samples were consolidated up to the desired confining pressures.
Shear loading was applied at an axial strain rate of 0.1 mm/min for drained tests and 0.3
mm/min for undrained ones. Cell pressure, volume change, pore pressure, load and
displacements were measured during triaxial compression tests by electronic transducers
and a calibrated data acquisition system. All the variables considered in testing program
are listed in Table 3.
3. Unconfined compression tests
Some specifications of the unconfined compression tests are described in Table 3. Fig. 2
indicates the variation of unconfined compressive strength (UCS) with cement content at
different curing times. Peak strength occurred at small strains between 0.2% and 0.7%. It
can be observed from Fig. 2 that the UCS increases with increasing cement content and
elapsed curing time. The lines intersect horizontal axis at cement content about 0.5%,
which is the minimum cement content to mobilize the shear strength of cemented soil and
formation of cemented bonds.

4. Analysis of the results


A summary of triaxial test results at failure and residual state is shown in Table 4.
Deviatoric stress (q), mean effective stress (p’) and specific volume (n) are defined using
the following equations:

Fig. 2. Variation of the unconfined compressive strength with cement content at


different curing times.

where s0 1 is the major effective principal stress, s03 is the minor effective
principal stress, and e is the void ratio.
4.1. Mode of failure
Fig. 3 shows the typical failure modes of the uncemented and cemented samples.
Although dilation occurred at different confining pressures, all uncemented samples in the
drained and undrained tests showed barreling mode without shear plane formation. In
lightly cemented samples (CC ¼ 1%), failure mode was a combination of barreling shape
and shear plane, although shear band was not obvious and barreling was the predominant
mode. Increase in cement content increased the thickness of the shear band. Cemented
samples with more cement content (CC > 1%) experienced a mode of brittle failure and
underwent significant dilation with apparent peak point in stressestrain curve. In the
drained and undrained conditions, cemented soils showed a shear band with a thickness
of 3e6 cm. Inclination of the shear band with horizontal axis decreased from 70_ to 55_
with increase in confining pressure from 50 kPa to 150 kPa.
4.2. Stress train behavior
Deviatoric stress axial strain curves are depicted in Figs. 4a, 5a, 6a and 7a for the
consolidated drained and undrained conditions. The drained and undrained tests are
indicated by the symbols “D” and “U”, respectively, and are followed by a number that
represents the value of confining pressure in kPa. Six tests were repeated twice to check
the repeatability of experiments, and the average stresse strain diagram is plotted in these
figures. All cemented samples in drained and undrained conditions showed an apparent
peak point
associated with the failure. After that, slope of the stress strain curve decreased to its residual
value in an axial strain of about 20%. For the uncemented samples under undrained condition,
the peak stress was not obvious and the softening behavior was not as clear as the drained tests.
Increase in cement content and reduction in confining pressure caused more softening in stress-
strain curve. Comparison of the results in the drained and undrained conditions showed that the
strain associated with the peak strength was larger
in undrained tests than that in drained ones. It can be concluded that the cemented soil
behavior is more brittle in the drained condition and bond degradation occurs easier when
volumetric strains can freely occur in the soil (Malandraki and Toll, 2001).
In order to understand the effect of confining pressure on the shear strength of cemented
soil, the maximum deviatoric stress (qmax) is normalized to the uncemented one in
drained tests in Fig. 8. It can be seen from Fig. 8 that, for lightly cemented sample (CC ¼
1%), normalized shear strength was approximately constant under different confining
pressures. However, the effect of confining pressure on the shear strength ratio increased
for other cement contents.
confining pressure increased contractive behavior at the start of the test. Figs. 4c, 5c, 6c
and 7c indicate the variation of the excess pore pressure with the mean effective stress in
the consolidated undrained tests. Positive pore pressure occurred at the beginning of
loading, followed by significant negative pore pressure at the final state. Same as the
volume change, increase in cement content and decrease in confining pressure increased
negative suction at the end of loading process. Haeri et al. (2005b) reported the results of
consolidated drained triaxial tests on a well graded gravely sand (containing 45% gravel)
cemented with Portland cement. The gradation curve is shown in Fig. 1 and the
maximum gravel size is 12.5 mm. Fig. 9 shows the comparison between the dilation and
excess pore pressure values for the present study and well graded gravely sand tested by
Haeri et al. (2005b). According to the previous studies, increase in gravel content
increases dilation in sandegravel mixtures (Evans and Zhou, 1995; Simoni and Houlsby,
2006; Hamidi et al., 2009). Although gravel content is larger inwell graded samples,
dilation in drained state or negative pore pressure in undrained condition is lower
compared with the poorly graded mixture tested. For the same gravel content, it can be
concluded that dilation in drained condition or negative pore pressure in undrained state
is larger in poorly graded sandegravel mixtures compared with the cemented well graded
gravely sands.

4.4. Stress paths

Stress paths for different cement contents are shown in Figs. 4d, 5d, 6d and 7d. The stress
path moved linearly with a slope of 3:1 in q-p’ space in the consolidated drained tests. It
reached a peak point which has been marked on the figures. After that, softening caused
reversal of the stress path with the same slope until residual stress state was reached. In
undrained condition, the peak point of the stress path was higher than the drained one due
to generation of significant negative pore pressure. However, the difference between peak
points of stress paths for the drained and undrained
tests decreases by increase in confining pressure. The same trend has also been reported
for a gypsum cemented well graded gravely sand (Haeri et al., 2005a); however, peak
points of the stress path in undrained condition were lower than the drained ones at high
confining pressures (over 300 kPa). Fig. 10 shows the axial strain contours of cemented
samples containing 2% cement in the drained and undrained conditions. Peak shear stress
has been observed in an axial strain level about 1% in drained condition and 3% in
undrained condition. Results of triaxial tests on the cemented well graded gravely sand
indicate that peak shear stress is associated with axial strains of 2.5% and 4.5% in drained
and undrained states, respectively (Haeri et al., 2005b). It confirms more brittle behavior
of cemented poorly graded sandegravel mixture compared with the cemented well graded
gravely sand in the same conditions.
4.5. Failure envelopes and shear strength parameters

Fig. 11 plots failure envelopes for all the tests. Numbers after the symbols “D” and “U”
show the value of cement content. Previousstudies have suggested a curved failure
envelope for cemented materials (Malandraki and Toll, 2001; Asghari et al., 2003; Baker,
2004; Sharma et al., 2011). However, the failure envelopes were not curved for the
studied soil. Moreover, the failure envelopes of drained tests were lower than those of
undrained ones which can be related to the large negative pore pressures induced during
undrained shearing. Different criteria can be used to determine the shear strength
parameters for the soil, including peak deviatoric stress at failure, maximum principal
stress ratio, maximum excess pore pressure, and zero Skempton’s pore pressure
parameter ðA ¼ 0Þ. The peak deviatoric stress is a commonly used method for
determination of the failure envelope in drained condition. In undrained state, Baxter et
al. (2011) recommended using A ¼ 0 as the failure criterion for cemented soils, because
it can eliminate the effects of large pore pressure gradients on the shear strength. Fig. 12a
plots failure envelopes for drained condition based on the peak deviatoric stress and Fig.
12b shows undrained failure envelopes based on both peak deviatoric stress and A ¼ 0
criteria. Table 5 also lists the shear strength parameters, c0 and 40.

According to Table 5, using A ¼ 0 criterion in undrained condition


yields consistent friction angle and cohesion intercept values with drained state. Baxter et
al. (2011) mentioned that increase in cement content increases the differences between
cohesion intercept values calculated in the drained and undrained conditions using A ¼ 0
criterion. However, results of the present study show that increase in cement content does
not particularly influence the relation between shear strength parameters in the

drained and undrained conditions. It can be related to the difference in gradation of the
cemented sandy gravel used in the present study and the fine cemented sand tested by
Baxter et al. (2011). Fig. 13 depicts the failure envelopes and residual state lines for
different cement contents in drained condition. Residual state line shows the relation
between deviatoric stress and mean effective stress in the final stage of loading. It can be
seen from Fig. 13 that the difference between the two lines increases with increasing
cement content, which confirms more brittle behavior as cement content increases. Fig.
14 shows variation of the effective principal stress ratio at failure ðs0 1=s0 3 Þ f for
drained and undrained conditions with confining pressure. The effective principal stress
ratio at failure decreases with increasing confining pressure and increases with increasing
cement content. Also, it is much higher in undrained state than drained state, especially at
lower confining pressures.

5. Stiffness and energy absorption


Stiffness of cemented soil was determined by calculating the secant modulus for half
shear strength in different confining pressures as shown in Fig. 15. We can note that the
stiffness increases with increase in cement content and confining pressure. The drained
stiffness is larger than the undrained one; however, the difference is less than 10%.
Required energy to induce deformation in cemented soil can be calculated by the area
under stress-strain curve. In the present study, absorbed energy for different axial strains
is normalized to the absorbed energy at 10% axial strain. Fig. 16 shows the variation of
the normalized absorbed energy for different cement contents and confining pressures in
the drained and undrained conditions. A major change in the slope of the curve can be
observed at 2% axial strain in drained state. As the cement content decreases, slope of the
curve became flattened and reached to the curve of the uncemented samples. For the
undrained condition, the slope of the curve was lower than that of the drained one. Also,
the change in the slope of energy absorption curve can be observed at axial strains more
than 2%.

Conclusions:
The present study deals with the engineering characteristics of cemented poorly graded
sand & gravel mixtures. The following conclusions can be drawn based on the test
results: (1) Under the similar conditions (the same cement type, cement content, gravel
content, and maximum gravel size), dilation and negative pore pressure induced in
cemented poorly graded sand & gravel mixture were larger than those of well graded
gravely sand. (2) Brittle behavior of cemented soils was more obvious in drained
condition than in untrained state. The axial strain at the peak shear strength was smaller
in drained condition than in untrained state. Failure envelopes were lower in the drained
condition than in the untrained state. (3) Shear strength parameters calculated under the
drained and untrained conditions were consistent when untrained shear parameters were
calculated based on zero Skempton’s coefficient ðA ¼ 0Þ criterion. Consistency between
the results can be observed for all cement contents. (4) A major change in the slope of the
absorbed energy curve was observed at axial strain of about 2% in the drained condition.
As cement content decreases, the slope of the curve became flattened and reached to the
curve of the uncemented soil. The change in the slope of energy absorption curve was
minor in untrained condition at axial strains more than 2%.
Conflict of interest
The authors wish to confirm that there are no known conflicts of interest associated with
this publication and there has been no significant financial support for this work that could
have influenced its outcome.

References
Acar Y, El-Tahir A. Low strain dynamic properties of artificially cemented sand.
Journal of Geotechnical Engineering 1986;112(11):1001e15.
Airey DW. Triaxial testing of naturally cemented carbonate soil. Journal of
Geotechnical Engineering 1993;119(9):1379e98.
Asghari E, Toll DG, Haeri SM. Triaxial behavior of a cemented gravelly sand Tehran
alluvium. Geotechnical and Geological Engineering 2003;21(1):1e28.
American Society for Testing and Materials (ASTM). Annual book of ASTM standards:
soils and rock. West Conshohocken, Philadelphia: ASTM; 1998.
Baxter CDP, Sharma MSR, Moran K, Vaziri H, Narayanasamy R. Use of ðA ¼ 0Þ as a
failure criterion for weakly cemented soils. Journal of Geotechnical and Geoenvironmental
Engineering 2011;137(2):161e70.
Baker R. Nonlinear Mohr envelopes based on triaxial data. Journal of Geotechnical
and Geoenvironmental Engineering 2004;130(5):498e506.
Bishop AW, Henkel DJ. The measurement of soil properties in triaxial tests. London:
Edward Arnold Ltd.; 1969.
Clough GW, Kuck WM, Kasali G. Silicate-stabilized sands. Journal of the Geotechnical
Engineering Division 1979;105(1):65e82.
Clough GW, Sitar N, Bachus RC, Rad NS. Cemented sands under static loading.
Journal of the Geotechnical Engineering Division 1981;107(6):799e817.
Clough GW, Iwabichi J, Rad NS, Kuppusamy T. Influence of cementation on liquefaction
of sands. Journal of Geotechnical Engineering 1989;115(8):1102e17.
Consoli NC, Rotta GV, Prietto PDM. Influence of curing under stress on the triaxial
response of cemented soils. Geotechnique 2000;50(1):99e105.
Consoli NC, Foppa D, Festugato L, Heineck KS. Key parameters for strength control
of artificially cemented soils. Journal of Geotechnical and Geoenvironmental
Engineering 2007;133(2):197e205.
Consoli NC, Viana da Fonseca A, Cruz RC, Heineck KS. Fundamental parameters for
the stiffness and strength control of artificially cemented sand. Journal of
Geotechnical Engineering 2009;135(9):1347e53.
Consoli NC, Cruz RC, Floss MF. Parameters controlling tensile and compressive
strength of artificially cemented sand. Journal of Geotechnical and Geoenvironmental
Engineering 2010;136(5):759e63.
Consoli NC, Cruz RC, Floss MF. Variables controlling strength of artificially cemented
sand: influence of curing time. Journal of Materials in Civil Engineering
2011;23(5):692e6.
Coop MR, Atkinson JH. The mechanics of cemented carbonate sands. Geotechnique
1993;43(1):53e67.
Cuccovillo T, Coop MR. Yielding and pre-failure deformation of structured sands.
Geotechnique 1997;47(3):481e508.
Cuccovillo T, Coop MR. On the mechanics of structured sands. Geotechnique
1999;49(6):741e60.
Amir Hamidi is an Associate Professor in the School of Engineering
at Kharazmi University of Tehran

You might also like