You are on page 1of 11

Engineering Geology 185 (2015) 5262

Contents lists available at ScienceDirect

Engineering Geology
journal homepage: www.elsevier.com/locate/enggeo

Specimen size effect on strength behavior of cemented paste backlls


subjected to different placement conditions
Erol Yilmaz a,b,, Tikou Belem a, Mostafa Benzaazoua a
a
b

Universit du Qubec en Abitibi-Tmiscamingue (UQAT), Department of Applied Sciences, 445 Boul. de l'Universit, Rouyn-Noranda, Qubec J9X 5E4, Canada
First Quantum Minerals Ltd., Cayeli Bakir Isletmeleri A.S., PO Box 42, Madenli, Cayeli, Rize TR53200, Turkey

a r t i c l e

i n f o

Article history:
Received 14 April 2014
Received in revised form 19 November 2014
Accepted 26 November 2014
Available online 5 December 2014
Keywords:
Paste backll
Scale effect
Placement conditions
Physical properties
Compressive strength

a b s t r a c t
This paper investigates the specimen size effect on the strength properties of cemented paste backll through unconned compressive strength tests. Paste backll samples were cast in different size molds (D H: 10 20 cm,
7.5 15 cm, and 5 10 cm) and subjected to the four placement conditions: cappeddrained CD; uncapped
drained UD; cappedundrained CU; and uncappedundrained UU. The unconsolidated samples were also
compared with consolidated backlls. Results show that the highest strengths were obtained from the CD samples cast in the mold 5 10 cm, followed by the UD, CU and UU samples. This could be explained by the removal of excess water within paste backll, the decreased number of micro-cracks in grains, and the wall effects.
With the drainage of excess water, particles are pulled together under capillary forces, resulting in an increase of
stiffness, accelerating binder hydration and thus, producing higher strengths. Further, specimen size effect on
water content, void ratio, mercury intrusion porosity, and degree of saturation of cemented paste backll were
presented. Some scale effect relationships were nally expressed and discussed.
2014 Elsevier B.V. All rights reserved.

1. Introduction
In recent years, new approaches and technologies have been developed to reduce dam failures and their subsequent environmental hazards, and enable mine operations to lessen the quantity of wastes
(Blanco et al., 2013). One of these approaches is the use of cemented
paste backll (CPB) with its main advantages of lower operating costs
and less wastes sent to tailings ponds (Bussire, 2007; Yilmaz, 2011).
CPB allows the disposal of sulde-rich tailings in underground stopes
avoiding largely serious environmental threats such as acid mine
drainage formation and tailings impoundment failure (Belem and
Benzaazoua, 2008; Yilmaz et al., 2014a). CPB can be described as a relatively complex composite material that consists typically of ltered wet
tailings (at a solid content of 7585 wt.%), a hydraulic binder, and water.
The main ingredients (tailings, cement, water, and seldom additives) of
CPB play a key role on its strength, its transportation and placement to
underground mined out stopes (Fall et al., 2008; Benzaazoua et al.,
2010).
A stope is dened as the site of ore production in an orebody.
After extraction of ore, the stope is ready for backlling. A steel reinforced shotcrete barricade is constructed across the undercut to

Corresponding author at: First Quantum Minerals Ltd., Cayeli Bakir Isletmeleri A.S., PO
Box 42, Madenli, Cayeli, Rize TR53200, Turkey. Tel.: +90 464 544 1544x304; fax: +90 464
544 6841.
E-mail addresses: yilmazer@fqml.com, erol.yilmaz@uqat.ca (E. Yilmaz),
tikou.belem@uqat.ca (T. Belem), mostafa.benzaazoua@uqat.ca (M. Benzaazoua).

http://dx.doi.org/10.1016/j.enggeo.2014.11.015
0013-7952/ 2014 Elsevier B.V. All rights reserved.

contain the initially uid paste backll. Paste backll is then placed
by end-of-pipe deposition from the overcut access (Fig. 1). In most
cases, initially a plug-ll of few meters high (up to 7 m) is poured
into the stope and then the residual ll is placed. The cement content
in the plug ll varies between 5 and 7 wt.% while the cement content
in the residual ll varies between 2 and 5 wt.%. The plug ll is usually
left 27 days for curing prior to the residual ll in order to avoid
excess pressure on the barricade (Belem et al., 2013). As mining
operations go ever deeper, CPB is placed into the stopes to provide
a stable platform for miners to work on and ground support for the
walls of the adjacent adits. Depending on its specic application,
CPB must have desirable mechanical properties to withstand underground stresses. In most cases, an adequate unconned compressive
strength (UCS) for a mine tailings backll containing 3 to 7 wt.% of
any binder varies between 700 and 2000 kPa (Brackebusch, 1994).
The US Environmental Protection Agency (US EPA US Environmental
Protection Agency, 1989) considers a stabilized material as satisfactory if it has a UCS of 345 kPa. However, the requirement for
minimum strength acquisition should be determined from the
design loads to which the material may be subjected. For example,
when CPB is used for underground disposal, a UCS between 150
and 300 kPa is required to eliminate the liquefaction at early ages
(Bloss, 2002). In open stoping operations, when free-standing wall
faces are exposed during pillar recovery, mechanical strengths
higher than 1000 kPa at 28 days are desired to retain CPB stability
(Belem and Benzaazoua, 2008). When CPB is used for roof support,
a UCS of higher than 4000 kPa is required (Grice, 1998).

E. Yilmaz et al. / Engineering Geology 185 (2015) 5262

53

Fig. 1. Overview of intrinsic and extrinsic factors which affect the quality and performance of cemented paste backll.

To mimic these curing and placement conditions in the stopes, CPB


samples in plastic molds are generally prepared, cured and subjected
to UCS tests in order to examine their mechanical behavior. The UCS
test is less expensive and can be easily practiced within routine quality
control programs at mine sites (Kesimal et al., 2004, 2005; Nasir and
Fall, 2008; Cihangir et al., 2012). The quality and performance of CPB
samples are greatly affected by its intrinsic and extrinsic factors
(Fig. 1): Intrinsic factors include all the parameters of tailings, cement
and water, and their interactions during curing while extrinsic factors
are those induced by the stope dimension, pasterock interaction,
placement conditions, curing temperature and time, self-weight or
time-dependent consolidation, and drainage or bleeding of excess
water. Most investigations have been conducted so far on these factors
of paste backlls (Fall et al., 2010; Yilmaz et al., 2010a; Abdul-Hussain
and Fall, 2011; Huang et al., 2011; Coussy et al., 2012; Ghirian and
Fall, 2013, 2014; Abbasy et al., 2014; Yilmaz et al., 2014b, 2015). To
assess the quality and strength of CPB samples, unconsolidated samples
with dimensions of 10 cm in diameter and 20 cm in height (based on a
height to diameter ratio of 2:1) are usually used in the laboratory
studies. Handling huge volumes of tailings to be used during testing is
time-consuming and costly. For this reason, some researchers
(Benzaazoua et al., 2004; Yilmaz et al., 2010b; Pokharel and Fall, 2013)
have used 5 10 cm plastic molds for UCS tests. Moreover, these
molds do not most often have drainage holes at the bottom and are
used as their top is capped.
Although the ultimate CPB design is based on the mechanical properties derived from lab-prepared CPBs using these conventional plastic
molds these lab studies are not particularly representative of the underground placement and curing conditions. Some works have already
shown that in situ strengths are always 24 times higher than those
obtained from laboratory strengths for a given CPB recipe and curing
time (Fourie et al., 2006; Belem et al., 2007; Grabinsky and Bawden,
2007; Yilmaz et al., 2010, 2011; Thompson et al., 2012; Dirige and
Archibald, 2014). These differences observed in the strength properties
could be explained in part by scale effects, mixing and in situ placement
techniques and curing conditions such as stope size versus mold size,
and state of CPB within the stope, such as back-lling rate, stress state,
strain rate, consolidation, arching and shrinkage effects (Li et al., 2005;

Helinski et al., 2007; Fahey et al., 2009; Belem et al., 2013). Curing
under applied pressure causes an increase in the unconned compression strength of cemented sandy soils and tailings (Consoli et al.,
2006; Yilmaz et al., 2009; Consoli and Foppa, 2014). The effect of porosity and cement content on soil strength has been recently studied by
Consoli (2014). A recent study by Festugato et al. (2013) reported the
effect of cement and bers on the cyclic strength of CPB samples.
There are in fact some problems with scaling up laboratory-prepared
CPB samples' strength results in order to determine the eld UCS properties. The difculties often lie in the selection of appropriate specimen
size and shape for laboratory tests (Hassani et al., 2007; Darlignton et al.,
2011; Ercikdi et al., 2014; Snyman et al., 2014). ASTM Standard C192/
C192M-13a (2012) shows that a minimum cylinder diameter is to exceed three times the maximum coarse aggregate size while Brown
(1981) recommends to have a cylinder size at least ten times the largest
grain size of rock. There is no standard specimen size and shape for the
UCS testing of paste backlls. Revell (2004) has shown that specimen
size plays a leading role on mechanical strength behavior of CPB. The
author found that, based on a height-to-diameter ratio of 2, small size
cylinders (4 cm in diameter) produced 610% higher compressive
strengths than large size cylinders (10 cm in diameter). As a result of
the logistical difculties associated with extracting undisturbed samples
of CPB samples and high costs of eld paste backll testing, understanding the combined effects of specimen size and placement conditions,
based on a range of underground stope congurations, is very important
in assessing in situ CPB quality and behavior over time.
In this paper, a laboratory investigation was undertaken to assess the
specimen size effects of curing and placement conditions on the
mechanical strength and physical properties (i.e. water content, degree
of saturation, void ratio and mercury intrusion total porosity) of CPB as
a function of different binder contents (3, 4.5 and 7 wt.%) and curing
times (7, 14 and 28 days). Consolidated paste backlls were tested in plastic molds of height/diameter (H/D) ratio of 2 and different sizes (D H =
10 20 cm, 7.5 15 cm, and 5 10 cm) under four placement conditions:
cappeddrained CD; uncappeddrained UD; cappedundrained CU;
and uncappedundrained UU. A newly-developed laboratory apparatus,
CUAPS (Curing Under Applied Pressure System) was used for these tests
(Benzaazoua et al, 2006; Yilmaz et al., 2010a). CUAPS allows more

54

E. Yilmaz et al. / Engineering Geology 185 (2015) 5262

condent determination of pertinent backll properties at laboratory


scale. This equipment can be used to simulate CPB placement and curing
processes, and then to assess the strength development of the samples
under consolidated and drained conditions via the application of external
pressure on the top surface of CPB specimens. Finally, simple correlations
were developed for assessing the scale effects of the CPB samples studied.
2. Materials and methods
The representative tailings material used in this study was obtained
from the LaRonde Mine's paste backll plant (Quebec, Canada) after
ltration process. In order to avoid any oxidation and prevent evaporation, mine tailings samples were stored in sealed plastic containers until
their utilization for experimental tests. Laboratory evaluation of mine tailings consisted of the determination of the grain size distribution (GSD),
solids' specic gravity, specic surface area, and mineralogical and chemical composition. To achieve the desired mechanical properties, the binder
and water were added to tailings to make various mixtures of paste backll. General use Portland cement (GU) and blast furnace slag (BFS), blended in the ratio 20/80 respectively were used as the binding agent while
tailings interstitial water was used as mixing water.
2.1. Physical and chemical properties of paste ingredients

Table 1
Physical characteristics of mine tailings.
Mine tailings
Specic gravity Gs
Specic surface area Ss (m2/g)
Fines content (b20 m; %)
D10 (effective particle size; m)
D30 (m)
D50 (average particle size; m)
D60 (m)
D90 (m)
Coefcient of uniformity Cu = D60/D10
Coefcient of curvature Cc = D302/(D60xD10)
Uniformity of graduation U = (D90-D60)/D50
Maximum dry unit weight d (kN/m3)
Optimum water content wopt (%)
Void ratio at optimum condition eopt

3.7
2.2
44
4.3
12.2
24.3
33.6
119.2
7.9
1.1
4.7
24.9
12.5
0.5

([SiO2 + Al2O3]/[CaO + MgO]) of the unhydrated cement was 0.8. A


Bruker AXS D8 Advance X-ray diffractometer equipped with a cobalt
anticathode was also used to determine the tailings' mineralogy. Quantitative mineralogical analysis was evaluated using Rietveld full-pattern
tting method with TOPAS software. The X-ray diffraction and semiquantitative analyses revealed that the most abundant mineral within
tailings was pyrite (47%), followed by quartz (32%), chlorite (9%),
paragonite (7%), muscovite (3%), talc (1%) and gypsum (1%). More
information on the characterization of tailings, binder and mixing
water can be found in Yilmaz et al. (2010b).

Grain size distribution (GSD) of the tailings sample was analyzed


using a Malvern Mastersizer laser diffraction particle size analyzer and
the obtained GSD curve was compared fairly well with those of the tailings used in Canadian mines (Fig. 2). Table 1 tabulates the physical
properties of the studied tailings. The tailings sample was found to
have a 44 wt.% fraction ner than 20 m, hence classifying as a medium
size tailings material. The sample was well-graded with a coefcient of
uniformity, Cu, of 7.9 and a coefcient of curvature, Cc, of 1.1. Based on
the Unied Soil Classication System (ASTM Standard D2487-11,
2011), the tailings sample was identied as low plasticity silt (ML).
The compaction tests, based on the ASTM standard (ASTM Standard
D1557, 2012), were also performed in the laboratory as shown in
Fig. 3. The main parameters obtained were also tabulated in Table 1.
For the tailings sample, the optimum water content, wopt, for the modied Proctor test was about 12.5 wt.%. The corresponding dry unit
weight, d, was 24.9 kN/m3. The calculated void ratio for the optimum
condition, eopt, was close to 0.5.
The chemical composition of the samples (Table 2) was determined
by a Perkin-Elmer Optima 3100 RL ICPAES (inductively coupled
plasmaatomic emission spectroscopy) after digestion. The tailings
contained 39.2% Fe2O3 and 51.3% total sulfur as determined by major
oxide and elemental analysis, respectively. The hydraulic index ratio

A total of 216 paste backll samples were prepared by blending


homogenously the tailings, binder and mixing water in an electrical
Hobart D300-1 mixer until a smooth consistency was achieved.
Depending on the viscosity of matrix, which changes greatly with the
binder content (3, 4.5 and 7% by dry mass) used in the mixtures, additional water was added to set the resulting CPB solid concentration percentage to 78 wt.%. The paste material was tested using the Abrams cone
(ASTM Standard C143/C143M, 2013 to achieve 7 slump for all CPB samples. To study the effects of specimen size and placement conditions in the
lab, samples were poured into different size molds (D H = 10 20 cm,
7.5 15 cm, and 5 10 cm) and different placement conditions (capped
drained CD, uncappeddrained UD, cappedundrained CU, and
uncappedundrained UU) for UCS testing. The cappeddrained
condition represents the placement of sample into the mold, which has
drainage holes at the bottom and a cover on the top while the
uncappeddrained condition refers to the placement of sample in the

Fig. 2. Grain size distribution (GSD) curve of mine tailings sample compared to a typical
range of GSD curves of eleven tailings sampled from Canadian hard rock mines.

Fig. 3. Typical moisturedensity curve obtained from laboratory compaction tests for the
mine tailings sample, accompanying with zero-air-voids curve.

2.2. Preparation and testing of paste mixtures

E. Yilmaz et al. / Engineering Geology 185 (2015) 5262


Table 2
Chemical composition of mine tailings, mixing waters and binder materials.

Al2O3
B2O3
BaO
CaO
Fe2O3
K2O
MgO
MnO
Na2O
Sc
SiO2
TiO2
ZnO

Detection
limit
(wt.%)

Mine tailings
(wt.%)

Tailings
pore water
(mg/L)

GUa
alone
(%)

BFSb
alone
(%)

GU:BFS =
20:80 wt.%
(%)

0.01
0.001
0.001
0.03
0.006
0.03
0.001
0.002
0.03
0.09
0.02
0.002
0.005

5.29
0.61
0.01
0.79
39.20
0.24
0.16
0.02
0.40
51.34
1.07
0.07
0.44

0.40
0.14
0.05
782.16
0.02
2.77
3.03
0.01
455.61
4069.88
1.91
n/a
0.13

4.86
0.03
0.01
65.76
2.44
0.83
2.21
0.04
2.11
3.67
19.51
0.26
0.03

10.24
0.48
0.09
31.41
0.55
0.51
11.29
0.97
2.01
3.27
36.22
0.61
n/a

8.39
0.13
0.03
42.82
0.64
0.55
6.19
0.18
2.03
3.35
30.91
0.48
n/a

General use Portland cement.


Ground granulated blast furnace slag.
c
Total sulfur was assumed expressed exclusively as SO2
4 (= 2.9956 S, based on
stochiometric conversion).
b

bottom-perforated mold without cover on the top. The cappedundrained placement condition, frequently used for quality control tests
at both laboratory and mine sites, represents the capped mold with
no drainage holes at the bottom. The uncappedundrained placement
condition represents the open-top mold with drainage holes at the
bottom.
In comparison with the unconsolidated samples a total of 108
consolidated-drained CPB samples, obtained using the CUAPS apparatus
were prepared for the prescribed curing times of 7, 14, and 28 days and
binder contents of 3, 4.5, and 7 wt.%. For the drained scenarios, 2-mm diameter holes were drilled through the bottom of each mold, 31 holes for
the 10 20 cm mold, 9 holes for the 7.5 15 cm mold, and 2 holes for
the 5 10 cm. Non-woven geotextile lter (AtexTEXEL Model 760912) was placed at the bottom to allow bleed water drainage. Paste samples were cast into the molds in one-third increments. After the mold
was lled, the paste was rammed in 25 blows using a small steel rod
in order to eliminate any large trapped air bubbles within CPB, as described in the ASTM C143 standard. After sealing, the molds were stored
in a humidity chamber set at 80% relative humidity and 25 C temperature (similar to underground conditions) over the desired curing times.
Fig. 4 shows the different size molds and the conditions conducted during the experimental tests.

55

2.3. Mechanical, physical and pore structure characterization tests


Following the curing periods of up to 28 days, the unconsolidated
backlls were extracted from the drained-, undrained-, capped-, and
uncapped plastic molds using compressed air while CUAPS-consolidated
samples were extruded from the Perspex mold using an adequate manual
push sample extruder. To determine the mechanical properties, the
extracted CPB samples were then subjected to unconned compression
tests (ASTM Standard C39/C39M, 2012). A servo-controlled press (MTS
Sintech 10/GL) that has a normal loading capacity of 50 kN and a uniform
loading rate of 1 mm/min was used. Prior to UCS testing, the two ends of
the CPB samples were rectied to ensure a uniform distribution of the
applied load. The mean values from triplicate tests were presented in
the results.
Some physical properties such as gravimetric water content wg,
degree of saturation Sr, dry density d and void ratio e were also determined on representative CPB samples after UCS testing. A required
portion of each specimen was oven-dried for about 3 days at 45 C in
order to determine the water content of the CPB sample. Solids' specic
gravity was determined by a Micromeritics AccuPyc 1330 helium
pycnometer. Using these data (e.g. wg, d and Gs), the remaining
physical parameters, including void ratio and degree of saturation
were calculated from the massvolume relationships. The physical
properties were evaluated only on the specimen size which gave the
highest mechanical strengths. In addition, the CPB's micro-structural
properties were determined using a mercury intrusion porosimeter
MIP (Micromeritics Autopore III 9420), which allows measurement of
a minimum accessible radius of 0.003 m. After UCS tests, representative CPB samples (D H: 12 24 mm) were taken from the middle
part of paste backll specimen which avoids stress concentration
effects. CPB samples were oven-dried at 45 C for at least 96 h and
then stored in a desiccator over silica gel to minimize pore alteration
due to hydration product destruction and moisture ingress. For each
sample, two MIP analyses were performed to obtain an average value.
3. Experimental results
3.1. Effect of specimen size on strength development of cemented paste
backll
Fig. 5 shows that specimen size played a key role on the strength
gain of CPB samples as a function of binder content and curing time.
For the research of the effect of specimen size on strength, a total
of four different CPB materials, cappeddrained CD, uncapped

Fig. 4. Photos illustrating the different size conventional plastic molds and test conditions: (a) empty molds, (b) molds lled with CPB mixtures and cured in the humidity chamber,
(c) cappeddrained CPB specimen cured within CUAPS, (d) CPB specimen under MTS (Material Testing Systems) Sintech 10/GL computer-controlled 50 kN testing machine, and
(e) CPB specimen being broken after UCS testing.

56

E. Yilmaz et al. / Engineering Geology 185 (2015) 5262

Fig. 5. Change in compressive strength with curing time for cappeddrained CD, uncappeddrained UD, cappedundrained CU, and uncappedundrained UU paste backll samples
cast in different size molds and prepared at binder contents of (a) 3 wt.%, (b) 4.5 wt.%, and (c) 7 wt.%.

drained UD, cappedundrained CU, and uncappedundrained U


U were studied. Fig. 5 shows the change in the compressive strength
with curing time for CD samples. For all the CPB samples tested, the
most suitable specimen size in connection with the rate of strength
development and the ultimate strengths was found to be D H =
5 10 cm, which provided higher UCS values for a given paste backll recipe, in comparison with the other specimen sizes of 10 20 cm
and 7.5 15 cm. For example, when CPB samples with 3 wt.% cement
content were cast in the mold 5 10 cm, the highest UCS of about
1605 kPa was generated in these samples after curing of 28 days.
Similarly, the plastic molds 7.5 15 cm and 10 20 cm gave strength
values of 1377 and 1131 kPa, respectively.
However, the CUAPS-consolidated samples always provided higher
strengths than those obtained from plastic molds. At 6 wt.% cement content, the strength acquisition of consolidated samples cured at 28 days
was 2150 kPa. Table 3 tabulates the percent-based mechanical strength
differences of consolidated samples cast in the mold 10 20 cm when
compared to the same size unconsolidated samples. Fig. 5 also shows
that, the UCS increases clearly with increasing the binder dosage for a
given curing time. It should be also noted that the CD samples meet
the lowest recommended UCS for both the reusability of treated waste
material and eliminating the risk of liquefaction. The mold 5 10 cm
gave 18% and 39% higher strength values than the molds 7.5 15 cm
and 10 20 cm for 3 wt.% binder. These values were 16% and 31% for
4.5 wt.% binder and 11% and 19% for 7 wt.% binder. The increase in the
strength as the size of the sample decreased at 3 wt.% binder was higher
than at higher binder contents (4.5 and 7 wt.%).
3.2. Effects of placement conditions on strength development of cemented
paste backll
In most cases, the water loss during lling and curing of the backll at
most modern underground mines can be explained by the consolidation,
drainage and evaporation processes, compaction, and self-desiccation

(i.e. water consumed by cement hydration). These are affected by the


stope geometry and the underground stope congurations. In this
study, only the effects of evaporation and drainage via the plastic
molds were studied. Fig. 6 shows the development of UCS with time
for different CPB samples cast in a mold size of D H = 10 20 cm,
7.5 15 cm, and 5 10 cm, and prepared at 3, 4.5 and 7 wt.% binder dosages. These different drainage congurations appeared to affect the
strength and stability, but the specimen size also played a key role in
the strength development of CPB.
From Fig. 6, it can be also seen that drained paste backlls generate better mechanical strengths than undrained paste backlls for a
given CPB recipe and curing time. This is due the fact that the drained

Table 3
The improvement of strength of consolidated samples compared with unconsolidated
samples (in percent).
Bw = 4.5 wt.%

Bw = 7 wt.%

Cappeddrained (CD) samples


7 days
63%
14 days
55%
28 days
47%

57%
48%
41%

41%
36%
34%

Uncappeddrained (UD) samples


7 days
71%
14 days
62%
28 days
58%

64%
55%
51%

46%
42%
38%

Cappedundrained (CU) samples


7 days
74%
14 days
70%
28 days
65%

70%
61%
55%

53%
49%
45%

Uncappedundrained (UU) samples


7 days
81%
14 days
78%
28 days
68%

76%
64%
59%

64%
56%
49%

Curing time

Bw = 3 wt.%

E. Yilmaz et al. / Engineering Geology 185 (2015) 5262

57

Fig. 6. Change in compressive strength with curing time for different CPB samples cast in a mold size of D H = 10 20 cm, 7.5 15 cm, and 5 10 cm, and prepared at binder contents of
(a) 3 wt.%, (b) 4.5 wt.%, and (c) 7 wt.%.

backlls have less water-to-cement w/c ratios, which favor the acceleration of cement hydration reactions with higher strengths. The results indicated are also conrmed for CPB samples cast in the mold
sizes D H = 7.5 15 cm and D H = 10 20 cm (Fig. 6). It is fairly
apparent that the highest uniaxial compressive strengths were obtained from the CD samples. These samples are compared in percent with the other samples (see Table 3). It can be inferred from
these ndings that the drained CPB samples have a lesser amount
of porosity. This is most often due to the fact that the water loss by
drainage gives rise to the settling of the backll and the resultant reduction of the CPB void ratio. Hence, higher strengths are achieved
when compared with the undrained backlls.
Fig. 7 presents the evolution of the mechanical strengths of CPB
prepared with 4.5 wt.% of the binder type GU-Slag@20:80 wt.% for
different curing times. It should be noted that the curing condition
used in most laboratories is the CU condition. From Fig. 7, it can be
seen that the higher mechanical strengths were obtained from the
small size specimen (d = 5 cm), regardless of curing condition, binder
content and curing time. These ndings clearly suggest the existence
of specimen size effect on the resultant strength gain of CPB samples.
3.3. Investigation of the physical parameters of the optimal paste backll
recipe
The mechanical properties of any paste backll material are
greatly affected by its physical (bulk or geotechnical index) properties, such as water content, void ratio, wet unit weight or specic
gravity, and degree of saturation. It is known that the physical properties that develop within CPB are affected by a combination of i) the

physical, chemical and mineralogical properties of paste ingredients;


ii) binder type and content; iii) curing conditions (with or without
effective stress) including temperature and time; iv) mixing and
placement techniques; and v) internal stresses occurred in the backll material based on self-weight or time-dependent consolidation
loadings.
In this study, the evaluation of physical property is made only on the
most ideal CPB recipe cast in the mold 5 10 cm, which always gives
higher strengths than other mold sizes. Fig. 8 shows the change in gravimetric water content wg of CPB. The overall observation is that the
water contents of the drained paste backlls are slightly lower than
the undrained ones, irrespective of the evaporation state. When the
binder dosage increased from 3 to 7 wt.%, the water content was
reduced mainly due to the initial w/c ratios and amounts of water
required for cement hydration. This suggests that, the higher the
cement content, the lower the water content is.
Fig. 8 shows the change in the void ratio e of different CPB samples as
a function of binder content and curing time. The drained CPB samples
showed smaller void ratios than the undrained ones because the
water loss by drainage would give rise to the settling of paste backll
and the resulting reduction of the void ratio within the backll.
Moreover, Fig. 8 illustrates that the drained CPB samples were less
saturated than the undrained ones at a given binder content and/or
curing time. The range of variation of degree of saturation Sr for the
undrained backlls was relatively low (99.594% for the UU samples
and 9992.5% for the CU samples). After curing of 28 days, the Sr of
CPB samples at 3, 4.5 and 7 wt.% binder contents was determined to
be 96.5%, 94.5% and 90.5%, respectively, indicating the decrease in
degree of saturation with increasing the binder content.

58

E. Yilmaz et al. / Engineering Geology 185 (2015) 5262

Fig. 7. Specimen size effect on the UCS of CPB prepared with 4.5 wt.% of binder type GU-Slag@20:80 wt.% cured in four different placement conditions: a) capped and drained CD;
b) uncapped and drained UD; c) capped and undrained CU; d) uncapped and undrained UU.

Fig. 8. Change in gravimetric water content, void ratio, and degree of saturation with curing time for CPB samples cast in a mold size of D H = 5 10 cm and prepared at binder contents
of (a) 3 wt.%, (b) 4.5 wt.%, and (c) 7 wt.%.

E. Yilmaz et al. / Engineering Geology 185 (2015) 5262

59

4. Empirical correlations of results


Fig. 10 shows a plot of the normalized compressive strength UCS*
(= UCS/d) versus the sample diameter d. The normalized UCS* is higher
for the small specimen diameter and decreases noticeably as the
diameter increases up to 10 cm). In the following subsections, two
models based on the specimen diameter d and reference diameter dref
were developed to predict the resulting UCS. Model 1 was expressed
as a function of d (based on curing time t, and binder content Bw)
while Model 2 was as a function of dref (based on diameter: 5, 7.5 or
10 cm).
4.1. General empirical model relating UCS to the specimen diameter d
Fig. 11 presents the correlation between mechanical strength and
the specimen diameter: a) raw data of UCS versus specimen diameter;
and b) normalized UCS versus the normalized time factor tn. The
correlation was based on a general power law: X = tn = (0.5 + t)/
(0.5 + Bw%); and Y = (tn)^3 (UCS/d)^ko, d in cm and UCS in kPa.
The plot of X against Y is given by Fig. 11b. The data are well tted
with a power law Y = aX^b where a = 1.1323, b = 3.0031 and ko =
0.025 (= 1/40) with r2 = 0.9999 (r = 1). The nal relationship
representing the UCS as a function of specimen diameter d, curing
time t and binder content Bw% is given as follows (Eqs. 1 and 2):
"

 #1


1 2t b3 k0
k0
UCS d; t; Bw% a  d 
1 2Bw%

or
"

 #1


1 2t k1 k0
k
UCS d; t; Bw% a  d 0 
1 2Bw%
Fig. 9. Cumulative (a) and incremental (b) mercury intrusion pore size distribution of 28day cured CPB samples subjected to different placement conditions (cappeddrained CD,
uncappeddrained UD, cappedundrained CU, and uncappedundrained UU) for
4.5 wt.% cement content.

where k1 = b3 = 3.00313 = 0.0031.


4.2. General empirical model relating the UCS to a reference diameter dref

Fig. 9 shows the cumulative and incremental pore size distribution


of the 28-day cured CPB samples prepared with the 4.5 wt.% binder.
The mercury intrusion porosity of CUAPS-consolidated backlls was
lower (38.2%) than that of the samples consolidated in any other plastic
molds (44.8% for CD samples, 45.8% for UD samples, 46.2% for CU
samples, and 47.4% for UU samples). It is apparent that the porosity
of undrained samples is greater than that of CUAPS-consolidated backlls due to the fact that the hydrationprecipitation appears to occur
to a lesser extent in the consolidated samples.

If the specimen size used in a laboratory is the smallest one d = 5 cm,


then a higher UCS/d ratio will be obtained when compared to the specimen diameter of d = 7.5 cm and d = 10 cm. On the contrary if the specimen diameter used in the experiments is d = 10 cm then the UCS/d
ratio will be the lowest. As no reference diameter does exist, any of
the three usual specimen diameters can be used as a reference diameter
depending on the situation. Therefore, it can be helpful to get even an
empirical relationship relating the UCS of a specimen having a diameter
d to the UCS of the specimen having the reference diameter dref (can be
5, 7.5 or 10 cm).
Fig. 12 presents the correlations between the mechanical strengths
and the specimen diameter: a) raw data of UCS versus the specimen
diameter; b) normalized UCS and diameter with the reference diameter
d = 10 cm; c) normalized UCS and diameter with the reference
diameter d = 7.5 cm; and d) normalized UCS and specimen diameter
with the reference diameter d = 5 cm. The variables of the correlation
determined using a general power law tting are as follows:
h

ik

t n 0:5 t=0:5 Bw ; X t n 6= dre f =d ; and Y


o
h

ik

t n 6  UCSd=UCS dre f
o

Fig. 10. Normalized UCS vs. specimen diameter for different placement/curing conditions
for CPB prepared with 7 wt.% of the binder type GU-Slag@20:80 wt.% at 28 days of curing.

The plot of X against Y is given in Figs. 12bd. The data are well tted
with a power law Y = aX^b. The nal relationship relating the UCS of a
specimen of diameter d to the UCS of a specimen of the reference

60

E. Yilmaz et al. / Engineering Geology 185 (2015) 5262

Fig. 11. Correlation between UCS and the specimen diameter: a) raw data of UCS versus
specimen diameter, b) normalized UCS versus the normalized time factor tn.

diameter dref as a function of curing time t and binder content Bw% is


given as follows (Eq. 3):


UCS d; t; Bw% a  UCSdre f 

1 2t
1 2Bw%

61k

0 b



dre f k0 b
d

where a, b = the tting constants, UCSdref = the unconned compressive strength corresponding to the reference diameter dref (kPa), d =
the target specimen diameter (cm), dref = the reference diameter
(cm), and ko = 0.125 (= 1/8). The above equation can be rewritten as
follows (Eq. 4):


UCS d; t; Bw% a  UCSdre f 

1 2t
1 2Bw%

61b 
 b
8
dre f 8

d

For the three possible reference specimen diameters and for the
ranges of curing conditions covered by the data in the present study
the tting constants are given in Table 4 along with the coefcient of
correlation r.
5. Discussion
In general, one of the reasons behind the large differences observed
between the mechanical strength properties of lab-prepared CPB samples and in situ CPB materials is regarded to be the specimen size and
scale effects. Understanding the scale effects is of great important for designing and modeling the overall quality and behavior of CPB. A number
of limited works including projects undertaken by several mining operations have been done on the scale effects of CPB samples. However,
many of these investigations have not been published because they
are condential, and the data for this aspect are not readily available
in the literature. The published data are specic to the individual test

mines and indicate the framework for developing a future in situ testing
study.
For example, Revell (2004) did an experimental work to identify the
impact of size on the ultimate mechanical strengths of CPB. The author
studied two cylindrical core samples: D H = 10 20 cm, and 4 8 cm,
which gives a height-to-diameter ratio of 2. It was shown the development of 6 to 10% higher strengths by the 4 cm diameter cylinders than
those by the 10 cm diameter cylinders. Moreover, Hassani et al. (2007)
conducted a laboratory work on three types of mine lls: one blended
tailingssand ll, a compositeaggregate paste ll (CAP), and rock ll.
For all the CPB mixes, the mechanical strength decreased with increasing specimen size. However, for the two mixtures that contain a blend
of full-plant precious metal tailings and either 50% or 75% sand, the
strength increased up to a cylinder diameter of 15 cm. In the rock ll,
the strength decreased with increasing specimen size irrespective of
binder content. The CAP specimens showed minimal strength dependence on specimen size.
It can be inferred from the present work that the CU samples at
3 wt.% binder content could fail to maintain the structural stability of
CPB at most modern underground mines when free-standing wall
faces are exposed during pillar recovery. The rest of the CU samples
can be used for several purposes at underground mines such as a
construction material, a major ground-support tool and a primary
mine tailings disposal method. The reason for higher strength of the
consolidated samples may be explained by the drainage of water and
the efcient packing of cementitious products. This would lead to the
improved CPB pore structure and a reduction in saturated hydraulic
conductivity. Both grain and pore renements also reduce the porosity
of CPB. These ndings are consistent with those earlier reported
(Ouellet et al., 2007; Yilmaz et al., 2011, 2014b; Belem et al., 2013;
Ghirian and Fall, 2014).
Results have also shown that the drained paste backlls (CD and U
D) continually produce higher mechanical strengths than the undrained
paste backlls (CU and UU). This could be well explained by the
removal of excess water within the backll, which helps accelerating
binder hydration and thus increasing the strength and stability of CPB.
In addition, the best UCS values were obtained from CD samples.
Additionally, the overall UCS results conrm that the UD samples can
be effectively used in the cut and ll mining, because, they offered
compressive strengths higher than 1000 kPa for a curing period of 28
days. However, the UD samples cannot be used for roof support
(based on the service life and structural integrity of CPB) at underground mines because the mechanical strengths obtained were under
the target strength of 4000 kPa (Grice, 1998).
Physical property tests indicate that the water content of the drained
paste backlls was slightly lower than the undrained ones. The water
content is reduced mainly due to the initial w/c ratios and amounts of
water required for cement hydration. There is no notable change in
the Gs values of the backll. The drained paste backlls show smaller
void ratios than the undrained ones because the water loss by drainage
gives rise to the settling of paste backll (largely increasing of the corresponding density) and the reduction of the void ratio within the backll.
The Sr of CPB samples having a binder content of 3, 4.5 and 7 wt.% is
96.5%, 94.5% and 90.5%, respectively. The higher the binder content,
the lower the degree of saturation. Mercury porosimeter results indicate
that the total porosity of the undrained paste backlls was greater than
that of the CPB samples consolidated under effective stress due to the
fact that the hydrationprecipitation appears to occur to a lesser extent
in the undrained samples. Application of effective stress during curing
improves CPB's pore structure, and causes lower porosities in comparison with other unconsolidated samples.
6. Conclusions
This paper presents the results of an experimental study aimed at
examining the effects of specimen sizes and placement conditions on

E. Yilmaz et al. / Engineering Geology 185 (2015) 5262

61

Fig. 12. Correlations between UCS and the specimen diameter: a) raw data of UCS versus diameter, b) normalized UCS and diameter with the reference diameter d = 10 cm, c) normalized
UCS and diameter with the reference diameter d = 7.5 cm, d) normalized UCS and diameter with the reference diameter d = 5 cm.

mechanical (i.e. unconned compressive strength) and physical (i.e.


water content, degree of saturation and void ratio) properties of the
CPB materials prepared using a slag blend binder consisting of 20 wt.%
general use Portland cement and 80 wt.% blast furnace slag. A total of
432 CPB samples were produced at a binder content of 3, 4.5 and
7 wt.% in different placement conditions: cappeddrained CD;
uncappeddrained UD; cappedundrained CU; and uncapped
undrained UU, and cast in three different size plastic molds (D H:
5 10 cm, 7.5 15 cm, and 10 20 cm) over three curing times
(7, 14 and 28 days).
The results from the experiments have shown that the diameter of
the specimen greatly affects the overall mechanical strength for a
constant height-to-diameter ratio. The most suitable specimen size
with regards to the rate of strength development and the ultimate
strengths was found to be D H = 5 10 cm, which constantly provided higher strength, regardless of the placement conditions, compared to
the other specimen sizes 7.5 10 cm and 10 20 cm. This is expected
owing to the small volume of specimen which provides the decreased
number of micro-cracks and pores in the homogenous matrix as compared with the larger specimen sizes. One of the major factors affecting
the strength and stability of CPB is the placement and curing conditions
entreated right after pouring the paste into the molds. The test work has
conrmed that the drained paste backlls always produce higher
strengths than the undrained ones as a result of the removal of excess
water within CPB, which helps accelerating binder hydration. Since in
situ backll conditions are different from those of laboratory-prepared
backlls, the best placement condition out of different placement regimes (e.g., CD, UD, CU and UU conditions) can be determined
Table 4
Fitting parameters for different reference diameters.
Reference diameter, dref

r2

Equation numbers

5 cm
7.5 cm
10 cm

0.7669
0.8901
1.3451

7.9690
7.9588
7.9644

0.9955
0.9956
0.9957

0.998
0.998
0.998

Fig. 12(a)
Fig. 12(b)
Fig. 12(c)

consistent with the achieved UCS results. It is generally expected from


a paste-backlled stope condition that the porosity decreases and the
strength increases with increasing depth into the placed paste backlls.
This is principally due to the drainage and consolidation occurred
during backlling. The acquisition of strength for paste backlls
becomes lower under the conditions of no drainage than drainage, as
shown by the test results.
There is no standard backll recipe and specimen size for the determination of the UCS of CPB materials. This work demonstrates the prime
importance of specimen size and placement conditions for the strength
development of CPB and hence for designing the most ideal paste
backll mixtures in the laboratory and operating a paste backll plant
in the eld. Notwithstanding this, further testing is still required to
acquire more consistent results for specimen size for the determination
of UCS of paste backll samples, principally on the paste cored samples
from paste-backlled stopes. This will provide evaluation of the in situ
mechanical strengths on which the eld backll samples are based.

Acknowledgments
The authors would like to express their sincere thanks and appreciation to the Natural Sciences and Engineering Research Council of Canada
(NSERC) Discovery Grant Program, Industrial NSERC-PolytechniqueUQAT Chair on Environment and Mine Wastes Management and Canadian Research Chair on Integrated Management of Mine Wastes for their
generous nancial support. Special thanks are extended to Nil Gaudet,
David Bouchard, Yvan Poirier and Alain Perreault of Unit de Recherche
et de Service en Technologie Minrale (URSTM Laboratory) for their
technical support in conducting a number of physical, mechanical and
microstructural tests, to Nathan Mutch of Lafarge North America Inc.
for providing the cement materials, to Professor Haci Deveci of Karadeniz
Technical University for his useful contributions on the earlier version of
this manuscript, and to two anonymous reviewers for their constructive
and helpful comments that signicantly improved the quality of the
manuscript.

62

E. Yilmaz et al. / Engineering Geology 185 (2015) 5262

References
Abbasy, F., Hassani, F.P., Madiseh, S.G., Ct, J., Nokken, M.R., 2014. An experimental study
on the effective parameters of thermal conductivity of mine backll. Heat Tran. Eng.
35 (13), 12091224.
Abdul-Hussain, M., Fall, M., 2011. Unsaturated hydraulic properties of cemented tailings
backll that contains sodium silicate. Eng. Geol. 4 (21), 288301.
ASTM Standard C143/C143M, 2013. Standard Test Method for Slump of Hydraulic Cement
Concrete. Annual Book of ASTM Standards, 04.02, West Conshohocken, PA http://dx.
doi.org/10.1520/C0192_C0192M.
ASTM Standard C192/C192M-13a, 2012. Standard Practice for Making and Curing Concrete Test Specimens in the Lab. Annual Book of ASTM Standards, 04.02, West
Conshohocken, PA http://dx.doi.org/10.1520/C0143_C0143M-12.
ASTM Standard C39/C39M, 2012. Standard Test Method for Compressive Strength of Cylindrical Concrete Specimens. Annual Book of ASTM Standards, 04.02, West
Conshohocken,PA http://dx.doi.org/10.1520/C0039_C0039M-12A.
ASTM Standard D1557, 2012. Standard Test Methods for Laboratory Compaction Characteristics of Soil Using Modied Effort (56,000 ft-lbf/ft3 (2,700 kN-m/m3)). Annual
Book of ASTM Standards, 04.08, West Conshohocken, PA http://dx.doi.org/10.1520/
D1557-12.
ASTM Standard D2487-11, 2011. Standard Practice for Classication of Soils for Engineering Purposes. Annual Book of ASTM Standards, 04.08, West Conshohocken, PA http://
dx.doi.org/10.1520/D2487-11.
Belem, T., Benzaazoua, M., 2008. Design and application of underground mine paste backll technology. Geotech. Geol. Eng. 26 (2), 147174.
Belem, T., El Atar, O., Benzaazoua, M., Bussire, B., Yilmaz, E., 2007. Hydrogeotechnical and
geochemical characterization of column consolidated cemented paste backll. Proceedings of the 9th International Symposium in Mining with Backll. Canadian Geotechnical Society, Montreal, Quebec, Canada, pp. 110 (April 29 May 2).
Belem, T., Benzaazoua, M., El Aatar, O., Yilmaz, E., Belem, T., Benzaazoua, M., El Aatar, O.,
Yilmaz, E., 2013. Effect of drainage and the pore water pressure dissipation on the
backlling sequencing. 23rd World Mining Congress, Canada, pp. pp. 110 (August
11-15).
Benzaazoua, M., Fall, M., Belem, T., 2004. A contribution to understanding the hardening
process of cemented pastell. Miner. Eng. 17 (2), 141152.
Benzaazoua, M., Belem, T., Yilmaz, E., 2006. Novel lab tool for paste backll. Can. Min. J.
127 (3), 3132.
Benzaazoua, M., Peyronnard, O., Belem, T., Fried, E., Stephant, D., Dublet, G., 2010. Key issues related to behavior of binders in cemented paste ll. The 13th Paste and Thickened Tailings. Australian Centre for Geomechanics, Canada, pp. 345364.
Blanco, A., Lloret, A., Carrera, J., Olivella, S., 2013. Thermo-hydraulic behaviour of the vadose zone in sulphide tailings at Iberian Pyrite Belt: waste characterization, monitoring and modelling. Eng. Geol. 165, 154170.
Bloss, M., 2002. Below ground disposal (mine backll). In: Jewell, R.J., Fourie, A.B., Lord,
E.R. (Eds.), Paste and Thickened Tailings: A Guide. University of Western Australia,
Nedlands, pp. 103126.
Brackebusch, F.W., 1994. Basics of paste backll systems. Min. Eng. 46 (10), 11751178.
Brown, E.T., 1981. Rock Characterization, Testing and Monitoring. ISRM Suggested
Methods, Oxford.
Bussire, B., 2007. Hydrogeotechnical properties of underground hard rock tailings from
metal mines and emerging environmental disposal approaches. Can. Geotech. J. 44
(9), 10191052.
Cihangir, F., Ercikdi, B., Kesimal, A., Turan, A., Deveci, H., 2012. Utilisation of alkaliactivated blast furnace slag in paste backll of sulphide mill tailings: effect of binder
type and dosage. Miner. Eng. 30, 3343.
Consoli, N.C., 2014. A method proposed for the assessment of failure envelopes of
cemented sandy soils. Eng. Geol. 169, 6168.
Consoli, N.C., Foppa, D., 2014. Porosity/cement ratio controlling initial bulk modulus and
incremental yield stress of an articially cemented soil cured under stress. Gotech.
Let. 4, 2226.
Consoli, N.C., Rotta, G.V., Prietto, P.D.M., 2006. Yieldingcompressibilitystrength relationship for an articially cemented soil cured under stress. Geotechnique 56 (1), 6972.
Coussy, S., Benzaazoua, M., Blanc, D., Mokowicz, P., Bussire, B., 2012. Assessment of arsenic immobilization in synthetically prepared cemented paste backll specimens. J.
Environ. Manag. 93 (1), 1021.
Darlignton, W.J., Ranjith, P.G., Choi, S.K., 2011. The effect of specimen size on strength and
other properties in lab testing of rock and rock-like cementitious brittle materials.
Rock Mech. Rock. Eng. 44, 513529.
Dirige, A.P.E., Archibald, J.F., 2014. Investigation of the inuence of homogeneous mixing
on the strength behavior of paste backll. Can. Geotech. J.
Ercikdi, B., Yilmaz, T., Kulekci, G., 2014. Strength and ultrasonic properties of cemented
paste backll. Ultrasonics 54 (1), 195204.
Fahey, M., Helinski, M., Fourie, A., 2009. Some aspects of the mechanics of arching in
backlled stopes. Can. Geotech. J. 46, 13221336.
Fall, M., Benzaazoua, M., Saa, E.G., 2008. Mix proportioning of underground cemented tailings backll. Tunn. Undergr. Space Technol. 23 (1), 8090.
Fall, M., Clestin, J.C., Pokharel, M., Tour, M., 2010. A contribution to understanding the
effects of curing temperature on mechanical properties of cemented tailings backll.
Eng. Geol. 114, 397413.

Festugato, L., Fourie, A., Consoli, N.C., 2013. Cyclic shear response of bre-reinforced
cemented paste backll. Gotech. Let. 3, 512.
Fourie, A., Helinski, M., Fahey, M., 2006. Filling the gapa geomechanics perspective.
Newsletter 26, 1824.
Ghirian, A., Fall, M., 2013. Coupled thermo-hydro-mechanical-chemical behavior of
cemented paste backll in column experiments. Part I: Physical and thermal processes and characteristics. Eng. Geol. 164, 195207.
Ghirian, A., Fall, M., 2014. Coupled thermo-hydro-mechanicalchemical behaviour of
cemented paste backll in column experiments. Part II: Mechanical, chemical and microstructural processes and characteristics. Eng. Geol. 170, 1123.
Grabinsky, M.W., Bawden, F.W., 2007. In situ measurements for geomechanical design of
cemented paste backll systems. CIM Bull. 100 (1103), 1828.
Grice, T., 1998. Underground Mining with Backll. Proceedings of the 2nd Annual Summit
on Mine Tailings Disposal Systems. Australian Mining Consultants, Brisbane,
Australia, pp. 114 (November 2425).
Hassani, F.P., Nokken, M.R., Annor, A., 2007. Physical and mechanical behavior of various
combinations of mine ll materials. CIM Bull. 100 (1101), 1826.
Helinski, M., Fahey, M., Fourie, A.B., 2007. Numerical modelling of cemented paste backll
deposition. J. Geotech. Geoenviron. Eng. ASCE 13 (10), 13081319.
Huang, S., Xia, K., Qiao, L., 2011. Dynamic tests of cemented paste backll: effects of strain
rate, curing time, and cement content on compressive strength. J. Mater. Sci. 46 (15),
51655170.
Kesimal, A., Yilmaz, E., Ercikdi, B., 2004. Evaluation of paste backll mixtures consisting of
sulphide-rich mill tailings and varying cement contents. Cem. Concr. Res. 34 (10),
18171822.
Kesimal, A., Yilmaz, E., Ercikdi, B., Alp, I., Deveci, H., 2005. Effect of properties of tailings
and binder on the short and long terms strength and stability of cemented paste
backll. Mater. Let. 59 (28), 37033709.
Li, L., Aubertin, M., Belem, T., 2005. Formulation of a three dimensional analytical solution
to evaluate stress in backlled vertical narrow openings. Can. Geotech. J. 42 (6),
17051717.
Nasir, O., Fall, M., 2008. Shear behaviour of cemented pastellrock interfaces. Eng. Geol.
101 (34), 146153.
Ouellet, S., Bussire, B., Aubertin, M., Benzaazoua, M., 2007. Microstructural evolution of
cemented paste backll: mercury intrusion porosimetry test results. Cem. Concr.
Res. 37 (12), 16541665.
Pokharel, M., Fall, M., 2013. Combined inuence of sulphate and temperature on the saturated hydraulic conductivity of hardened cemented paste backll. Cem. Concr.
Compos. 38, 2128.
Revell, M.B., 2004. PasteHow Strong Is It? Proceedings of the 8th International Symposium on Mining with Backll. The Nonferrous Metal Society of China, Beijing, China
(September 1921).
Snyman, B.J., van der Spuy, B., Correia, L.D.C., 2014. A critical look at uniaxial test procedures applied in the backll industry. Proceedings of the 11th International Symposium on Mining with Backll. Australian Centre for Geomechanics, Perth, Australia
(May 2022).
Thompson, B.D., Grabinsky, M.W., Bawden, W.F., 2012. In-situ measurements of
cemented paste backll at the Cayeli Mine. Can. Geotech. J. 49 (7), 755772.
US EPA US Environmental Protection Agency, 1989. Stabilization/solidication of CERCLA
and RCRA wastes: physical tests, chemical testing procedures, technology screening,
and eld activities, EPA/625/6-89/022. Risk Reduction Engineering Laboratory Cincinnati, USA.
Yilmaz, E., 2011. Advances in reducing large volumes of environmentally harmful mine
waste rocks and tailings. Mineral Resour. Manag. 27 (2), 89112.
Yilmaz, E., Benzaazoua, M., Belem, T., Bussire, B., 2009. Effect of curing under pressure on
compressive strength development of cemented paste backll. Miner. Eng. 22
(910), 772785.
Yilmaz, E., Investigating the hydrogeotechnical, microstructural properties of cemented
paste backlls using the versatile CUAPS apparatus, 2010a. Ph.D. Thesis, Universit
du Qubec en Abitibi-Tmiscamingue UQAT, Rouyn, Rouyn-Noranda, Qubec,
Canada (http://depositum.uqat.ca/34/1/erolyilmaz.pdf).
Yilmaz, E., Belem, T., Benzaazoua, M., Bussire, B., 2010b. Assessment of the modied
CUAPS apparatus to estimate in situ properties of cemented paste backll. Geotech.
Test. J. 33 (5), 351362.
Yilmaz, E., Belem, T., Bussire, B., Benzaazoua, M., 2011. Relationships between microstructural properties and compressive strength of consolidated and unconsolidated
cemented paste backll. Cem. Concr. Compos. 33 (6), 702715.
Yilmaz, E., Benzaazoua, M., Bussire, B., Pouliot, S., 2014a. Inuence of disposal congurations on hydrogeological behaviour of sulphidic paste tailings: a eld experimental
study. Int. J. Miner. Process. 131, 1225.
Yilmaz, E., Belem, T., Benzaazoua, M., 2014b. Effects of curing and stress conditions on hydromechanical, geotechnical and geochemical properties of cemented paste backll.
Eng. Geol. 168, 2337.
Yilmaz, E., Belem, T., Bussire, B., Mbonimpa, M., Benzaazoua, M., 2015. Curing time effect
on consolidation behaviour of cemented paste backll containing different cement
types and contents. Constr. Build. Mater. 75, 99111.

You might also like