You are on page 1of 8

Energy Geotechnics – Wuttke, Bauer & Sánchez (Eds)

© 2016 Taylor & Francis Group, London, ISBN 978-1-138-03299-6

Thermomechanical properties of a new small-scale reinforced


concrete thermo-active pile for centrifuge testing

A. Minto, A.K. Leung, D. Vitali & J.A. Knappett


University of Dundee, Dundee, Scotland, UK

ABSTRACT: The use of thermo-active geo-structures has been recognised to be a sustainable engineering
solution that can reduce carbon emissions from civil infrastructure. Physical modelling in a geotechnical cen-
trifuge has been increasingly used to study the behaviour of this kind of geo-structures and their interaction
with the surrounding soil under cyclic heating/cooling loads. Previous studies have been limited by the choice
of materials used to model thermo-active geo-structures (such as aluminium and conventional concrete), due
to inaccurate scaling of the thermal properties and the inability to capture the quasi-brittleness and strength
properties of reinforced concrete (RC) due to improper scaling of aggregate sizes. The paper aims to develop a
new thermally-enhanced plaster-based model concrete which can realistically reproduce both the thermal and
mechanical properties that are representative of concrete at prototype scale. The new model concrete, combined
with steel wire reinforcement (i.e. geometrically scaled reinforcing bars and stirrups), was then used to create
1:20 scaled RC thermo-active piles. Effects of temperature on their thermomechanical behaviour, including the
coefficient of thermal expansion, moment capacity and flexural stiffness, were investigated. The suitability of
using the newly-developed RC thermo-active piles for future centrifuge testing is discussed.

1 INTRODUCTION internal diameter, and hence the second moment of


area, of the model pile. In terms of thermal proper-
Thermo-active geo-structures such as pile foundations ties, aluminium has a coefficient of thermal expansion
and diaphragm walls have been used to exploit shal- (CTE) of 22.2 με/◦ C (where μ is micron; and ε is
low geothermal energy to provide heating and cooling thermally-induced axial strain). In their aluminium
to buildings (Brandl, 2006). A number of field tests piles, an epoxy resin with a thickness of 1.5 mm had to
(Laloui et al., 2006, Bourne-Webb et al., 2009, Murphy be applied on the pile shaft to protect the strain gauges.
and McCartney, 2015) have been conducted to investi- As the CTE of epoxy is 85 με/◦ C (Rotta Loria et al.,
gate the thermomechanical behaviour of these types of 2015), the CTE of the overall instrumented aluminium
structures and their interaction with the surrounding pile should be higher than 22.2 με/◦ C, which is thus
soil. It was demonstrated that cyclic heating/cooling at least 30% higher than the CTE of unrestrained RC
processes introduced additional stresses and strains to (15–16 με/◦ C; Goode III et al., 2015). This makes it
the thermo-active geo-structures. In order to improve difficult for this kind of aluminium pile to realistically
understanding of the soil-structure interaction in more capture the soil-structure interaction of thermo-active
detail, small-scale physical model tests in a geotech- piles that are made of RC in the prototype. Numerical
nical centrifuge have been conducted (Stewart and studies (Yavari et al., 2014) have shown that the coef-
McCartney, 2014, Ng et al., 2015). By elevating the ficient of thermal expansion of a thermo-active pile
acceleration of a physical model by N times Earth’s plays a significant role in affecting the mobilisation
gravity (i.e., N -g), the stress levels experienced by and distribution of pile axial loads under a combina-
the soil in a small-scale physical model are simi- tion of mechanical and thermal loadings. Moreover,
lar to that in a full scale prototype, enabling the this type of elastic aluminium pile is not able to mimic
stress dependency of the soil behaviour to be correctly the nonlinear quasi-brittleness feature of RC under ten-
captured. A major challenge of testing a thermo- sile and bending stresses such as when subject to lateral
active geo-structure in the centrifuge is the selection loading conditions (Karihaloo and Huang, 1991).
of appropriate model materials that model both the Stewart and McCartney (2014) modelled a thermo-
mechanical and thermal properties realistically and active pile using RC with scaled steel reinforce-
simultaneously. ment and scaled aggregates in their centrifuge tests.
In the centrifuge tests performed by Ng et al., Although the thermal properties of prototype RC
(2015), aluminium alloy was used to model a thermo- might be properly scaled using RC as a model pile,
active pile. Scaling of the mechanical properties such the use of gravel with a maximum diameter of 6 mm
as the axial or flexural rigidity of the aluminium (in model scale) to model coarse aggregates could pose
pile can be achieved through careful selection of the a scaling problem of the mechanical properties. At 24g

37
which Stewart and McCartney (2014) adopted in their
centrifuge tests, the prototype diameter of the gravel
was 144 mm, which was 7 times larger than the diam-
eter of coarse aggregates in prototype concrete. This
size effect would potentially lead to an over-strength
of RC (Litle and Paparoni, 1966, Belgin and Sener,
2008). This modelling method is thus limited to rela-
tively low scaling factors (or g-level) to minimise any
over-strength of RC.
In order to more realistically capture the nonlin-
ear quasi-brittleness feature and failure mechanisms
of concrete in centrifuge at higher scaling factors,
Knappett et al. (2011) developed model concretes
using plaster-based mortars. In the mortar mix, fine sil-
ica sand was used to geometrically scale the aggregate Figure 1. Particle-size distribution of silica sand and copper
found in concrete. Such model concrete was shown to powder.
have representative mechanical strengths, in terms of
unconfined compressive strength and modulus of rup- distribution of the silica sand and copper powder were
ture. The model concrete has been successfully used measured using a laser diffraction analyser. The results
for modelling various engineering structures such as are compared in Fig. 1. It can be seen that both the sand
piles (Al-Defae and Knappett, 2014) and bridge piers and copper powder were uniformly graded, and that the
(Loli et al., 2014) in the centrifuge where simultane- size of copper powder was finer than that of the sand.
ous modelling of stiffness and strength are crucial.
For modelling thermo-active RC geo-structures, this
type of model concrete requires further modification 2.2 Thermomechanical properties of the new
to ensure correct scaling of the thermomechanical model concrete
properties.
In order to investigate the effects of the copper pow-
This study aims to develop a new type of model
der content on the thermal conductivity of the model
concrete that can realistically scale the mechanical
concrete, a series of laboratory testing was carried out
and thermal properties of real concrete for future
using a hot-box apparatus developed by Jones et al.
centrifuge testing of concrete energy geo-structures
(2007). It is an apparatus that can create a temperature
for larger scaling factors. An application of the new
gradient across a slab-shaped specimen (45 mm width,
model concrete to produce RC thermo-active piles
150 mm long and 150 mm height) and also can mea-
is presented. Temperature effects on thermomechan-
sure the corresponding heat flux. At the steady state,
ical properties, including the coefficient of thermal
the thermal conductivity of the specimen can be deter-
expansion, moment capacity and flexural stiffness, of
mined by dividing the heat flux by the temperature
a model pile were tested. By comparing the model and
gradient, according to Fourier’s law.
prototype properties, the suitability of testing this type
Model concrete mixed with five different percent-
of new RC model pile in centrifuge is discussed.
ages of copper powder (by volume), 0%, 1.5%, 3%,
6% and 12% were tested. The test results depicted in
Fig. 2 show that the thermal conductivity of the model
2 NEW MODEL CONCRETE
concrete originally designed by Knappett et al. (2011)
was 0.4 W/(m·K), which was lower than the typical
2.1 Constituents of the model concrete
range of concrete (i.e., 0.9 to 1.1 W/(m·K); Kanbur
The new model concrete developed in this study et al. (2013)). When copper powder was added, there
is based on the design previously proposed by was almost a linear increase in the thermal conductiv-
Knappett et al. (2011). The original design consisted of ity with the amount of copper added. This shows that
a mixture of β-form surgical plaster (manufactured by the addition of copper powder was effective to enhance
Saint Gobain), water and fine silica sand (Congleton the thermal properties of the model concrete. In par-
HST95). The sand was used to geometrically scale and ticular, 6% and 12% copper powder contents appeared
approximate the size of aggregates found in concrete. to match the prototype range reasonably well.
Knappett et al. (2011) suggested that a water/plaster Although adding copper powder could significantly
(W/P) ratio of 0.9:1 and a sand/plaster (S/P) ratio of improve the thermal properties of the new model con-
1:1 would result in a model concrete that can realisti- crete, one concern is any detrimental effects of such
cally mimic the mechanical properties of concrete in addition on the mechanical properties. For this pur-
prototype. pose, a series of four-point bending (FPB) tests were
In order to properly scale and mimic the thermal conducted to measure the modulus of rupture (fr ) of
properties of prototype concrete, a new constituent, prismatic specimens (25 × 25 × 250 mm) when differ-
copper powder (manufactured by Phoenix Scientific), ent percentages of copper powder were added to the
was added to the design mix to enhance the thermal model concrete. The testing procedures outlined by
conductivity of the model concrete. The particle-size Knappett et al. (2011) were adopted.

38
due to the rapid evolution of the pore structure during
the hydration process (Song et al., 2009).
Fig. 4(c) shows the model concrete mix at × 300
where both the sand and copper particles are visible.
There was a gap between the irregular-shaped sand
particles and plaster, hence creating some weakened
interfaces. However, such gapping was not found for
the copper particles. It must be pointed out that no
chemical bonding was formed along the plaster, sand
and copper particles after the hydration process. These
constituents were bonded physically through weak
inter-particle van der Waal force. Such physical bond
is an important feature of the model concrete to real-
istically mimic the non-linear quasi-brittleness nature
Figure 2. Effects of copper powder content on thermal con- of concrete in prototype, which cannot be captured
ductivity of model concrete. Error bars represent standard by existing elastic model piles made of aluminium,
errors (n = 3). and which avoids the potential over-strength of using
cement as the binder.

3 NEW MODEL RC THERMO-ACTIVE PILE

3.1 Structural analysis and design


Before producing a RC thermo-active pile, a detailed
structural analysis and design was carried out. In
future centrifuge tests, model thermo-active piles with
a square cross-section will be used to stabilise a silty
slope in a 1:20 scale model tested at 20 g. Preliminary
analysis and a literature review of previous studies
using piles to reinforce soil slopes (Al-Defae and
Figure 3. Effects of copper powder content on fr of the new Knappett, 2014, Hayward et al., 2000) suggests that
model concrete. Error bars represent standard errors (n = 6). a moment capacity of 250 kNm would be sufficient
for the piles to maintain the slope stability for repre-
Fig. 3 shows the variations of copper powder con- sentative soil properties and slope angles. Based on the
tent with normalised fr (by the values obtained from targeted moment capacity, the pile reinforcement was
0% copper powder content, fr0 ). It is clear that the designed according to the design procedures outlined
addition of copper powder up to 12% did not cause in Eurocode 2: Design of Concrete Structures (BSI,
significant changes in the mechanical properties. The 2004). Hence, at model scale, the thermo-active pile
addition of copper powder could thus enhance the will have a square cross section with the side of 25 mm
thermal conductivity of the model concrete, while and a length of 250 mm. In the following discussion,
simultaneously maintaining its ability to mimic the all dimensions are expressed in model scale, unless
mechanical properties of concrete. stated otherwise.
The arrangement of the reinforcement is detailed in
Fig. 5. It was a doubly-reinforced concrete pile, since
2.3 Micro-structure of the new model concrete
to allow bending in either or both horizontal direc-
In order to have a better understanding of the inter- tions in future centrifuge testing. All the longitudinal
action between the plaster, sand grains and copper reinforcement bars chosen were 1.25 mm in diameter,
powder as well as the uniformity of these constituents while the diameter of shear reinforcement (aka stir-
within the mixture, the micro-structure of the new rups) was 0.6 mm. The thickness of concrete cover was
model concrete was imaged by a scanning electron 2 mm. The total area of steel reinforcements (As ) was
microscope (SEM). Fig. 4(a) shows an SEM image at 12.93 mm2 for a cross-sectional area of 625 mm2 of the
a magnification of × 55. The image depicts some sil- model pile. This means that the reinforcement ratio,
ica sand physically interlocked with the flake-shaped As /bd (where b and d is the breath and the depth of the
plastic matrix. The distribution of the sand appears model pile, respectively), is 2.1%, which falls within
uniform. a typical range found in high-strength doubly rein-
At a higher magnification of ×500 (Fig. 4(b)), forced concrete beams under flexural loads (Rashid
the fine copper particles can be viewed. They were and Mansur, 2005).
spherical in shape and physically embedded into the The model reinforcement used in this study was
fibre-like irregular structures of the plaster. This kind a stainless steel (Grade 316), which was manufac-
of micro-structure is a key feature of the β-form plaster tured by Ormiston Wire Ltd. It has a yield strength

39
Figure 5. Cross-section and reinforcement details of a 1:20
model RC thermo-active pile.

Figure 6. Overview of the formwork used to cast two model


piles.

reinforcement bars and stirrups were coated with


epoxy resin and then immersed in a pool of HST95
silica sand for 24 hours to ensure an intimate mechan-
ical frictional contact between the reinforcements and
the sand, preventing the reinforcements from slipping
inside the model concrete.
In order to allow water to be circulated inside
the new model concrete, two pairs of closed-loop
U-shaped silicon pipes were attached to the internal
surface of the stirrups (Fig. 5), which is a typi-
cal arrangement for a thermo-active pile prototype
(Loveridge and Powrie, 2013). Placing the pipes closer
to the pile perimeter aimed to minimise their thermal
interaction and the heat loss to the concrete material,
hence achieving a more uniform distribution of pile
temperature. Cecinato and Loveridge (2015) showed
that circulation pipes spaced too closely can reduce
the heat transfer efficiency to the surrounding soil.
Figure 4. SEM images of at (a) ×55; (b) ×500; and The diameter of the silicon pipe was 3 mm, which was
(c) ×300 magnification. the minimum internal size of the pipe required to min-
imise the thrust force caused by the change in water
of 460 MPa (Knappett et al. 2011) and a coefficient of flow direction at the U-turn of the pipe (see Fig. 6).
thermal expansion of 18.5 με/◦ C (Lee et al., 2007). Three thermocouples were attached to both ends
and the centre of the reinforcement bar for monitoring
the temperature distribution along the pile.
3.2 Model preparation
When the reinforcements, silicon pipe and ther-
Formwork with a dimension identical to the model mocouples were fixed in position, the mix of model
pile (25 mm × 25 mm × 250 mm; Fig. 6) was con- concrete (including plaster, water, silica sand and cop-
structed to cast the model piles. Before casting, all per powder) was poured into the formwork. The mix

40
was allowed to cure for 28 days in a room maintained
at an ambient temperature of around 20◦ C.
In order to heat and cool the model pile, a heating
system (Julabo Ltd; Model F12) that can control and
maintain constant water temperature between 1 and
99◦ C was connected to the silicon pipes. Change in
pile temperature was monitored by the three thermo-
couples embedded in the model pile.

4 TESTING OF THE MODEL PILE

4.1 Test plan


Laboratory testing was undertaken to determine the Figure 7. Typical setup of a FPB test for model pile.
thermo-mechanical properties of the small-scale RC
thermo-active pile made by the new model concrete. The CTE tests were started by submerging each
The primary aim was to ensure the design of the model model pile in a water bath with a controlled temper-
pile was mechanically and thermally representative ature of 50◦ C for at least two hours. When a uniform
of the prototype RC thermo-active pile. An impor- distribution of pile temperature of about 50◦ C was
tant thermal property that needed to be characterised reached at the steady state (as indicated by the three
was the coefficient of thermal expansion (CTE). Upon thermocouples), each model pile was removed from
pile temperature changes, the pile expansion (upon the water bath and any change in pile length between
heating) and contraction (upon cooling) would have the two reference points was measured immediately
direct effect on the soil-pile contact, hence affecting using a calliper with an accuracy of 0.01 mm. The
the soil-pile interaction and the mobilisation of pile measurements were taken as quickly as possible to
axial load (Yavari et al., 2014). minimise heat loss from each model pile.
Another test series aimed to determine any tem- For the measurements of the pile moment capac-
perature dependency on the flexural properties, which ity, the four-point bending (FPB) testing method was
is a key thermo-mechanical property that governs the adopted (Balendran et al., 2002). Two pile tempera-
pile behaviour when the piles are subjected to lateral tures were considered: circulating water at ambient
loading in slope stabilization applications. temperature (20◦ C) and an elevated level of 50◦ C.
In both types of tests, effects of the percentage At each pile temperature, model piles cast with three
of copper powder addition were studied. This aimed different amount of copper powder (0%, 6% and 12%
to determine any optimum amount of copper powder by volume) were tested. 17 FPB tests were carried out
that could realistically scale both the mechanical and (7 tests at 0%, 4 tests at 6% and 6 tests at 12%).
thermal properties of the model pile simultaneously. A typical set up for a FBD test is shown in Fig. 7.
The model pile was roller-supported, having a span
4.2 Test procedures (L) of 210 mm. Vertical loading was applied symmet-
rically at two locations, both of which were 35 mm
The testing method of CTE followed a modification of away from the centre of the pile to ensure that the
the procedures outlined in AASHTO TP 60. CTE was loads and reactions were spaced at L/3 as per typical
calculated by the change in the length of a model pile FPB testing procedures (Huurman and Pronk, 2009).
when there is an increase in pile temperature. After By using the heating system, water was controlled at a
casting the model piles, they were coated with a sil- constant temperature of 20 or 50◦ C and was circulated
icon conformal coating spray to ensure a waterproof to the model pile. When all thermocouples showed
condition, as would be required in future centrifuge a temperature matching the targeted value set on the
tests. Then, two reference points were defined in each heating system, test began by loading the model pile at
model pile to determine their initial pile length. The a rate of 0.6 mm/min. The test was complete when the
two reference points were fixed along the axis of the pile section exhibited large deformation and formed
two different longitudinal sides, which are orthogo- prominent cracking.
nal with respect to each other. From Fig. 5, ‘side 1’
refers to the four longitudinal reinforcement bars laid
in parallel, while ‘side 2’ refers to the adjacent side,
5 RESULTS AND DISCUSSION
where no longitudinal reinforcement bars were added.
By comparing the test results obtained from sides 1
5.1 Coefficient of thermal expansion
and 2, any effects of heat-induced elongation of the
reinforcement bars on the overall length of the model Table 1 summarises the measured values of CTE for
pile may be quantified. In total, three tests were con- the Piles A to C. It is found that the CTE of the
ducted, one with no copper powder added (Pile A), and model pile without copper powder (i.e., Pile A) was
two replications with 6% copper powder content (Piles 15.7 με/◦ C, which was close to the values (<3% dif-
B and C). ference) obtained from side 2 of Piles B and C when

41
Table 1. Summary of the CTE test.

Pile Side Copper powder (%) CTE (με /◦ C)

A 1 0 15.7
B 1 6 16.5
2 15.3
C 1 6 16.0
2 15.5

6% of copper powder was added. This is because the


CTE of copper (i.e., 17.5 με/◦ C; (Davis, 2001)) was
not very significantly different than that of the plaster
(13.9 με/◦ C; (Hyer, 2009)) and quartzite (i.e. sand par-
ticles), which ranged from 7.0 to 13.5 με/◦ C. (Johnson
and Parsons, 1944). This suggests that the addition of
copper powder did not introduce significant effects on
the CTE.
It can also be seen in Table 1 that the CTE obtained
from side 1 (where reinforcement bars were present)
was slightly higher than that on side 2, by not more than
8%. This was somewhat expected because the CTE
of plaster (i.e., 13.9 με/◦ C) did not differ much from
that of the steel reinforcement used (i.e., 18.5 με/◦ C).
By taking the average measured values obtained from
sides 1 and 2, the CTE of Pile B (15.9 με/◦ C) was
close to that of Pile C (15.8 με/◦ C). These average
Figure 8. Effect of copper powder on (a) moment capacity,
values match well with the typical range of CTE of (b) flexural stiffness of model piles at 20◦ C and 50◦ C. Error
unreinforced concrete (i.e., 9.4 to 14.4 με/◦ C; depend- bars represent standard error.
ing on the aggregate mineralogy; Choi and Chen,
2005). Goode III et al., (2015) reported that the added (Fig. 3(b)). This implies that the presence of the
measured CTE of their RC pile (which was made silicon pipes and the water being circulated in model
of conventional concrete) under unrestrained con- pile played a significant role on the thermo-mechanical
dition (i.e., free to expand upon pile heating) was responses. The CTE of silicon is ∼2.3 με/◦ C (Vasiliev,
15–16 με/◦ C, which was also very close to the aver- 2008), while that of water ranges from 200 με/◦ C at
age value of the model RC thermo-active pile created 20◦ C to 450 με/◦ C at 50◦ C (Kell, 1975). This sug-
in this study. gests that the thermal expansion of the water has a
greater influence than the expansion of the silicon
pipes. When the copper powder content is higher, the
5.2 Temperature effects on moment capacity and
thermal expansion of the copper particles, silicon pipe
flexural stiffness
and the circulating water might have created more
In this study, pile failure was defined at the onset of weakened interfaces, hence reducing the Mult of the
the first observable flexural cracking on the surface model pile.
of the model pile. The rationale of defining this fail- Flexural stiffness (EI, where E and I are theYoung’s
ure criterion is that as flexural cracking develops, the modulus and the moment of inertia of the un-cracked
model pile would no longer be serviceable as an energy section of a model RC, respectively) can be obtained
geo-structure. This is because the embedded pipes, by determining the gradient of the linear portion of a
which have limited flexural resistance, would be split, moment-curvature plot. The linear portion is defined
causing leakage of the circulating water. as the range of Mult from zero to a value where the
Fig. 8(a) shows the effects of copper powder first crack was observed in the model piles. Effects of
percentage on normalised ultimate bending moment copper powder and pile temperature on normalised EI
(Mult ) at different pile temperatures. At both tempera- are shown in Fig. 8(b). At pile temperature of 20◦ C, the
tures, there were not very substantial changes in Mult copper powder percentage has negligible effects on the
as the amount of copper powder increases from 0 to EI. In contrast, at an elevated level of 50◦ C, an increase
6%, given the presence of sample variability. When the in copper powder percentage causes a significant drop
percentage of copper powder increased further to 12%, of EI.
significant drops of Mult were found. The drop under Based on the test results presented in this paper,
the elevated pile temperature at 50◦ C was much more it appears that 6% was an optimum percentage of
substantial (up to 86%). Interestingly, such a drop was copper powder that could provide a model RC thermo-
not observed in plain model concrete, where fr was active pile with a reasonably close CTE (Table 1)
almost unaffected by the percentage of copper powder and thermal conductivity (Fig. 2) when compared to

42
the properties of a prototype thermo-active pile. At Al-Defae, A. H. & Knappett, J. A. 2014. Centrifuge Mod-
6% copper powder content, the effects of temper- eling of the Seismic Performance of Pile-Reinforced
ature on the mechanical properties, including both Slopes. Journal of Geotechnical and Geoenvironmental
Mult and EI may be practically negligible (see Fig. Engineering, 140.
Balendran, R., Zhou, F., Nadeem, A. & Leung, A. 2002.
8). This is consistent with the BS EN 1992-1-2:2004 Influence of steel fibres on strength and ductility of nor-
Clauses 3.2.2 and 3.2.3, which respectively state that mal and lightweight high strength concrete. Building and
any temperature effects on the mechanical properties environment, 37, 1361–1367.
of concrete and steel are negligible when the tempera- Belgin, C. M. & Sener, S. 2008. Size effect on failure
ture is below 100◦ C.A study reported by Li and Purkiss of overreinforced concrete beams. Engineering Fracture
(2005) also shows that theYoung’s modulus of concrete Mechanics, 75, 2308–2319.
is reasonably constant for temperatures between 20 Bourne-Webb, P. J.,Amatya, B., Soga, K.,Amis, T., Davidson,
and 60◦ C. C. & Payne, P. 2009. Energy pile test at Lambeth College,
London: geotechnical and thermodynamic aspects of pile
response to heat cycles. Geotechnique, 59, 237–248.
6 CONCLUSION Brandling, H. 2006. Energy foundations and other thermo-
active ground structures. Geotechnique, 56, 81–122.
A new type of model concrete that is highly suitable British Standards Institution. 2004. Eurocode 2: Design of
Concrete Structures. London, BSI.
for realistically modelling thermo-active energy geo- BS EN 1992-1-2:2004: Eurocode 2: Design of concrete struc-
structures in centrifuge testing is developed. The new tures. Part 1.2: General rules – Structural fire design.
model concrete is a mixture of plaster, silica sand, British Standards Institution, London, 2004.
water and copper powder, which results in a material Cecinato, F. & Loveridge, F. A. 2015. Influences on the ther-
that has both mechanical and thermal properties scaled mal efficiency of energy piles. Energy, 82, 1021–1033.
simultaneously. The test results showed that adding 6% Choi, J. H., & Chen, R. H. L. 2005. Design of continu-
or 12% copper powder content to the new model con- ously reinforced concrete pavements using glass fiber
crete could increase the thermal conductivity to a value reinforced polymer rebars. Publication No. FHWA-HRT-
close to those found in prototype concrete. Such an 05-081. Washington, D.C.
Davis, J. R. 2001. Copper and copper alloys, ASM interna-
addition was shown not to cause significant changes tional.
in modulus of rupture. Goode III, J. & McCartney, J. 2015. Centrifuge modeling
This study also demonstrated an application of this of end-restraint effects in energy foundations. Journal of
type of new model concrete to create 1:20 model rein- Geotechnical and Geoenvironmental Engineering, ASCE,
forced concrete (RC) thermo-active piles. The effects DOI: 10.1061/(ASCE)GT.1943-5606.0001333.
of copper powder content (ranging from 0% to 12%) Hayward, T., Lees, A., Powrie, W., Richards, D. & Smethurst,
on the thermomechanical behaviour of the model RC J. 2000. Centrifuge modelling of a cutting slope stabilised
pile were quantified. It was found that copper powder by discrete piles, Transport Research Laboratory.
content of 6% is an optimum amount that would pro- Huurman, M. & Pronk, A. 2009. Theoretical analysis of the 4
point bending test.AdvancedTesting and Characterization
vide the model pile with a highly representative value of Bituminous Materials, A. Loizos, MN Partl, T. Scarpas,
of the coefficient of thermal expansion as compared and IL Al-Qadi, eds., CRC Press, Boca Raton, 749–759.
to prototype. At this percentage of addition, effects of Hyer, M. W. 2009. Stress analysis of fiber-reinforced
pile temperature between 20 and 50◦ C on pile bending composite materials, DEStech Publications, Inc.
moment capacity and flexural stiffness are practically Johnson, W. H. & Parsons, W. H. 1944. Thermal expansion
negligible. This correctly models the thermomechani- of concrete aggregate materials, US Government Printing
cal behaviour of prototype RC piles. Such a model RC Office.
thermo-active pile was also capable of mimicking the Jones, M. R., Zheng, L., McCarthy, A., Dhir, R. K., &
nonlinear quasi-brittle nature of real concrete, which is Yerramala A. 2007. Increasing the use of foamed concrete
incorporating recycled and secondary aggregates. WRAP
a key feature that is not achievable using elastic model Project Report:AGG79-001.
piles. Kanbur, B. B., Atayilmaz, S. O., Demir, H., Koca, A. &
Gemici, Z. 2013. Investigating the thermal conductivity
of different concrete and reinforced concrete models with
ACKNOWLEDGEMENTS
numerical and experimental methods. Recent Advances in
Mechanical Engineering Applications, Recent Advances
The authors would like to acknowledge the stu- in Mechanical Engineering Series, 95–101.
dentships and the research cost supported by the Karihaloo, B. L. & Huang, X. 1991. Tensile response of
Energy Technology Partnership (ETP), Scottish Road quasi-brittle materials. Pure and Applied Geophysics, 137,
Research Board (SRRB) from Transport Scotland and 461–487.
the EPSRC Doctoral Training Award. Kell, G. S. 1975. Density, thermal expansivity, and com-
pressibility of liquid water from 0. deg. to 150. deg..
Correlations and tables for atmospheric pressure and satu-
REFERENCES ration reviewed and expressed on 1968 temperature scale.
Journal of Chemical and Engineering Data, 20, 97–105.
American Association of State Highway and Transporta- Laloui, L., Nuth, M. & Vulliet, L. 2006. Experimental
tion Officials 2007. Provisional Test Method for the and numerical investigations of the behaviour of a heat
Coefficient of Thermal Expansion of Hydraulic Cement exchanger pile. International Journal for Numerical and
Concrete. Washington DC, AASHTO. Analytical Methods in Geomechanics, 30, 763–781.

43
Lee, H.-Y., Lee, S.-H., Kim, J.-B. & Lee, J.-H. 2007. Creep– pile performance in saturated sand. Canadian Geotechni-
fatigue damage for a structure with dissimilar metal welds cal Journal, 52, 1045–1057.
of modified 9Cr–1Mo steel and 316L stainless steel. Rashid, M. & Mansur, M. 2005. Reinforced high-strength
International Journal of Fatigue, 29, 1868–1879. concrete beams in flexure. ACI Structural Journal, 102,
Li, L.Y. & Purkiss, J. 2005. Stress–strain constitutive equa- 462–471.
tions of concrete material at elevated temperatures. Fire Rotta Loria, A. F., Gunawan, A., Shi, C., Laloui, L., &
Safety Journal, 40(7), 669–686. Ng, C. W. W. 2015. Numerical modelling of energy piles
Litle, A. W. & Paparoni, M. 1966. Size Effect in Small- in saturated sand subjected to thermo-mechanical loads.
Scale Models of Reinforced Concrete Beams. Journal Geomechanics for Energy and the Environment, 1, 1–15.
Proceedings, 63. Song, K.-M., Mitchell, J. & Gladden, L. F. 2009. Observ-
Loli, M., Knappett, J. A., Brown, M. J., Anastasopou- ing microstructural evolution during plaster hydration.
los, I. & Gazetas, G. 2014. Centrifuge modeling of Diffusion Fundamentals, 10, 22.1–22.3.
rocking-isolated inelastic RC bridge piers. Earthquake Stewart, M. A. & McCartney, J. S. 2014. Centrifuge Modeling
Engineering & Structural Dynamics, 43, 2341–2359. of Soil-Structure Interaction in Energy Foundations. Jour-
Loveridge, F.A. & Powrie, W., 2013, April. Pile heat exchang- nal of Geotechnical and Geoenvironmental Engineering,
ers: thermal behaviour and interactions. Proceedings of ASCE, 140.
the Institute of Civil Engineers: Geotechnical Engineering Vasiliev, L. 2008. Micro and miniature heat pipes–Electronic
(Vol. 166, No. 2, pp. 178–196). component coolers. Applied Thermal Engineering, 28,
Murphy, K. D. & McCartney, J. S. 2015. Seasonal Response of 266–273.
Energy Foundations During Building Operation. Geotech- Yavari, N., Tang, A. M., Pereira, J.-M. & Hassen, G. 2014.
nical and Geological Engineering, 33, 343–356. A simple method for numerical modelling of energy
Ng, C. W. W., Shi C., Gunawan, A., Laloui, L. & Liu, H. L. pile’s mechanical behaviour. Geotechnique Letters, 4,
2015. Centrifuge modelling of heating effects on energy 119–124.

44

You might also like