You are on page 1of 12

This article appeared in a journal published by Elsevier.

The attached
copy is furnished to the author for internal non-commercial research
and education use, including for instruction at the authors institution
and sharing with colleagues.
Other uses, including reproduction and distribution, or selling or
licensing copies, or posting to personal, institutional or third party
websites are prohibited.
In most cases authors are permitted to post their version of the
article (e.g. in Word or Tex form) to their personal website or
institutional repository. Authors requiring further information
regarding Elsevier’s archiving and manuscript policies are
encouraged to visit:
http://www.elsevier.com/copyright
Author's personal copy

Minerals Engineering 22 (2009) 1307–1317

Contents lists available at ScienceDirect

Minerals Engineering
journal homepage: www.elsevier.com/locate/mineng

Saturated hydraulic conductivity of cemented paste backfill


M. Fall a,*, D. Adrien b, J.C. Célestin a, M. Pokharel a, M. Touré a
a
Department of Civil Engineering, University of Ottawa, Ontario, Canada K1N 6N5
b
Department of Civil Engineering, ENS, Cachan, France

a r t i c l e i n f o a b s t r a c t

Article history: The key design parameters of cemented paste backfill (CPB, a mix of tailings, water and binder) are
Received 28 May 2009 strongly influenced by its saturated hydraulic conductivity (permeability). However, our understanding
Accepted 6 August 2009 of the permeability of CPBs, as well as factors that affect it and its evolution with time, is limited. Hence,
Available online 5 September 2009
a laboratory investigation is conducted to study the hydraulic conductivity of CPBs and develop a model
for predicting its evolution with time. The results show that the hydraulic conductivity of CPB is time-
Keywords: dependent. As the curing time increases, the hydraulic conductivity decreases. The permeability is also
Cemented paste backfill
affected by the mix components. The permeability decreases as the binder content increases or the w/
Tailings
Hydraulic conductivity
c ratio decreases. Medium tailings with 45% fine particles confer lower hydraulic conductivity to the
Environmental performance CPB. The sulphate can have two opposite effects on the permeability of CPBs, contributing to an increase
Durability or decrease. However, the magnitude of the influence of the mix components depends on the curing time
and is generally more pronounced at early ages (67 days). Moreover, this study demonstrates that the
hydraulic conductivity decreases with curing temperature and time for the studied CPBs. However, the
effect of curing temperature on the hydraulic conductivity of CPBs is more significant in early age samples
(up to 7 days) and depends on the binder type. Furthermore, the mechanical damage can significantly
increase the hydraulic conductivity. Finally, the authors propose a simple function for the prediction of
the evolution of the hydraulic conductivity of CPB with time. The validation results show that the devel-
oped model is able to predict the time-dependent change of the hydraulic conductivity with good
accuracy.
Ó 2009 Elsevier Ltd. All rights reserved.

1. Introduction 85%, water that is either fresh or mine processed, and hydraulic
binder, which is usually 2–7% by weight. This technology has eco-
Financial revenues generated by the mining industry are vital nomical as well as environmental benefits for the mining industry.
for the economy and well being of several countries and popula- The use of CPBs in voids created during mining operations allows
tions in the world. However, the mining industry generates a large the mining industry to extract additional ores. CPB technology sig-
amount of solid waste, such as waste rock and tailings. The dis- nificantly reduces the quantity of mine waste that has to be man-
posal of such waste can create several problems (Ritcey, 2005). aged on the earth’s surface (Benzaazoua et al., 2004; Yilmaz et al.,
Public perception and strict government regulations regarding 2004; Orejarena and Fall, 2008). In addition to this, the hardened
the disposal of such waste compel the mining industry into think- CPBs provide ground support for the surrounding mine structure
ing and developing new strategies which are environmentally and also a safe working environment for the mine workers (Has-
sound and cost effective. sani and Archibal, 1998; Kesimal et al., 2003; Fall and Benzaazoua,
In this scenario, recycling of such waste is gaining popularity in 2005).
Canada and other parts of the world. Cemented paste tailing (CPT), Mechanical performance, stability of barricades, environmental
also called cemented paste backfill (CPB), is a relatively new tech- performance and durability belong to the most important design
nology which is practiced extensively in several mining sites criteria of CPB structures. Once placed, CPBs have to satisfy certain
worldwide (Fall et al., 2008). CPB is a heterogeneous material pro- loading requirements to ensure a safe underground working envi-
duced by mixing tailings with a solid percentage between 70% and ronment for all mining personnel. This is mainly dependent on the
mechanical properties of the CPB. Moreover, to carry fresh and
hydrating CPB materials during stope filling and thereby prevent
* Corresponding author. Address: Department of Civil Engineering, University of CPB from flowing into the mine working areas, retaining wall
Ottawa, 161 Colonel by, Ottawa (Ontario), Canada K1N 6N5. Tel.: +1 613 562
structures called barricades (permeable free draining retaining
5800x6138.
E-mail address: mfall@eng.uottawa (M. Fall). walls) or bulkheads (retaining walls that are impermeable) are

0892-6875/$ - see front matter Ó 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.mineng.2009.08.002
Author's personal copy

1308 M. Fall et al. / Minerals Engineering 22 (2009) 1307–1317

built in each of the access ways into the stope prior to stope filling. 2.1.2. Tailings
The stability of these retaining walls is significantly affected by the The tailings that were used were natural tailings (NT) and
pore water pressure developed behind the barricades or bulkheads ground silica (SI). The NT has a grain size curve close to that of SI
and the drainage ability of the wall. These two factors are strongly (Fig. 1). The NT originated from hard magmatic rocks and is char-
influenced by the saturated hydraulic conductivity (permeability) acterized by the dominance of sulphide minerals (pyrite), quartz
of the hydrating CPB. Furthermore, one of the most important and paragonite. The main objective for selecting SI as tailings
parameters affecting environmental performance and durability material is to control with accuracy, the chemical and mineralogi-
of CPBs is permeability. Susceptibility to AMD and ability to release cal composition of the tailings. This maintains the level of uncer-
contaminants into the mine areas and/or groundwater (after mine tainties at a minimum level when studying the isolated impact of
flooding) are relevant environmental design criteria for CPB struc- the mix components of CPB and temperature on its hydraulic con-
tures. The sensitivity of CPBs to AMD is mainly dependent on the ductivity. NT may contain various (uncontrollable) minerals and
reactivity (oxidation potential) of the tailings contained in the chemical elements that can influence the outcome of the study.
CPB. In turn, this reactivity is not only dependent on the types The SI contained 99.8% SiO2 and shows a grain size distribution
and quantity of sulphide minerals present in the CPB system, but close to the average of nine Canadian mine tailings (Fig. 1). Table
also the ease with which fluids, such as oxygen and water, enter 1 gives the main physical and chemical characteristics of the NT
and move through the CPB matrix, i.e. on the permeation proper- and SI. It can be observed that NT and SI, with about 41–45 wt.%
ties of the CPB. These permeation properties can be assessed using fine particles (<20 lm), can be classified as medium tailings (med-
knowledge of the hydraulic conductivity of the CPB. Furthermore, ium tailings have between 35% and 60% of particles with size
hydraulic conductivity is one of the main parameters controlling 620 lm). Furthermore, to evaluate the impact of tailings fineness
the groundwater flow rate through the CPB structure once it is on the hydraulic conductivity of CPBs, two other SI tailings materi-
flooded. In other words, it significantly affects the leaching poten- als (SI-35, SI-20) with a percentage of fine particles (passing
tial and transport of contaminants through the CPB to groundwater 20 lm) of 35%, 20% by weight, respectively, were used. These tail-
(Levens et al., 1996). In addition, hydraulic conductivity can give ings had the same chemical and physical characteristics (except
relevant information about the pore structure, such as coarseness grain size distribution) as SI. This allowed us to isolate the grain
and connectivity, and the cracking of the CPB. Poorer pore struc- size factor and evaluate its impact on CPB hydraulic conductivity.
tures, including coarse pores and high connectivity of the pores,
and cracks can allow and accelerate fluid transfer, such as oxygen 2.1.3. Water
and water, between the CPB and surrounding media, thereby Tap water was mainly used to mix the binders and tailings.
resulting in increased potential oxidation of the sulphide minerals However, specific amounts of sulphate concentrates (FeSO4.7H20)
contained in the tailings and reducing service-life through sulphate were added to a specific volume of distilled water to create mixing
attacks. waters with well known sulphate concentrations (0 ppm,
During the past 15 years, research efforts have been primarily 5000 ppm, 15,000 ppm, and 25,000 ppm), i.e. with different chem-
spent on understanding the factors that affect the mechanical ical compositions.
properties of CPBs (e.g., Landriault, 1995; Kesimal et al., 2003; Yil-
maz et al., 2004; Fall et al., 2005, and Fall et al., 2007). However, 2.2. Specimen preparation and mix proportions
technical data on the hydraulic conductivity of CPBs is quite lim-
ited. Furthermore, most of the studies were carried out on CPBs Around 300 CPB samples made of PCI (PCI-CPB) or PCI/Slag
cured at ‘‘laboratory” curing temperatures (20–25 °C). This means (Slag-CPB) were prepared to investigate the saturated hydraulic
that the impact of colder (0–20 °C) or warmer (25–50 °C) curing conductivity of CPBs as well as their microstructure. The tailings
temperatures on the hydraulic properties of CPBs is still not under- materials, binders and water were mixed until a homogeneous
stood. In addition, the CPB is an evolutive porous medium, i.e. its paste was obtained. The slump (consistency) of the CPBs was kept
properties (e.g., hydraulic conductivity) change with time. There constant at 18 cm (the most frequent slump value used for CPB de-
is currently no available model for predicting the evolution of the signs), except for mixes prepared with a water/cement ratio (w/c
hydraulic conductivity of CPBs with time. Thus, there is a need to ratio) = 5, 6 and 10. The consistency of the paste mixtures was
substantially increase our knowledge about the hydraulic conduc- measured by a slump test in accordance to ASTM C 143-90. The
tivity of CPBs and the factors that can affect it, and to develop a produced CPB mixes were poured into curing cylinders that were
model that will allow the prediction of time-dependent evolution 5 cm in diameter and 10 cm in height. The specimens were then
of the hydraulic conductivity of CPBs. This has inspired the authors sealed (to avoid evaporation of water) and cured in environmental
to conduct the present research, experimentally study the satu- chambers at specific curing temperatures of 2 °C, 20 °C, 35 °C and
rated hydraulic conductivity of CPBs and the factors affecting it 50 °C for periods of 1, 7, 28, 56 and 90 days. Afterwards, the CPB
as well as develop a model for predicting the evolution of the samples were subjected to various tests, including hydraulic con-
hydraulic conductivity of CPBs with time. The main results of this ductivity tests on undamaged and mechanically pre-damaged
research are presented in the coming sections. CPB samples, and microstructural analysis.

2.3. Testing of specimens


2. Materials and experimental program
2.3.1. Hydraulic conductivity tests on undamaged CPB samples
2.1. Materials used
To study the effects of curing time, mix components of CPBs and
curing temperature on the hydraulic conductivity of CPBs, satu-
The materials used include binders, tailings, and water.
rated hydraulic conductivity tests were conducted on undamaged
CPB samples cured at various temperatures (2 °C, 20 °C, 35 °C
2.1.1. Binders and 50 °C) and times (mostly at 1, 3, 7, 28, 56 to 90 days). In early
Portland cement Type I (PCI), and blast furnace slag (Slag) were ages (1 to 7 days), a rigid wall permeameter in falling head mode
used as binders. PC I was used alone or blended with Slag. The following ASTM D 5856-00 was used. This method proved to be
blending ratio of PCI and Slag was 50/50. These binders are often impractical for advanced age (28–90 days) sample testing. At these
used for CPB mixtures in mines located in Eastern Canada. latter periods, the hardened samples required higher hydraulic
Author's personal copy

M. Fall et al. / Minerals Engineering 22 (2009) 1307–1317 1309

100
SI
90
NT
80 9mines

Volumepercent (cumulative)
70

60

50

40

30

20

10

0
0.01 0.1 1 10 100
Grain size (µm)

Fig. 1. Grain size distribution of the tailings (SI, NT) used and the average grain size distribution of tailings from nine Canadian mines.

Table 1 compressive load on the CPB samples. The stress or load level (ap-
Physical properties of the tailings used. plied stress/peak stress) was preset and applied under cyclic load-
Element Gs D10 (lm) D30 (lm) D50 (lm) D60 (lm) Cu Cc ings. The permeability of the samples was measured after each
SI 2.7 1.9 9.0 22.5 31.5 16.6 1.3 cycle of loading and unloading, i.e. at different load levels (load lev-
NT 3.6 3.6 13.2 28.3 39.5 11.0 1.2 els up to 90% of the evaluated UCS were considered) in the pre-
peak phase or in other words, at different mechanical damages
(D). D is defined as follows:
gradient to obtain infiltration by the permeant. Another difficulty E
arose that required the destruction of the cell after testing as the D¼1 ð1Þ
E0
hardened paste was almost impossible to remove from the perme-
ameter cell. About 120 tests had to be conducted in such condi- with: D: damage, 0 6 D 6 1, E: Young’s modulus of damaged CPB,
tions; therefore, the choice of using the flexible wall technique E0: initial Young’s modulus of CPB (determined in the elastic range
was addressed. The procedure for this method is described in of the stress–strain curve of CPB).
ASTM D5084-00 and was conducted in the constant head mode. Each test was performed at least twice to ensure the repeatabil-
Each hydraulic conductivity test was repeated at least twice and ity of the results obtained.
the average value was considered as the saturated hydraulic con-
ductivity of the sample tested. 2.3.3. Microstructural analysis
The microstructure (pore structure, binder hydration products)
of some CPB samples was also evaluated. The microstructure of the
2.3.2. Hydraulic conductivity tests on mechanically damaged CPB
studied CPB samples was investigated by mercury intrusion poros-
samples
imetry (MIP), X-ray tests and scanning electron microscopy (SEM)
To experimentally clarify the impact of mechanical damage or
observations. Before MIP-testing, all samples were first dried at
the opening of cracks induced by mechanical loads on the hydrau-
50 °C up to mass stabilization. Drying at this temperature did not
lic conductivity of CPBs, hydraulic conductivity tests were con-
appear to cause cracking. The MIP tests allowed the evaluation of
ducted on mechanically pre-damaged CPB samples by using a
the pore size distribution in the dried CPB samples. X-ray tests
triaxial cell (flexible wall technique) in accordance to ASTM
were carried out on some CPB samples as well as some prepared
D5084-00. Prior to conducting permeability tests on damaged
hardened cement pastes with a high w/c ratio (w/c = 2; to simulate
CPB samples, the uniaxial compressive strength (UCS) and the ini-
the cement matrix of CPB) by using a Scintag XDS 2000 XRD. The X-
tial permeability (k0) of the studied CPB samples were determined.
ray tests and SEM observations allowed the studying of binder
The UCS of the CPB samples was determined by using a computer-
hydration products that formed in the CPB system. The SEM obser-
controlled mechanical press (MTS 10/GL) with a normal loading
vations were done with Hitachi S4800 FEG–SEM. The samples were
capacity of 50 kN. The UCS test was performed in accordance to
observed in backscatter electron mode.
standard ASTM C 39. The compression tests were carried out at a
constant deformation rate of 1 mm/min. The axial deformation
was automatically measured by using an internal linear variable 3. Results and discussion
differential transformer (LVDT) connected to an electronic data
acquisition system. A measurement of the k0 of the CPB (undam- 3.1. Effect of curing time on the saturated hydraulic conductivity of
aged) was performed at 20 °C by using the aforementioned flexible CPB
wall technique. Afterwards, hydraulic conductivity tests in accor-
dance to ASTM D5084-00 were carried out over pre-damaged sam- The CPB samples were prepared with plain PCI and NT or SI.
ples. Mechanical damage was achieved by applying uniaxial The total binder content was fixed at 4.5% and the slump at
Author's personal copy

1310 M. Fall et al. / Minerals Engineering 22 (2009) 1307–1317

1.0E-04 that region # 1 corresponds to the main period of cement hydra-


CPB made of SI tion, during which the larger amount of hydration products are
8.0E-05 produced (Taylor, 1993). From the results presented above, it
Permeability, k (cm/s)

CPB made of NT
can be concluded that since the permeability of CPBs decreases
6.0E-05 with curing time, their fluid transport ability (water, oxygen) will
decrease with time. In other words, their durability and environ-
4.0E-05 mental performance may increase.

2.0E-05 3.2. Effect of CPB’s mix components on its saturated hydraulic


conductivity
0.0E+00
0 10 20 30 40 50 60
The main components of CPBs include the binder, mixing water
Curing time (days)
and tailings. The nature (e.g., chemical and physical characteristics)
Fig. 2. Time-dependent change of the permeability of CPBs containing SI or NT as and amount of these components used to prepare CPB materials in
tailings materials (4.5% PCI; slump = 18 cm). the mine plants vary greatly. Therefore, the effects of the mix com-
ponents of CPBs and their variations on the permeability of CPBs
that are cured at 20 °C were investigated. The main results ob-
18 cm for all mixes. The samples obtained were cured at 20 °C for tained are summarized in Fig. 4a–e.
periods of 1, 7, 28 and 56 days. Fig. 2 illustrates the influence of
curing time on the evolution of the permeability of the studied
CPBs. From this figure, it is obvious that regardless of the tailings
used, the curing time significantly affects the hydraulic conductivity
a 1.4E-04 3.0%PCI
of CPBs. As the curing time increases, the hydraulic conductivity 1.2E-04 4.5%PCI
Permeability,k (cm/s)
decreases. The reason for this behaviour is the increasing degree 6.0%PCI
1.0E-04
of cement hydration with time, i.e. in the amount of cement
hydration products (e.g., CSH, CH, ettringite) formed. The binder 8.0E-05
hydration leads to a progressive refinement of the pore structure 6.0E-05
and reduction of the porosity of the CPB materials by means of
4.0E-05
filling the pores in the cemented matrix of the CPBs and blocking
the interconnected pores by the aforementioned hydration prod- 2.0E-05
ucts. This refinement of the microstructure of CPBs with increas-
0.0E+00
ing curing time is graphically demonstrated in Fig. 3. This figure 0 10 20 30 40 50 60
represents the results of MIP tests performed on two samples Curing time (days)
made of tailings NT and cured for periods of 7 and 28 days, respec-
tively. It can be seen that the pore size distribution of the CPB b 1.0E-04
cured at 7 days is coarser than that of the specimen cured at 4.5%PCI
28 days. In the graph shown in Fig. 2, it is also possible to clearly 8.0E-05
Permeability,k (cm/s)

4.5%PCI/ Slag
distinguish two different regions with a time of 0–7 days (region #
1) and over 7 days (region # 2) in the permeability of CPBs with 6.0E-05
changes that are time-dependent. The highest decrease of the
permeability of CPBs with curing time occurs in region # 1, while 4.0E-05
region # 2 is characterized by a slight decrease of the k-value of
CPBs with curing time. This behaviour can be explained by the fact 2.0E-05

0.0E+00
0 10 20 30 40 50 60
12
Curing time (days)
28 days
c
Incremental Hg intrusion porosity (%)

10
7 days 1.6E-04
w/ c = 5
1.4E-04
w/ c = 7.6
Permeability,k (cm/s)

8 1.2E-04
w/ c = 10
1.0E-04

6 8.0E-05

6.0E-05

4 4.0E-05

2.0E-05

2 0.0E+00
0 10 20 30 40 50 60
Curing time (days)
0
0.001 0.01 0.1 1 10 100 Fig. 4. Effect of CPB’s mix components on its permeability: Effect of (a) binder
Pore size (µm) content (PCI; slump = 18 cm; SI used); (b) binder type (4.5% cement;
slump = 18 cm; SI used); (c) w/c ratio (4.5% PCI; SI used); (d) tailings fineness
Fig. 3. MIP pore size distribution of paste backfill cemented with 4.5% PCI vs. curing (4.5% PCI; slump = 18 cm; SI used); and (e) sulphate content (4.5% PCI;
temperature and pore size range (w/c = 7.6; slump = 18 cm; NT used). slump = 18 cm; SI used).
Author's personal copy

M. Fall et al. / Minerals Engineering 22 (2009) 1307–1317 1311

to the physical (filler) and chemical (pozzolanic) effect of the min-


d 1.6E-04 4.2E-03(cm/s)
20% Fine eral admixture, as shown in other cemented materials (Manmohan
1.4E-04
32% Fine and Mehta, 1981). It is known (Manmohan and Mehta, 1981; Hoo-
1.2E-04
45% Fine ton, 2000) that Slag improves pore structure in two manners: its
Permeability, k (cm/s)

1.0E-04 small particle size results in a filler effect in which the Slag parti-
8.0E-05 cles bridge the spaces between cement particles and the spaces be-
tween cement particles and tailings grains; and the Slag reacts
6.0E-05
pozzolanically with calcium hydroxide (CH) resulting from the
4.0E-05 hydration reactions of PC to produce a greater solid volume of C–
2.0E-05 S–H gel. This in turn, leads to an additional reduction in capillary
0.0E+00
porosity during hydration (by obstruction of pores and voids)
0 10 20 30 40 50 60 (Manmohan and Mehta, 1981), and a denser packing within the ce-
Curing time (days) mented matrix, thereby reducing the fluid transportability of the
CPB. These results then imply that the partial replacement of PCI
e 8.0E-05 by Slag should result in an increase of the environmental perfor-
0 ppm
7.0E-05 mance of CPBs.
5000 ppm
Permeability, k (cm/s)

6.0E-05 15000 ppm


3.2.2. Effect of w/c ratio
25000 ppm
5.0E-05 Fig. 4c illustrates the influence of the w/c ratio on the hydraulic
4.0E-05 conductivity of CPBs with respect to curing time (1, 7, 28, and
56 days). CPB samples with a w/c = 5, 7.6 and 10, respectively, were
3.0E-05
prepared. The binder (PCI) content was kept constant at 4.5%. The
2.0E-05 workability of the sample with w/c = 5 was poor. This w/c ratio was
1.0E-05
only selected from a research point of view. From Fig. 4c, it can be
seen that the permeability of CPB decreases as the w/c ratio de-
0.0E+00 creases. This is because a higher w/c ratio confers a higher capillary
1 3 7 28 90
Curing time (days)
porosity to the CPBs (Fall et al., 2005). However, the amplitude of
the impact of the w/c ratio on CPB’s permeability is more pro-
Fig. 4 (continued) nounced at early ages (67 days). This can be explained by the fact
that a lower w/c ratio leads to lower porosity and faster binder
hydration (at early ages) and thereby, refines the pore structure
3.2.1. Effect of binder content and type
and blocks the connectivity of the pores within the cement of CPBs
Fig. 4a shows the effect of the binder proportion (3%, 4.5% and
with lower w/c ratio quicker.
6% PCI) on the permeability of the CPB prepared at a constant
slump (18 cm) in relationship to the curing time. It can be observed
3.2.3. Effect of tailings fineness and type
that the binder content significantly affects the permeability of the
Typical results of the effect of tailings fineness on CPB perme-
CPB, especially at early ages. The permeability of the CPB decreases
ability in relationship with the curing time are presented in
as the binder content increases. This is because higher binder con-
Fig. 4d. From this figure, it appears that the proportion of fine tail-
tent leads to more precipitation of hydration products (e.g., CSH,
ings particles (<20 lm), i.e., the fineness of the tailings materials
CH, ettringite) in the cemented matrix of the CPB, thereby causing
used in the CPB mixture, has a large effect on the hydraulic conduc-
a refinement of the pore structure and reduction in the porosity of
tivity of CPBs. It can be noted that medium tailings with 45% fine
the CPB (Taylor, 1993; Fall et al., 2008). However, from Fig. 4b, it
particles confer lower hydraulic conductivity to the CPB. The CPB
can be noted that aside from the binder content, the type of binder
specimens made of coarse tailings (20% and 32% fineness) gener-
used also has an impact on the hydraulic conductivity of CPBs.
ates higher permeability for a given consistency. The observed
Fig. 4b shows the variation of the hydraulic conductivity of CPBs
with binder type (100% PCI, PCI blended with Slag at a ratio of
50/50) in relationship to the curing time. It is obvious that the type 7
of binder or the mineral admixture has an influence on the CPB’s 30% Fine
32%Fine
permeability. However, the magnitude of this influence depends 6 45%Fine
45% Fine
on the curing time. For curing times that are 61 day, the perme-
ability of CPBs made of PCI (PC-CPB) is higher than that of CPBs
Hgintrusion porosity (%)

5
containing Slag (Slag-CPB) as mineral admixtures. The measured
permeability value of Slag-CPB was about 10% higher than that of
4
PCI-CPB. This can be mainly attributed to the filler effect of the
mineral admixture Slag, which in addition to the binder hydration
3
products, contribute to filling the pores of the CPB matrix, thereby
refining its pore structure (Manmohan and Mehta, 1981; Fall et al.,
2008). However, an observation of Fig. 4b indicates that for a cur- 2

ing time higher than 1 day and lower than 28 days, the permeabil-
ity of PCI-CPB is lower than that of Slag-CPB. This is because the 1
hydration process is slower and delayed in Slag compared with
PCI (Demirboga, 2007). For this reason, the binder hydration in- 0
duced porosity reduction in CPB made with only PCI is faster than 0.001 0.01 0.1 1 10 100 1000
that in CPB containing Slag (Celestin and Fall, 2009). For curing Pore diameter (µm)

times that are P28 days, the Slag-CPB specimens show slightly
Fig. 5. Effect of tailings fineness on pore size distribution of CPB specimens
lower permeability than those made of PCI. This can be attributed cemented with PCI/Slag for a given w/c ratio (w/c = 6).
Author's personal copy

1312 M. Fall et al. / Minerals Engineering 22 (2009) 1307–1317

increase of the permeability of CPB for a proportion of tailings fine- more porous cemented matrix. From Fig. 4e, it can be observed that
ness lower than 45% can be attributed to a reduction of the packing as the curing time increases from 1 to 28 days, the hydraulic con-
density of the tailings materials and coarsening of the pore struc- ductivity values of the sulphate-bearing CPB specimens are lower
ture as the fineness content decreases from 45% to 20%. The first than those of sulphate-free CPBs specimens. This lower permeabil-
assumption is supported by the results of the experimental inves- ity associated with sulphate concentration can be explained by the
tigations conducted by Fall et al. (2005) on the effect of tailings fact that, as mentioned in Fall and Benzaazoua (2005), sulphate can
fineness on the microstructure of CPBs. This study has shown that contribute to the decrease of CPB porosity. The reason can be
the packing density of the tailings decreases as the proportion of attributed to the precipitation of secondary hydrated minerals
fine tailings becomes lower than 45%. Lower packing density is (such as gypsum, ettringite and brucite) in the empty capillary
associated with higher volume of void spaces between the tailings pores that are present in the CPB. This is graphically demonstrated
particles available for cement hydration products. Indeed, coarse by the results of SEM observations presented in Fall and Benzaa-
tailings particles produce bodies that have fewer particle–particle zoua (2005). From this study, it can be observed that secondary hy-
contacts per unit volume, and thus, larger pores between the par- drated minerals precipitate in the pores that were inside the
ticles. The latter assumption is in perfect agreement with the re- cement matrix and thus, reduce its porosity. However, from
sults of the MIP tests (Fig. 5) performed on two Slag-CPB samples Fig. 4e, it can be also seen that for advanced curing times (90 days),
prepared with the same binder content (4.5%) and w/c ratio (w/ the CPB specimens with high initial sulphate concentrations
c = 6), but with tailings that had two different finenesses (32% (15,000 and 25,000 ppm) behave differently than those with low
and 45%). It can be observed that the CPB with 32% fine tailings initial sulphate content (5000 ppm) or without sulphate. The
has a coarser pore structure. Its volume of macropores >1 lm is hydraulic conductivity of the 15,000 and 25,000 ppm CPBs is
the highest. Furthermore, the threshold diameter (the pore size slightly higher than that of 0 and 5000 ppm CPB specimens as well
corresponding to the highest rate of mercury intrusion per change as higher than the permeability values obtained at 28 days curing.
in pressure, (Manmohan and Mehta, 1981)) value of the CPB with a This reveals a coarsening of the pore structure of 15,000 and
45% fineness (1.4 lm) is 4 times lower than that of the CPB with a 25,000 ppm CPBs as the curing time increased from 28 to 90 days
32% fineness (4.8 lm). In other words, this means that the hard- and also in comparison to the 0 and 5000 ppm samples. These
ened CPB with 45% fineness has lower fluid transportability, since higher values of permeability can be explained by the formation
the threshold diameter is characteristic of a capillary network con- of excessive amounts of expansive minerals (ettringite, gypsum)
nected to the surface of the cemented porous media (Manmohan resulting from the reaction between the sulphate ions and C3A of
and Mehta, 1981). Contrary to the tailings fineness, Fig. 2, which the cement (formation of ettringite) as well as between the sul-
illustrates the time-dependent evolution of CPB specimens made phate ions and cement hydration product; CH (formation of gyp-
of tailings SI and NT, reveals that the mineralogical composition sum). These expansive minerals could have exerted excessive
of tailings does not significantly affect the permeability of CPBs. pressure on the pore structure of the 15,000 and 25,000 ppm CPBs,
resulting in formation of microcracks in the cemented matrix. This
3.2.4. Effect of initial sulphate content will cause an increase in the permeability of the CPBs. This
Fig. 4e shows the evolution of the permeability of CPBs contain- assumption is fully supported by the results of SEM observations
ing 4.5% PCI with time for various initial sulphate concentrations conducted on a high sulphate-bearing CPB sample made of tailings
(0, 5000, 15,000, and 25,000 ppm). From this figure, it can be seen NT and 4.5% PCI (Fig. 6). From Fig. 6, it can be clearly noted that the
that the initial sulphate content in the CPB material has an impact pores within the CPB which are filled with secondary gypsum min-
on its hydraulic conductivity. Depending on the curing time and erals have damaged the CPB by their destructive expansive action.
sulphate concentration, the sulphate can have two opposite effects
on the permeability of CPBs, contributing to an increase or de-
3.3. Effect of curing temperature on the saturated hydraulic
crease. At 1 day curing, the permeability values of CPB specimens
conductivity of CPB
with 15,000 and 25,000 ppm are slightly higher than those with
0 and 5000 ppm. This can be attributed to the fact that high sul-
The temperature in backfill operations is a function of several
phate concentrations inhibit cement hydration at early ages (Fall
parameters that can be classified as internal (e.g., heat produced
and Benzaazoua, 2005), which in turn, results in the formation of
by binder hydration) and external (e.g., temperature of rock mass
adjacent to or surrounding the CPB structure, depth mine temper-
ature, geographical position, heat generated by mining operations,
heat produced during CPB transport through the pipelines). These

8.E-05

7.E-05 1dayPCI 7days PCI


1dayPC/Slag 7days PC/Slag
rmeability,k (cm/s)Pe

6.E-05
90days PCI 90days PC/Sla
5.E-05
28days PCI 28days PC/Sl
4.E-05

3.E-05

2.E-05

1.E-05

0.E+00
0 10 20 30 40 50 60
Temperature (οC)
Fig. 6. SEM images showing secondary hydrated minerals that have damaged the
CPB by their destructive expansive action. SEM image showing CPB damaged by the Fig. 7. Coupled effect of temperature and binder on the hydraulic conductivity of
destructive action of the expansive mineral phase formed within the CPB pores. CPB at various curing temperatures (4.5% binder; slump = 18 cm; SI used).
Author's personal copy

M. Fall et al. / Minerals Engineering 22 (2009) 1307–1317 1313

various thermal sources and loads on the CPB may affect its perme- interconnected to provide an easy path for liquid flow. As the tem-
ability, or in other words, its drainage ability, and environmental perature increases up to 50 °C, a significant drop is observed in the
and durability performances. Consequently, the impact of curing hydraulic conductivity and this decrease continues with longer
temperatures on the hydraulic conductivity of CPB was studied. curing time. This means that the hardening process has evolved
The main results obtained are summarized in Fig. 7 and discussed following minimum binder stoichiometry (Mindess and Young,
below. 1981). The rate of formation of hydration products has a direct ef-
fect on the short and long-term refinement structure of the solid
3.3.1. Combined effects of curing temperature and time matrix (Fall and Benzaazoua, 2005) and is therefore, responsible
From Fig. 7, the hydraulic conductivity of PCI-CPB samples for the drop in hydraulic conductivity encountered at these tem-
cured at 2 °C, 20 °C, 35 °C and 50 °C for various curing times can peratures. This is because of the increase of curing temperature
be observed. From this figure, it is noticed that the hydraulic con- which significantly accelerates the hydration rate of the cement
ductivity decreases with curing time and temperature for the stud- paste (Lothenbach et al., 2007). Fall et al. (2007) also observed
ied CPBs. At one day of curing time, the hydraulic conductivity is the increase in the formation of CH, C–S–H and calcite with in-
always higher than all of the other advanced curing times. The crease of curing temperature by performing a thermal analysis
hydraulic conductivity appears to be affected in almost the same on hardened cemented paste of CPBs. This is further confirmed
extent by temperature and curing time. For example, on the one by the results of XRD tests performed on cemented paste of CPBs
hand, the difference in hydraulic conductivity drops from cured at various temperatures by Pokharel (2008). These results
6.97  105 to 1.40  105 cm/s (5.57  105) for 1 day samples, show that high curing temperatures allow CPB samples to form
6.02  105 to 1.54  106 cm/s (5.86  105) for 7 day samples high amounts of hydration products. These products fill the voids
and from 4.72  106 to 1.54  106 cm/s (3.18  106) for and reduce the porosity so that the sample has low hydraulic con-
90 day samples in terms of differences when the temperature ductivity at higher curing temperatures. This temperature induced
changes from 2 °C to 50 °C. These variations expressed in terms refinement of the pore structure of CPBs is illustrated by Fig. 8. This
of percentage represent a decrease in hydraulic conductivity of figure shows the main results of the effect of temperature on the
79%, 74% and 67%, respectively. These results show that the MIP curve (porosity and pore size distribution) of PCI-CPB after
changes in hydraulic conductivity with temperature are more sig- 7 days of curing. From this figure, it appears that CPBs cured at
nificant in early age samples (up to 7 days). On the other hand, the higher temperatures have a finer pore structure. The volume of
changes recorded with varying curing times from 1 to 90 days at macropores >1 lm is highest in CPB samples cured at 20 °C. From
2 °C, 20 °C, 35 °C and 50 °C are 6.50  105 cm/s, 2.08  105 cm/ this figure, it can be also observed that a higher curing temperature
s, 2.21  105 cm/s, and 1.25  105 cm/s in terms of differences means smaller threshold diameters; i.e. the pore structure of the
or 93%, 90%, 97%, and 89% in terms of decrease percentages, respec- CPB is finer. It is generally accepted that the threshold diameter
tively. Changes are slightly more significant at low curing temper- is a unique transport length scale of major importance in the per-
atures than at higher temperatures. An examination of samples meability properties of cemented material (Garboczi, 1990). There-
cured at 2 °C show that they all bear the highest hydraulic conduc- fore, the 7 days PCI-CPBs cured at higher temperatures have lower
tivity. The higher values can be explained with the fact that hydra- hydraulic conductivity than CPBs cured at colder temperatures
tion products are formed slowly under low reactivity conditions (Fig. 7). However, the effect of curing temperature on the hydraulic
caused by low temperatures leading to a slow chemical reaction conductivity of CPBs depends on the binder type as discussed
in the binder (slow hydration, Lothenbach et al., 2007). This confers below.
high porosity and pore connectivity to the cementitious materials
in comparison to the sample cured at high temperatures (Fall 3.3.2. Combined effects of curing temperature and binder type
et al., 2007). A considerable amount of time (90 days) was required In Fig. 7, the hydraulic conductivity recorded on four groups of
for the formation of sufficient hydration products which produces samples (1, 7, 28 and 90 days) made of 4.5% PCI and 4.5% PCI/Slag
a relatively dense cementitious system that fills the voids and re- (50/50) cured during the same period of time and temperature
duces the pore connectivity. Thus, at 1 day, the presence of non ab- conditions (2 °C, 20 °C, 35 °C and 50 °C) are also presented to ob-
sorbed water is very high (Jensen and Hansen, 2002), liquid phase serve the coupling effect of temperature and binder type. The dis-
is continuous, and thereby, voids are possibly sufficiently well tinct behaviour of the two different binders is obvious. One can
notice that at 1 day of curing time and temperature of 2 °C, the
hydraulic conductivity of samples containing Slag is higher than
60 the ones made of neat PCI. However, at higher temperatures of
20°C 20 °C, 35 °C and 50 °C, the hydraulic conductivity of samples con-
Incremental Hg intrusion porosity (%)

35°C taining Slag is lower than the one made of neat PCI. This cross
50
50°C behaviour at different temperatures leads to invoking the cemen-
ted and pozzolanic properties of the two binders. First, as a hydrau-
40 lic binder, PCI possesses low activation energy and can initiate the
formation of hydration products rapidly due to the glassy phase
30 and C3A in its minerals (Mindess and Young, 1981). By its side,
as a pozzolan, Slag contains crystalline particles that generally re-
quire either the presence of an alkaline environment or CaO and
20
water to develop binding qualities (Palmer et al., 2000; Wang
et al., 2004). Furthermore, these pozzolans bear higher activation
10 energy for starting the chemical reaction. Therefore, it can be in-
ferred from this graph that at 2 °C, while hydration has started in
0 PCI, it is necessary for the mixes containing Slag to wait until the
0.001 0.01 0.1 1 10 100 1000 pore water becomes sufficiently alkaline prior to starting the poz-
Pore size diameter
zolanic reaction in order for the Slag to contribute to the formation
Fig. 8. Effect of curing temperature on the pore size distribution of PCI-CPB (4.5% of hydration products. This delay in the apparition of hydration
PCI; w/c = 7.6; NT used; curing time = 7 days). products leaves the voids in the samples containing Slag partially
Author's personal copy

1314 M. Fall et al. / Minerals Engineering 22 (2009) 1307–1317

300

250

200
Intensity,CPS

monosulphate
150 CaSO4 Ca(OH)2 Ca(OH)2

CaCO3

ettringite ettringite CSH Ca(OH)2 Ca(OH)2


100

50

0
5 10 15 20 25 30 35 40 45 50 55 60

Degrees 2-theta-Cu Kα Radiation

300

250

monosulphate
200 CaCO3
CaSO4 Ca(OH)2
Intensity,CPS

Ca(OH)2

150

CaCO3 Ca(OH)2
ettringite CSH Ca(OH)2
100
CaCO3

50

0
5 10 15 20 25 30 35 40 45 50 55 60
Degrees 2-theta - Cu Kα Radiation

Fig. 9. XRD result of Slag-CPB cured at (a) 20 °C and (b) 35 °C.

free of precipitates and therefore, the CPB is less dense. Such a produced by PCI, leading to a decrease in pore volume and hence,
material is more porous at the beginning, allowing an easier flow an increase in the length of the interconnected pore path (Doven
of permeant fluid than in the case of samples made of neat PCI. and Pekrioglu, 2005). This increase in hydration products within
However, at temperatures ranging from 20 °C to 50 °C and for all the Slag-CPB samples with higher curing temperatures is demon-
studied curing times that are P7 days, the hydraulic conductivity strated by the results of XRD tests performed on cement paste of
is lower in samples containing Slag. This suggests that the pores Slag-CPB cured at 20 °C and 35 °C, and shown in Fig. 9a and b,
in Slag samples are extensively filled by precipitation products. respectively. The XRD result in Fig. 9a shows the presence of a
This filling has even surpassed the void size of samples made of higher quantity of hydration products in the sample cured at
neat PCI. This is due to the fact that aside from the binding effect, 35 °C. The intensity of CH is found higher for the sample cured at
additional C–S–H has been formed from the poor crystalline CH 35 °C (Fig. 9b) than at 20 °C (Fig. 9a). It is observed that the
Author's personal copy

M. Fall et al. / Minerals Engineering 22 (2009) 1307–1317 1315

intensity of CH at 18° 2-theta is 75 CPS in the case of the sample at impair environmental performance and durability of CPB struc-
20 °C whereas at 35 °C, this intensity is approximately 95 CPS. In tures by creating preferential paths for the penetration of various
the same way, the intensity of CH at 34°, 47°, 51° 2-theta for the types of potentially aggressive agents (e.g., oxygen, ions) and/or
sample cured at 35 °C is higher than that of the 20 °C samples. facilitating the leaching of backfill materials and transport of con-
The XRD results presented above are in good agreement with those taminants through the CPB to groundwater. Hence, the relation-
of differential thermal analysis (TG/DTA) performed by Fall et al. ship between load-induced cracking or mechanical damage and
(2007) on hardened cement pastes PCI/Slag (50/50) with a high CPB permeability is studied. For the sake of simplicity, the study
w/c ratio (w/c = 2). The comparison of the TG/DTA diagrams of is restricted to compressive loads.
the PCI/Slag cement pastes cured at different temperatures (20 °C To evaluate the impact of mechanical (compressive) damage or
and 35 °C) clearly shows that the amounts of CSH, ettringite, CH, cracking on the permeability of CPBs, hardened PCI-CPB samples
and calcite that are formed increase with the curing temperature with various cement contents (4.5% PCI, 6.0% PCI) and the same
(from 20 °C to 35 °C). In addition, the filler effect of the Slag grains slump (18 cm) were prepared. They were then cured at various
as explained above; a behaviour that PCI grains are not able to temperatures (20 °C and 25 °C for 4.5% PCI-CPB samples, 50 °C for
achieve, is a mechanism that also contributed to the observed low- 6.0% PCI-CPB samples). This allowed us to obtain CPB specimen
er hydraulic conductivity of Slag-CPB samples. with different UCS values (355 kPa, 550 kPa, 2700 kPa) and to
From the results presented above, it can be concluded that high- study the relationship between strength and the hydraulic re-
er curing temperatures contribute to reducing the fluid transport sponse of CPB subjected to mechanical damage. The prepared
ability of CPBs, i.e. improving the environmental performance of CPB specimens were subjected to several mechanical loading–
the studied CPBs. unloading cycles following the procedures explained in Section
2.3.2. Hydraulic conductivity determinations were made after each
cycle. Typical results of this study are presented in Fig. 10a and b.
3.4. Effect of mechanical damage on the saturated hydraulic
Fig. 10a and b illustrate typical examples of the impact of stress
conductivity of CPBs
level (applied stress/peak stress) and mechanical damage (D) on
the relative hydraulic conductivity of the CPBs studied, respec-
During their service life, CPB structures can be subjected to var-
tively. The relative hydraulic conductivity is expressed as the ratio
ious mechanical loads (e.g., rock wall closure, service loads, rock-
between the permeability obtained in the damaged CPB after N cy-
burst, self-weight) that may cause cracking of the CPB materials.
cles of loading/unloading and the permeability initially in the no
Since CPBs are designed to satisfy expected loading requirements,
damaged CPB (k0). These figures reveal that the applied stress level
the aforementioned loads are generally not significant enough to
and mechanical damage influence the hydraulic conductivity of
cause an important structural degradation of CPB structures (those
CPBs. Under low to medium stress levels that correspond to 50–
that are well designed from a mechanical point of view), i.e. lead-
60% of the UCS, the hydraulic conductivity tends to decrease
ing to their mechanical failure or collapse. However, with time,
slightly and/or then remains relatively constant. This can be attrib-
degradation accumulates and may lead to micro-cracking of the
uted to the contracting volumetric behaviour of the CPB material
CPB materials, resulting in permeability variations. Cracks could
under uniaxial compressive loading. Beyond the aforementioned
threshold stress level, the permeability of CPB starts to increase,
10.0 first very slightly, then significantly from about 80% of the UCS or
9.0
6.0%PCI-2700 kPa a mechanical damage value of 0.06 for the high strength CPB spec-
4.5%PCI-550 kPa imens (6% PCI-2700 kPa) and 0.15 for the low strength CPB speci-
Relative permeability (k/k0)

8.0
4.5%PCI-355 kPa mens (4.5% PCI-550 kPa, 4.5% PCI-355 kPa (Fig. 10b). This
7.0
increase in permeability results from the formation and propaga-
6.0
tion of connected networks of microcracks. Similar observations
5.0
were made in other cementitious materials (Picandet et al.,
4.0
2001). However, it can be noted from Fig. 10 that the increase in
3.0 permeability observed (beyond the threshold stress) for the high-
2.0 strength CPBs is much higher than that for the low-strength CPBs.
1.0 Furthermore, this increase is observed at a damage value of 0.06 for
0.0 the high-strength CPBs, while the low-strength CPBs show increase
0 0.2 0.4 0.6 0.8 1 1.2
in the permeability as D reaches of value of 0.15. This can be ex-
Applied stress/Peak stress
plained by the fact that high-strength CPBs show more quasi-brit-
10.0
tle behaviour than medium and low-strength CPBs (Fall et al.,
6.0%PCI-2700 kPa 2008). This brittle behaviour leads to a higher generation of micro-
9.0
4.5%PCI-550 kPa cracks in the pre-peak region of the stress–strain curves of high-
Relative permeability (k/k0)

8.0 4.5%PCI-355 kPa


strength CPBs as demonstrated in previous studies on the mechan-
7.0
ical behaviour of CPBs (Fall et al., 2008).
6.0
5.0 3.5. Predicting the time-dependent evolution of the saturated
4.0 hydraulic conductivity of CPBs
3.0
2.0 Engineers and mining operators need analysis tools that can
1.0 give indications on how the CPB (an evolutive porous medium)
0.0 hydraulic conductivity will change with time. Those tools and indi-
0 0.05 0.1 0.15 0.2 0.25 0.3 cations will be useful during the preliminary design phase of a CPB
Mechanical damage (D)
structure for verifying the validity of the experimental results, and
Fig. 10. Evolution of the relative hydraulic conductivity (k/k0) of various CPBs with
also estimating the possible variation of permeability that can be
the (a) ratio of the maximum applied stress to peak stress (applied stress/UCS) and expected during the service of a given CPB structure. In turn, this
(b) mechanical damage (D). will facilitate the decision making process in the final design of a
Author's personal copy

1316 M. Fall et al. / Minerals Engineering 22 (2009) 1307–1317

in Fig. 11. It can be noticed that there is a good agreement between


Adrien (2008)
7.0E-05
the predicted and experimental values. This equation can thus be
Mukesh (2008)
R² = 0.97
considered a useful tool for the estimation of the saturated hydrau-
Godbout (2005)
6.0E-05 lic conductivity of CPB.
Celestin (2008)
Predicted k (cms)

5.0E-05 Trendline
4. Summary and conclusions
4.0E-05
This study has demonstrated that the hydraulic conductivity of
3.0E-05 CPBs is significantly influenced by mix components (binder type
and proportions, w/c ratio, tailings fineness, and sulphate content),
2.0E-05
the curing time and temperature, stress level applied on the CPBs
1.0E-05 and mechanical damage.

1.0E-08
0.0E+00 1.0E-05 2.0E-05 3.0E-05 4.0E-05 5.0E-05 6.0E-05 7.0E-05
 The hydraulic conductivity of the CPB decreases as the w/c ratio
Measured k (cm/s)
decreases, while it increases as the binder content decreases.
Fig. 11. Predicted versus measured k-values for various CPBs. The type of binder used has an impact on the hydraulic conduc-
tivity of CPBs. The CPBs made of coarse tailings (20% and 32%
fineness) generate higher permeability for a given consistency.
CPB structure, but not eliminate the need for testing. For these rea- Depending on the curing time and sulphate concentration, the
sons, a simple predictive model for saturated hydraulic conductiv- sulphate can have two opposite effects on the permeability of
ity is developed in this research to estimate the time-dependent CPBs, contributing to an increase or decrease. However, the
evolution of hydraulic conductivity of undamaged CPBs. To keep magnitude of the influence of CPB’s mix components on the
the paper at a reasonable length, only the key points of the model hydraulic conductivity is strongly influenced by the curing time;
development and validation will be presented here. Other details this effect is more pronounced at early ages (67 days).
on the model development and validation are presented elsewhere
 The hydraulic conductivity of CPB is time-dependent. It
(Adrien, 2008).
decreases as the curing time increases. The highest decrease of
The UCS (unconfined compressive strength) is undoubtedly the
the permeability of CPBs with curing time occurs at early ages
geotechnical property that is most often evaluated in mine cemen-
of curing (0–7 days).
ted backfill practice. Furthermore, assuming that the development
of the UCS of CPB is an adequate indication of its binder hydration  The hydraulic conductivity decreases with curing temperature
degree and microstructure (pore size distribution, porosity) devel- and time for the studied CPBs. The results show also that the
opment and the latter strongly affects the hydraulic conductivity of changes in hydraulic conductivity with temperature are more
CPB, a direct relationship may exist between UCS and the perme- significant in early age samples (up to 7 days). However, the
ability of CPB at any curing time. Hence, using the experimental effect of curing temperature on the hydraulic conductivity of
data from this study with regards to the permeability of CPBs CPBs depends on the binder type.
and from UCS tests (Celestin, 2008; Pokharel, 2008) performed  This study reveals that the applied stress level and mechanical
on CPB samples having the same mix composition as those tested damage influence the hydraulic conductivity of CPBs. Under
in this study, Eq. (2) was formulated (based on regression analysis low to medium stress levels that correspond to 50–60% of the
of the data sets by using Excel software) to predict the time-depen- UCS, the hydraulic conductivity tends to decrease slightly and/
dent evolution of the permeability (k, [lT1]) of CPBs: or then remains relatively constant. Beyond the aforementioned
threshold stress level, the permeability of CPB starts to increase,
K ¼ kT  A  ðUCSt =UCSmax ÞB ð2Þ first very slightly, then significantly from about 80% of the UCS
or a mechanical damage value of 0.06 for the high strength
where kT [lT1] is the saturated hydraulic conductivity of the tailings CPB specimens and 0.15 for the low strength CPB specimens.
used, UCSt (kPa) of the CPB for a given time, UCSmax (kPa) is the max-
imal UCS of the CPB (i.e. the UCS of the mature CPB); and A and B are A simple function is proposed for predicting the evolution of the
adimensional fitting parameters to be determined for each CPB mix. hydraulic conductivity of CPBs. The proposed function can be use-
The desired UCS values can be obtained by performing uniaxial com- ful during the preliminary design phase of a CPB structure, for ver-
pression test (according to ASTM C 39) on the CPB samples. kT [lT1] ifying the validity of the experimental results, and also to
value can be determined by conducting saturated hydraulic conduc- estimating the possible variation of permeability that can be ex-
tivity tests on the tailings used for the preparation of the CPB. More- pected during the service of a given CPB structure.
over, the value of kT can be estimated from a suitable empirical This study will provide useful information for understanding
relationship for predicting the saturated hydraulic conductivity of the fluid transport ability of CPBs. The authors believe that the out-
fine soils and/or tailings (e.g., Chapuis and Aubertin, 2003). come of this research will help mining operators and the engineer-
To verify and demonstrate the applicability of the model, the ing community working in the CPB system field for consideration
developed predictive model was validated against experimental in the design, environmental and durability evaluation of the
data resulting from additional (new) hydraulic and UCS tests per- aforementioned structures, and will also help in predicting the po-
formed on various CPBs in this study. Furthermore, to better eval- tential formation of AMD in CPB systems.
uate the capabilities of the developed model, hydraulic
conductivity and UCS tests conducted by other researchers (e.g., Acknowledgement
Celestin, 2008; Pokharel, 2008; Godbout, 2005) on CPB samples
with various mix components (binder contents and types, sulphate The writers would like to acknowledge the National Sciences
contents, w/c ratio, fineness) and cured at various times (7– and Engineering Research Council of Canada (NSERC), the Univer-
90 days) and temperatures (20–50 °C) were also used for valida- sity of Ottawa, Cement Lafarge Inc., and the National Research
tion purpose. The main results of the validation are summarized Council (NRC) of Canada.
Author's personal copy

M. Fall et al. / Minerals Engineering 22 (2009) 1307–1317 1317

References Hooton, R.D., 2000. Canadian use of ground granulated blast-furnace slag as
supplementary cementing material for enhanced performance of concrete.
Canadian Journal of Civil Engineering 27, 754–760.
Adrien, D., 2008. Experimental observations and numerical modeling of saturated
Jensen, O.M., Hansen, P.F., 2002. Water-entrained cement-based materials: II.
hydraulic processes in cemented backfill structures. Research Report, University
Experimental observations. Cement and Concrete Research 32, 973–978.
of Ottawa, 58p.
Kesimal, A., Ercikdi, B., Yilmaz, E., 2003. The effect of desliming by sedimentation on
Benzaazoua, M., Fall, M., Belem, T., 2004. A contribution to understanding the
paste backfill performance. Journal Minerals Engineering 16, 1009–1011.
hardening process of cemented paste fill. Minerals Engineering 17 (2), 141–
Landriault, D. 1995. Paste backfill Mix design for Canadian Underground Hard Rock
152.
Mining. In: Proceedings of the 97th Annual General Meeting of CIM. Rock
Celestin, J.C., 2008. Geotechnical properties of cemented paste backfill and tailings
Mechanics and Strata Control Session. Halifax, Nova Scotia, pp. 229–238.
liners: effect of mix components and temperature. Master Thesis (M.A.Sc.),
Levens, R.L., Marcy A.D., Boldt, C.M.K., 1996. Environmental Impacts of Cemented
University of Ottawa, 222p.
Mine Waste Backfill. RI 9599. United States Bureau of Mines, 23p.
Celestin, J.C., Fall, M., 2009. Thermal conductivity of cemented paste backfill
Lothenbach, B., Winnefeld, F., Alder, C., Wieland, E., Lunk, P., 2007. Effect of
materials and factors affecting it. International Journal of Mining Reclamation
temperature on the pore solution, microstructure and hydration products of
and Environment 2, 1–17.
Portland cement pastes. Cement and Concrete Research 37, 483–491.
Chapuis, R.P., Aubertin, M., 2003. On the use of the Kozeny–Carman equation to
Manmohan, D., Mehta, P.K., 1981. Influence of pozzolanic slag, and chemical
predict the hydraulic conductivity of soils. Canadian Geotechnical Journal 40,
admixtures on pore size distribution and and permeability of hardened cement
616–628.
pastes. Journal of Cement, Concrete and Aggregates (CCA) 3, 63–67.
Demirboga, R., 2007. Thermal conductivity and compressive strength of concrete
Mindess, S., Young F., 1981. Concrete. Prentice-Hall Inc., Englewood Cliffs, New
incorporation with mineral admixtures. Building and Environment 42, 2467–
Jersey 07632, 671p.
2471.
Orejarena, L., Fall, M., 2008. Mechanical response of a mine composite material to
Doven, A.G., Pekrioglu, A., 2005. Material properties of high volume fly ash cement
extreme heat. Journal of Engineering Geology and Environment 67, 387–396.
paste structural fill. Journal of Materials in Civil Engineering 17, 686–693.
Palmer, B.G., Edil, T.B., Benson, C.H., 2000. Liners for waste containment constructed
Fall, M., Benzaazoua, M., 2005. Modeling the effect of sulphate on strength
with class F and C fly ashes. Journal of Hazardous Materials 76, 193–216.
development of paste backfill and binder mixture optimization. Cement
Picandet, V., Khelidj, A., Bastian, G., 2001. Effect of axial compressive damage on as
Concrete Research 35, 301–314.
permeability of ordinary and high-performance concrete. Cement and Concrete
Fall, M., Benzaazoua, M., Ouellet, S., 2005. Experimental characterization of the
Research 31, 1525–1532.
effect of tailings fineness and density on the quality of cemented paste backfill.
Pokharel, M. (2008). Geotechnical and environmental response of paste tailings
Journal Minerals Engineering 18, 41–44.
systems to coupled thermo-chemical loadings. Master Thesis (M.A.Sc.),
Fall, M., Nasir, O., Celestin, J.C., 2007. Paste backfill response in deep mine
University of Ottawa, 248p.
temperature conditions. In: Symposium MineFill, Montreal, Canada, CD-Room.
Ritcey, M.G., 2005. Tailing management in gold plants. Journal of Hydrometallurgy
Fall, M., Benzaazoua, M., Saa, E., 2008. Mix proportioning of underground cemented
78, 3–20.
paste backfill. International Journal of Tunnelling and Underground
Taylor, H.F.W., 1993. The Chemistry of Cements. Academic Press, London. 475p.
Construction 23, 80–90.
Garboczi, E.J., 1990. Permeability, diffusivity, and microstructural parameters: A Wang, K., Shah, S.P., Mishulovich, A., 2004. Effects of curing temperature and NaOH
critical review. Cement and Concrete Research 20 (4), 591–601. addition on hydration and strength development of clinker-free CKD-fly ash
Godbout, J. 2005. Evolution des proprieties des remblais miniers cimentés en pate binders. Journal of Cement and Concrete Research 34, 299–309.
durant le curage. Mémoire de Maitrise, Ecole Polytechnique de Montréal, 212p. Yilmaz, E., Kesima, A., Ercidi, B., 2004. Strength development of paste backfill
Hassani, F., Archibal, J., 1998. Mine Backfill, CD-Rom. Canadian Institute of Mine, simples at Long term using different binders. In: Proceedings of 8th Symposium
Metallurgy and Petroleum. MineFill04, China, pp. 281–285.

You might also like