You are on page 1of 10

Small-Scale Modeling of Thermomechanical Behavior of

Reinforced Concrete Energy Piles in Soil


Rui Zhao 1; Anthony Kwan Leung 2; Davide Vitali 3; Jonathan Adam Knappett 4; and Zheng Zhou 5
Downloaded from ascelibrary.org by Ahsanullah University of Science and Technology on 12/28/21. Copyright ASCE. For personal use only; all rights reserved.

Abstract: Small-scale physical model tests have been increasingly used to study the thermomechanical soil–pile interaction, but existing
model piles are highly simplified and do not have representative thermal properties or the quasi-brittle mechanical behavior of RC. This study
aims to overcome these shortcomings by presenting a new type of model RC. This consists of a mortar (plaster, sand, and water) with copper
powder added to tune the mixture’s thermal properties, along with a steel reinforcing cage. Fine sand was used to represent geometrical
scaling of the prototype aggregates to correctly capture the quasi-brittle structural response. Adding copper powder content of 6% (by
volume) matched the coefficient of thermal expansion and thermal conductivity of prototype concrete without changing the axial and flexural
properties of model piles. In 1g soil–structure interaction tests, the model pile was able to serve as an effective heat exchanger for transferring
heat from a water-carrying pipe embedded within the mortar to the surrounding soil. The model RC exhibited cyclic pile head settlement due
to repeated pile heating and cooling. DOI: 10.1061/(ASCE)GT.1943-5606.0002225. © 2020 American Society of Civil Engineers.

Introduction 2015; Chen and McCartney 2017; Zhao et al. 2017), and, more
recently, centrifuge modeling (McCartney and Rosenberg 2011; Ng
Energy piles have been increasingly used in urban spaces as heat et al. 2015; Goode and McCartney 2015; Stewart and McCartney
exchangers for not only providing structural support to the super- 2014). Centrifuge modeling enables small-scale physical models to
structure above but also for exploiting the renewable near-surface be tested within an elevated gravity field, such that soil stress levels
geothermal energy for heating, ventilation, and air conditioning are identical to those experienced by much larger prototypes at
purposes via a ground-source heat pump (Brandl 2006). Energy homologous points and under much more controlled test conditions
piles are typically made of RC, with pipe loops attached to the than are possible in field tests. A challenge to study the soil–energy
reinforcement cage for circulating heat-carrying fluid to facilitate pile interaction meaningfully using a centrifuge is selecting an
soil–concrete–pipe heat transfer. The alternate summer and winter appropriate model of the pile that can simultaneously scale both
operation of an energy pile continually heats and cools the pile and the thermal and mechanical properties of the structure at prototype
the surrounding soil. Cyclic pile heating and cooling mobilizes soil scale. Aluminum alloy and concrete are the two common choices
stress and strain (Ng et al. 2015; Stewart and McCartney 2014; of model material. For instance, Ng et al. (2015) and Yavari
Rotta Loria and Laloui 2017; Zhao et al. 2017; Elzeiny et al. et al. (2014) modeled their piles using an elastic aluminum alloy.
2018; Xiao et al. 2018), consequently affecting the pile capacity However, only either stiffness or strength can be scaled, but not
and settlement. Understanding the behavior of thermoactive pile both, because of the differences in intrinsic properties between
requires improved understanding of the load transfer mechanisms aluminum and RC (Knappett et al. 2011). Another limitation of
due to cyclic thermal loads and thermomechanical soil responses. using this approach is its inability to capture the thermal pile behav-
Thermomechanical soil–pile interaction has been studied by dif- ior. The coefficient of thermal expansion (CTE) of aluminum
ferent methods, namely, field tests (e.g., Mimouni and Laloui 2015; (22.2 × 10−6 με=°C) is up to three times higher than that of RC
Bourne-Webb et al. 2009; Murphy and McCartney 2015), numeri- (6 to 9 × 10−6 με=°C) (Bourne-Webb et al. 2009), while the ther-
cal modeling (e.g., Rotta Loria and Laloui 2017; Rotta Loria et al. mal conductivity of aluminum (205 W=m · K) is 100–200 times
higher than that of RC (∼1 W=m · K) (GSHP 2012). This mismatch
1
Research Student, Discipline of Civil Engineering, School of Science of model and prototype thermal properties means that aluminum
and Engineering, Univ. of Dundee, Dundee DD1 4HN, UK. alloy is unable to study the efficiency of pipe–concrete–soil heat
2
Assistant Professor, Dept. of Civil and Environmental Engineering, transfer properly.
Hong Kong Univ. of Science and Technology, Clear Water Bay, Kowloon, Reinforced concrete has been used (e.g., Stewart and McCartney
Hong Kong (corresponding author). ORCID: https://orcid.org/0000-0002
2014) as an option to model the thermomechanical behavior of
-5192-5033. Email: ceanthony@ust.hk
3
Research Student, Discipline of Civil Engineering, School of Science energy piles. Although the thermal properties in this case are more
and Engineering, Univ. of Dundee, Fulton Bldg., Dundee DD1 4HN, UK. realistically modeled, the particle size of coarse aggregates in RC
ORCID: https://orcid.org/0000-0003-3756-1568 is not scaled geometrically. This means that the particle size
4
Professor, Discipline of Civil Engineering, School of Science and distribution is the same between the model and prototype scale,
Engineering, Univ. of Dundee, Fulton Bldg., Dundee DD1 4HN, UK. which would potentially result in overstrength of the model material
5
Research Student, Dept. of Civil and Environmental Engineering, (e.g., Belgin and Sener 2008). Proper scaling of the particle size in
Hong Kong Univ. of Science and Technology, Clear Water Bay, Kowloon, an RC model is crucial for correctly capturing the failure mechanism
Hong Kong.
and the cracking process of this kind of quasi-brittle and nonlinear
Note. This manuscript was submitted on August 17, 2018; approved on
October 29, 2019; published online on February 13, 2020. Discussion per- material (Bažant 2005). To overcome this shortcoming, attempts
iod open until July 13, 2020; separate discussions must be submitted for have been made to use plaster-based mortars to capture the quasi-
individual papers. This paper is part of the Journal of Geotechnical and brittleness of RC. Knappett et al. (2011) created various types of
Geoenvironmental Engineering, © ASCE, ISSN 1090-0241. mortar in which fine silica sand is used to geometrically scale

© ASCE 04020011-1 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2020, 146(4): 04020011


the coarse aggregate in RC. This type of model concrete mimicked content. The mean strength of 6% copper powder normalized by that
the nonlinear quasi-brittle behavior of concrete and has representa- of the nonthermal model concrete (0% copper) is 1.00  0.37. One-
tive mechanical strengths (i.e., unconfined compressive strength way ANOVA showed that the normalized strengths between the 0%
from 26.3 to 81.7 MPa and modulus of rupture from 2.02 to and 6% cases are statistically not different (p-value ¼ 0.592 > 0.05).
4.42 MPa). These types of model concrete have been successfully This means that the addition of copper powder did not adversely
used for modeling various nonenergy structural elements in centri- affect the compressive strength of the new model concrete.
fuge tests, including pile-reinforced slopes (Al-Defae and Knappett Effects of copper powder on the modulus of rupture of the new
2014) and RC bridge piers on rocking shallow foundations (Loli model concrete were also determined via four-point bending (FPB)
et al. 2014). tests, following BS EN 12390-5 (BSI 2009b). Six replications of
This study aims to build on this previous work, developing a new beam samples (25 × 25 × 250 mm) were tested for each copper
Downloaded from ascelibrary.org by Ahsanullah University of Science and Technology on 12/28/21. Copyright ASCE. For personal use only; all rights reserved.

model RC that is thermally enhanced to more realistically scale the powder content. The moduli of rupture for 6% and 12% copper
thermomechanical properties of an RC energy pile representative to powder, normalized by the value obtained at 0%, were 1.06  0.02
those in the prototype. A series of structural tests was undertaken to and 1.08  0.14, respectively. As previously, ANOVA suggests a
characterize the thermal, mechanical, and coupled thermomechanical p-value of 0.143 (>0.05), meaning that there is no statistically
properties of the model RC. To illustrate the capabilities of the new significant difference between these tests. This suggests that the
model RC, it was used to construct model energy piles for evaluating addition of copper power did not substantially affect the modulus
their axial thermomechanical interaction in soil at 1g. of rupture.

New Model Concrete


Design of Model RC Energy Piles

Constituents of the New Model Concrete Having assessed the fundamental thermal and mechanical proper-
ties of the model concrete, it was subsequently used to produce
The new model concrete is based on the plaster-based mortars
small-scale model RC energy piles, designed for future centrifuge
pioneered by Knappett et al. (2011). Their model concrete composes
model tests following the high gravitational acceleration (high-g)
of either α-form dental stone or β-form surgical plaster, water, and
scaling laws summarized in Table 1. Two types of piles were con-
silica sand (Congleton HST95, Bent Farm, Congleton, Cheshir). To
sidered, one for testing axial behavior and the other for flexural
enhance the thermal properties of their mix for modeling the behav-
behavior.
ior of energy piles, copper powder (Phoenix Scientific Industries,
The first pile type is a 1/35th model energy pile (Type 1) de-
East Sussex, UK) was incorporated into the mix. Copper powder
signed as a part of a tall building foundation, which means that the
has a much higher thermal conductivity (400 W=m · K) than plaster
pile was expected to be subject to both axial and lateral loading at
(0.29 W=m · K) and quartz (i.e., 6.5 W=m · K). The diameter of the
the pile head. Considering the physical constraints of the model
copper powder used ranges from 0.002 to 0.1 mm, and D10 (the par-
size, a pile geometry of 0.63 × 0.63 × 10.5 m (in prototype) was
ticle diameter at 10% passing), D30 , and D60 are 0.013, 0.026, and
chosen. Following Broms (1964), this pile geometry would neces-
0.042 mm, respectively. The Mix 2 design mix from Knappett et al.
sitate a design moment capacity of approximately 200 kNm in sandy
(2011) was adopted here. This utilizes β-form surgical plaster at a
soil (friction angle of 32°), which matches the range of typical values
water/plaster ratio of 0.9∶1 and sand/plaster ratio of 1∶1, to which
found in the literature (e.g., Al-Defae and Knappett 2014; Hayward
various amounts of copper powder were added. The mix was left
et al. 2000). Following Eurocode 2 [EN 1992-1-1(CEN 2004a)],
to air cure at room temperature (20°C–24°C) for 28 days prior to
four longitudinal steel bars of 35-mm diameter each and 10 shear
testing.
links of 22-mm diameter each with 1.05-m spacing are needed

Thermal and Mechanical Behavior


A series of tests were conducted to characterize the thermal proper-
Table 1. Summary of relevant scaling factors for thermomechanical soil–
ties of the new model concrete and to study whether the addition
structure interaction problems in 1g and high gravitational acceleration
of copper power might affect the fundamental mechanical proper- (high-g)
ties. Thermal conductivity tests were performed on slab samples
(45 × 150 × 150 mm) by using a hot box apparatus. Three copper Quantity 1g scale factor High-g scale factor
powder contents (0%, 6%, and 12% by volume) were considered. General
Three replications were tested for each mix. Following BS 874-3.2 Length N N
(BSI 1990), each specimen was subjected to a controlled heat gra- Density 1 1
dient across its thickness and heat flux was measured. By Fourier’s Stress N 1
law, the thermal conductivity was determined. As expected, adding Strain (mechanical) N 0.5 1
a more thermally conductive material (i.e., copper powder) into the Stiffness N 0.5 1
Structural properties
mix increased the thermal conductivity of the model concrete from
Bending stiffness, EI N 4.5 N4
0.38  0.01 to 0.73  0.08 and then to 0.94  0.04 W=ðm KÞ Axial stiffness, EA N 2.5 N2
(mean  standard error of mean) as the copper powder content Axial force N3 N2
increased from 0% to 6% to 12%, respectively. Comparing with Axial displacement N 1.5 N
the values for concrete typically used to construct energy piles Bending moment N4 N3
[1–4 W=ðm KÞ] (GSHP 2012), copper powder contents ranging Temperature related
from 6% to 12% are deemed appropriate for capturing the thermal Strain (thermal) N 0.5 1
properties of energy piles. Time (diffusion) N2 N2
Unconfined compression tests [BS EN 12390-3 (BSI 2009a)] Temperature change N 0.5 1
were conducted. Cubic samples (100 mm) were tested for compres- Source: Data from Iai et al. (2005).
sive strength (fc;100 ) with three replications for each copper powder Note: N = prototype divided by model.

© ASCE 04020011-2 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2020, 146(4): 04020011


Thermomechanical Behavior of Model Energy Pile

Free Thermal Expansion


The first series of tests were free thermal expansion tests, aiming
to determine the CTE of the model pile Type 1. Three pairs of strain
gauges (steel foil; Model SGD-3/120-LY11, Omega Engineering,
Norwalk, Connecticut) and thermocouples were attached to the pile
surface at 105, 170, and 235 mm along the pile length to measure
axial strain mobilization and pile surface temperature, respectively.
Downloaded from ascelibrary.org by Ahsanullah University of Science and Technology on 12/28/21. Copyright ASCE. For personal use only; all rights reserved.

To minimize temperature effects, half-bridge connections were ap-


plied to each pair of strain gauges (National Instruments 2016).
One gauge was mounted along the principal axis of strain (i.e., axial
direction), while the other was mounted perpendicularly. A layer of
Fig. 1. Cross section and reinforcement layouts of the model RC epoxy resin with a thickness of about 1 mm was applied on the
energy piles of Type 1 and Type 2 (all dimensions expressed at model surface of each strain gauge for protection. The epoxy–sand inter-
scale). face has a friction coefficient of 0.53 (Ramadan et al. 2013), which
is well within the range for the concrete–sand interface reported by
Di Donna et al. (2016) (0.49–0.58). In fact, along the 300-mm-long
model pile, the three stain gauges occupy a length of only 3 mm.
(Fig. 1). The longitudinal reinforcement ratio is 1.0%. Both scaled
Effects of the presence of external strain gauges on the soil–pile
longitudinal bars and shear links were made of stainless steel wire
interface are thus negligible. Each model pile was supported by
(Grade 316, Ormiston Wire, Middlesex, London), which has a CTE
two stainless steel rollers, so the pile was allowed to expand and
of 18.5 με=°C. Stainless steel with a yield strength of 460 MPa and
contract freely upon heating–cooling cycles. Four heating–cooling
a model diameter of 1 mm was used for producing the longitu-
cycles were introduced to the model pile by varying the temperature
dinal bars, while that with a yield strength of 380 MPa and a model
of water between ambient (∼22°C) and 50°C (i.e., heating phase).
diameter of 0.6 mm was adopted for producing the shear links. All
Each pile was heated until the average pile surface temperature
reinforcement was coated with fine sand using epoxy for providing
reached a thermal equilibrium, and it was then cooled to the
mechanical interlocking with the mortar [after Knappett et al.
ambient temperature. Each cycle took 40–50 min. Inlet and outlet
(2011)].
temperature (T in and T out ) were measured by thermocouples. These
To allow heat energy to be circulated in and out of the model
measurements were made at the silicone pipes where they enter the
energy piles, two flexible silicone pipes with a model internal diam-
top of the model pile. Subsequently, another four cooling–heating
eter of 1.5 mm each were used. The silicone pipe has a tensile
cycles (i.e., cooling phase) were applied to the same model piles but
strength of 10 MPa and the CTE ranges from 70 to 110 με=°C.
varying the temperature of water between ambient (∼22°C) and 5°C.
Each pipe was attached to shear links between the longitudinal
During testing, the ambient temperature also varied and this could
bars with a few spots of rapid-set glue forming a U-shape loop ar-
introduce additional thermal expansion and contraction within the
rangement as shown in Fig. 1, as commonly adopted in the field
strain gauges. To compensate for this effect, a control experiment
(e.g., Loveridge and Powrie 2013; Mimouni and Laloui 2015). The
was conducted in parallel, consisting of an identical strain gauged
longitudinal bars, the shear links, and the two U-shaped silicone
model pile placed next to the tested energy pile, but which was not
piles were assembled in a multipart formwork, into which the mor-
subjected to heating–cooling cycles. Any thermal strain obtained
tar mix with 6% copper powder was poured and subsequently air
from this control pile was subtracted from the values obtained from
cured for 28 days.
the tested pile.
The second pile type is a 1/24th model pile designed as part of
As can be seen in Fig. 2(a), during all four heating cycles, the
a discretely spaced pile row for slope stabilization, and thus it
loop inlet water temperature (T in ) was consistently higher than the
would be subjected to predominately lateral loading. According to
outlet’s (T out ) by approximately the same temperature difference of
the test results of nonenergy pile reinforcement systems reported by
3°C. The temperature difference between T in and T out was likely
Al-Defae and Knappett (2014) and Jeong et al. (2003), a mean pile
because a portion of the heat energy of the water that was circulat-
moment capacity of 230 kNm was used as a design value. Follow-
ing within the model pile was transferred to the model concrete and
ing Eurocode 2 (EN 1992-1-1, 2004), a doubly reinforced 0.6 ×
model reinforcement due to the temperature gradient between the
0.6 m square pile with a length (L) of 6 m and a reinforcement
pipes and the surrounding materials. During the cooling cycles
ratio of 2.1% is required. The model diameter of the longitudinal
[Fig. 2(b)], there was little difference between T in and T out under
bars and shear links used in this case were 1.25 and 0.6 mm, re-
the smaller temperature difference (ΔT ¼ −17°C) compared to the
spectively. The preparation procedures were otherwise identical to
heating cycles (ΔT ¼ þ28°C). The pile surface temperature during
those of pile Type 1, and utilized the same materials for the model
all heating and cooling cycles was respectively lower and higher
concrete (i.e., 6% copper) and steel wire.
than the water temperature due to thermal resistances of the differ-
In order to supply different temperatures continually to the
ent components of the mortar. Considering that the heat flow in the
model energy piles during testing, the silicone pipes embedded
system obeys Fourier’s law, the pile surface temperature (T p ) can
in each model pile were connected to a heating–cooling system
be estimated by
(Model F12-ED, Julabo, Seelbach, Germany). The system has a
water bath, where any temperature between 1°C and 99°C can X
be controlled. The system also has a pump, which is able to produce Tp ¼ Tf þ q Ri ð1Þ
a water flow rate up to 13.5 mL=s. This is sufficient to generate a i

turbulent flow regime within the embedded pipeline, hence maxi-


mizing the convective heat transfer mechanism from the pipes to where T f = water temperature, which is the mean of T in and
the model concrete (Loveridge and Powrie 2013). T out (Loveridge and Powrie 2013); Ri = thermal resistance of

© ASCE 04020011-3 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2020, 146(4): 04020011


Downloaded from ascelibrary.org by Ahsanullah University of Science and Technology on 12/28/21. Copyright ASCE. For personal use only; all rights reserved.

(a) (b)

Fig. 2. Time histories of water temperature (inlet and outlet) and pile surface temperature: (a) during four heating cycles; and (b) during four cooling
cycles, for the free thermal expansion tests.

component i, which in this case includes the silicone pipes, copper was almost uniform. It is thus a reasonable assumption that the tem-
powder, sand, and plaster; and q = heat transferred per unit length perature measured at the pile surface at each elevation, and hence
thermal strain, was the same as that along the cross section. At any
mSc ðT out − T in Þ pile depth, a linear and reversible response is observed. This sug-
q¼ ð2Þ
Lp gests that for the temperature range examined, the model RC en-
ergy pile made of the thermally enhanced mortar behaved thermally
where m = applied mass flow rate of water (1.6 × 10−3 kg=s); Sc = elastically. By estimating the gradient of the linear curves, the CTE
specific heat of water (4,187 J=kg K); and Lp = pipe length of the model pile is 8.9  1.4 με=°C, which is within the range of
(600 mm). To capture the effects of (1) pipe–mortar thermal resis- the linear thermal expansion of concrete (8.5–10 με=°C) (Bourne-
tance and (2) pipe-to-pipe heat exchange within the model pile Webb et al. 2009; Pasten and Santamarina 2014).
when estimating Ri , the calculation method proposed by Marcotte
and Pasquier (2008) was adopted, whereby a shape factor of 11 Compression Tests
was applied to take the U-loop pipe arrangement into account.
Hence, the equivalent thermal resistance of the model RC pile was The second series of tests conducted were uniaxial compression tests
0.25 mK=W (prototype), which is close to the typical range from for model pile Type 1. The toe of each free-standing model pile was
0.05 to 0.23 mK=W reported in the literature (Bourne-Webb et al. clamped and structurally fixed at a platform. A full moment fixity
2009; Mimouni and Laloui 2015). By Eq. (1), time histories of cal- was thus achieved. The pile head was subjected to vertical compres-
culated T p are superimposed in Fig. 2. Close agreement was found sion using an Instron 5985 load frame (High Wycombe, Bucking-
with the measurements. This verifies the mechanism of heat con- hamshire) under a loading rate of 0.05 MPa=s. To investigate any
duction within the model pile, similar to that normally seen in full- effect of temperature on the pile axial behavior, three levels of tem-
scale RC cases. perature were examined, ambient (19°C), heating, and cooling. For
Fig. 3 relates the thermally induced axial strain of pile Type 1 the heating test, the model pile was heated continuously by main-
with the pile surface temperature change. A recent field study re- taining the T in at 51°C. When a thermal equilibrium was reached,
ported by Faizal et al. (2018) shows that the temperature distribu- an axial load test was conducted (while T in was maintained). The
tion over the cross section of a full-scale 0.6-m-diameter bored pile axial behavior of the model pile was also tested after subjecting
to continuous cooling by maintaining the T in at 9°C. Three test rep-
lications were conducted for each temperature level.
Fig. 4 shows typical pile load–displacement curves during the
compression at different temperatures. Regardless of the pile tem-
perature applied, the axial load increased almost linearly and
reached a distinct peak value between 3 and 3.6 MN (prototype)
at an axial strain of around 0.005. In foundation designs, the ulti-
mate structural capacity of pile would rarely be reached because the
soil underneath would have normally failed under a lower load.
After mobilization of the peak strength, there was a significant re-
duction of stress. For the application of a piled building foundation,
the compressive resistance (N max ) and axial rigidity (EA, where E
is Young’s modulus and A is gross cross-section area of the model
pile) are of most interest. One-way ANOVA was carried out to de-
termine any statistical difference at the 5% significance level of
the pile axial behavior at the three temperature levels considered.
The compressive strength has no statistically significant difference
between the 9°C and 19°C cases (p-value ¼ 0.227 > 0.05). This
means that the temperature reduction (from 19°C to 9°C; i.e., 10°C
difference) was not significant enough to introduce a noticeable dif-
Fig. 3. Relation between thermally induced axial strain and pile surface
ference in the pile capacity. However, the p-value between the 19°C
temperature change.
and 51°C cases is 0.025 < 0.05, meaning that they are statistically

© ASCE 04020011-4 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2020, 146(4): 04020011


Downloaded from ascelibrary.org by Ahsanullah University of Science and Technology on 12/28/21. Copyright ASCE. For personal use only; all rights reserved.

Fig. 5. Typical moment–curvature curves obtained from four-point


bending tests of model pile Type 2 at three different temperatures,
Fig. 4. Typical axial force–displacement curves obtained from free- all expressed in prototype scale following N ¼ 24 and using the
compression tests of model pile Type 1 at three different temperatures, high-g scaling factors in Table 1.
all expressed in prototype scale following N ¼ 35 and using the high-g
scaling factors in Table 1.

workmanship. Indeed, the model RC exhibited a similar coefficient


of variation (COV) of bending strength as the full-scale RC (<10%)
different. For EA, no significant statistical difference is found for (Knappett et al. 2018). As can be seen in the inset of the figure,
either the pair of 9°C and 19°C (p-value ¼ 0.136 > 0.05) or 19°C after tests, prominent bending cracks were recreated, realistically
and 51°C (p-value ¼ 0.588 > 0.05). It is well known that below replicating the important quasi-brittle behavior found in the proto-
100°C, the mechanical properties of prototype concrete and steel type RC.
reinforcement, including compressive strength and elastic modulus, Fig. 6 shows the effects of temperature and copper powder con-
are insensitive to temperature change (Takeuchi et al. 1993). For the tent on peak bending moment and flexural rigidity (EI, taken as the
case of model concrete, some temperature dependency on the com- initial linear portion of the moment–curvature curve between 30%
pressive strength was found, whereas the EA was much less sensi- and 50% of the peak moment). Without copper powder (0%), both
tive. The modeling, however, is deemed satisfactory because for moment capacity [Fig. 6(a)] and EI [Fig. 6(b)] apparently increase
foundation applications (1) it is mainly EA that affects pile head with an increase in pile temperature, but these trends are statisti-
settlement, and (2) it is the soil shear strength, not the pile compres- cally not significant (p-value is always greater than 0.05). For the
sive strength, that limits the axial capacity of a soil–pile system. 6% case, no specific trend can be found between pile temperature
and its flexural properties (p-value is always greater than 0.05).
FPB Tests This is consistent with the fact that the mechanical properties of
concrete and steel are less affected when the temperature applied
The third series of tests was conducted to characterize any effects of is less than 100°C (Takeuchi et al. 1993). With 12%, copper, how-
temperature on flexural behavior of the model pile Type 2 via FPB ever, evident drops of both bending moment and EI can be seen.
tests. Water of different temperatures [5°C, ambient (20°C), 35°C, Indeed, p-values in both cases are less than 0.05, meaning that the
and 50°C] was circulated to each model pile until a thermal equi- addition of 12% copper powder affected the flexural properties
librium was reached, as indicated by the three thermocouples at- statistically significantly.
tached on the pile surface in each test. Each model pile was then Based on the observed properties of the new model concrete and
loaded to give a constant moment over the central third of the beam the thermomechanical axial and flexural behavior of the model
to obtain a moment–curvature curve. A test was complete when the piles, with support of statistical testing, it appears that an optimum
pile midsection exhibited large deformation and formed prominent copper powder content that could enhance the thermal properties of
cracking (at tensile strains of 0.016%–0.048%). Model piles mixed the model concrete by Knappett et al. (2011) but without adversely
with 0%, 6%, and 12% copper powder were examined. Three rep- affecting the mechanical response of the model structure is 6%. The
licates were tested for each amount of copper powder and each theoretical peak bending moment and EI predicted by KSU_RC
temperature level. were 325 kNm and 41 MNm2 , respectively. These values are
Fig. 5 compares typical moment–curvature relationships of the deemed acceptably close to the laboratory-measured average peak
model piles tested at different temperatures. A curve predicted moment (336 kNm) and average EI (33 MNm2 ).
by an open-source section-analysis software, KSU_RC (Esmaely
2013), at the ambient temperature of 20°C, is also shown for refer-
ence. The nonlinear stress–strain behavior of RC used in KSU_RC Thermomechanical Soil–Pile Interaction at 1g
was that proposed by Mander et al. (1988) and the behavior of steel
reinforcement is linear elastic–perfectly plastic. All model piles, To illustrate the effectiveness of using the new mode RC to model
regardless of the level of temperature tested, show an abrupt drop the axial thermomechanical behavior of energy piles in soil, 1g
of moment near the peak value as major cracks formed. Ductile physical model tests were performed. These tests are illustrative
behavior is observed because all model piles do not show any sign and they aim to explore whether key thermomechanical pile–soil
of ultimate collapse after a large curvature higher than 0.05 m−1 . interaction phenomena could be reproduced qualitatively, ahead
Some of the observed differences of responses were associated of subsequent centrifuge tests (not reported here). Although rel-
with experimental uncertainties such as material variabilities and evant scaling laws and nondimensional groups were utilized to

© ASCE 04020011-5 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2020, 146(4): 04020011


Downloaded from ascelibrary.org by Ahsanullah University of Science and Technology on 12/28/21. Copyright ASCE. For personal use only; all rights reserved.

Fig. 6. Temperature effects on (a) peak bending moment; and (b) flexural rigidity of model pile Type 2, all expressed in prototype scale following
N ¼ 24 and using the high-g scaling factors in Table 1.

ensure similitude in behavior, there is no intention to quantitatively base, where N q is the bearing capacity coefficient, which is a func-
compare prototype responses with field observations in this simple tion of friction angle (ϕ); σv0 is vertical effective stress; and A is pile
1g experiment. sectional area, and (2) βσv0 BL along the shaft, where β ¼ K tan δ;
K is earth pressure coefficient; δ is soil–pile interface friction angle;
and B is pile circumference]
Scaling Laws and Similitude
   
Considering that the sand used in the 1g tests has the same density Qpile Qpile
¼ ð4aÞ
as that of the prototype and the shear modulus (G) of such a soil is N q ðϕÞσv0 A m N q ðϕÞσv0 A p
proportional to the square root of effective confining pressure, the
so-called Type II scaling laws proposed by Iai et al. (2005) may be    
used. The scaling laws for soil and pile properties are summarized Qpile Qpile
¼ ð4bÞ
in Table 1. 4βðϕÞσv0 L m βðϕÞσv0 BL p
In order to achieve similitude, the relative soil–pile stiffness and
strength must be scaled appropriately. Nondimensional relative These relationships are satisfied if the friction angle is the same
axial stiffness can be described by the ratio of the elastic stiffness in the model and prototype. This is ensured in the centrifuge (high-
of the soil to the axial pile structural stiffness (Randolph and g) because the effective stress levels are representative of prototype
Wroth 1978) conditions. At 1g, the operative friction angle would be higher in
 2   2  the model than the prototype due to increased dilation at low con-
LG L G
¼ ð3Þ fining effective stress. This would suggest that structural failure
EA m EA p would be biased in the model; however, in the tests presented here,
the capacity was observed to be controlled by soil failure.
where the subscripts m and p = model and prototype scale, respec- To scale the heat flow in soil, the nondimensional number,
tively. This similitude is satisfied using both the 1g and high-g scal- Fourier’s number (which is derived from the second law of heat
ing laws from Table 2. Nondimensional relative axial strength can diffusion), may be used
be described by the ratio of the structural pile strength (Qpile ) to the
bearing capacity of the pile [proportional to (1) N q σv0 A at the pile    
at at
¼ ð5Þ
L2 m L2 p

Table 2. Properties of soil used in the physical model tests where a = soil thermal diffusivity; and t = heat diffusion time.
Because (1) the same sand was considered in the model and proto-
Parameter HST95 silica sand
type scale (i.e., scaling factor of a is 1) and (2) the scaling factor
State independent of L is N (Table 2), for similitude, the scaling factor for t must be
Specific gravity, Gs 2.63 scaled by N 2.
Maximum void ratio, emax 0.769 According to Table 1, mechanical strain is scaled by N 0.5. To
Minimum void ratio, emin 0.467
consider thermomechanical coupling, the scale factor for mechani-
Critical friction angle, ϕcri (degrees) 32
Coefficient of thermal expansion (με=°C) 0.55–0.75
cal strain has to be the same as that for thermal strain (εT ), so it
State dependent should also be scaled by N 0.5. Considering thermal elasticity
Peak friction angle, ϕp (degrees) 44
Dilation angle, φ (degrees) 14.5 εT ¼ αT ΔT ð6Þ
Young’s modulus, E (MPa) 38.7
Unit weight, γ (kN=m3 ) 16.6 where αT = coefficient of thermal expansion; and ΔT =
Thermal conductivity [W=ðm KÞ] 0.54 temperature change. Again, because (1) εT is nondimensional,
Specific heat [J=ðkg KÞ] 830 (2) the same sand is considered in the model and prototype scale
Note: The physical properties of the HST95 silica sand are from Al-Defae (i.e., scaling factor of αT is 1), and (3) εT is scaled by N 0.5, ΔT must
and Knappett (2014). be scaled by N 0.5.

© ASCE 04020011-6 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2020, 146(4): 04020011


Test Setup and Procedures influence zone of the soil temperature induced by pile heating and
cooling is typically ∼4D. Hence, the pile heating and cooling proc-
In this test, a 1/35th scale model energy pile Type 1 (6% copper
esses introduced by the model pile in this study are not expected to
powder) was installed wish-in-placed in the center of a uniform
be subjected to any thermal boundary effect. The distance between
bed of Congleton HST95 silica sand in a model container (Fig. 7).
the pile toe and the bottom of the sand bed (8D) was sufficient to
As the model pile was suspended in position, the sand was air plu-
avoid boundary effects on the pile base capacity (Knappett et al.
viated surrounding the pile to produce a uniform sand deposit with
2016). From this point on, all dimensions are expressed in prototype
a height of 350 mm at an average relative density Dr of 74%. The
scale.
mechanical properties of the sand at this specific Dr are summa-
The test consisted of four phases, representing different propor-
rized in Table 2. The model pile was instrumented with a vertical tions of the pile capacity (i.e., working loads for different factors of
Downloaded from ascelibrary.org by Ahsanullah University of Science and Technology on 12/28/21. Copyright ASCE. For personal use only; all rights reserved.

array of three thermocouples and strain gauges for measuring pile safety). In each phase, (1) the maintained vertical working load was
surface temperature and any thermally induced pile axial strain at 5-, initially applied using the Instron 5985 load frame under load con-
70-, and 135-mm depth. The pile embedment depth was 200 mm, trol at a rate of 0.05 MPa=s, followed by (2) four heating–cooling
leaving the top 100 mm above the soil surface. The model box was cycles, and (3) mechanical unloading to zero head load. Time his-
hydraulically insulated, but not thermally. The model pile was in- tories of the mechanical and thermal loads applied are given in
stalled at 11 times the diameter (D, i.e., 11D) from the side boun- Fig. 8. The working loads applied in the first, second, third, and
dary and 8D from the bottom. Based on the limited field data fourth phases represented 0%, 13%, 51%, and 89% of the ultimate
available in the literature (e.g., Mimouni and Laloui 2015), the radial pile capacity (1,741 kN; much less than the ultimate structural
capacity in Fig. 4), respectively. The ultimate capacity was deter-
mined by the sum of the shaft and base resistances, following
Terzaghi’s bearing capacity equation and the calculation proce-
dures outlined in Eurocode 7 [EN 1997-1 (CEN 2004b)]. For the
cohesionless sand with a friction angle of 32°, the bearing capacity
factors associated with overburden pressure (N q ) and cohesion (N c )
are 23 and 0 (Knappett and Craig 2019), respectively. During
thermal cycling, water with T in varying between 12°C and 45°C
(ambient temperature of 22°C) was circulated within the model
piles [Fig. 8(b)]. Identical heating–cooling cycles were applied
in all four phases. Pile head displacement, pile surface temperature,
pile axial load mobilization, and soil temperature were monitored
continuously throughout the test.

Test Results
Fig. 9 shows the thermally induced pile head displacement. In gen-
eral, the pile head heaved upon heating and settled upon cooling, as
expected. When the pile head was not constrained (i.e., 0% work-
ing load), the pile head displacement was almost fully reversible.
When the pile head was more mechanically constrained (i.e., higher
applied working load), irreversible ratcheting pile head settlement
was increasingly significant. The same behavior was identified
qualitatively in both the laboratory 1g test (Yavari et al. 2014) and
centrifuge testing (Ng et al. 2015) using aluminum alloy pile mod-
els and from numerical simulations (Pasten and Santamarina 2014;
Zhao et al. 2017). The model pile, with the use of the new model
concrete, can reproduce the soil–structure interaction behavior of
RC energy piles.
The time histories of the thermomechanical strains during all
four phases are presented in Fig. 8(c). Based on the coefficient
of thermal expansion of the model pile determined from the free
thermal expansion tests (Fig. 3) and the known time history of tem-
perature in this test, theoretical free thermal strain was determined.
The corresponding time history is superimposed in Fig. 8(c) for
reference. As expected, the amplitude of strain fluctuation at all
three depths in any loading stage cycle was less than that of the
free thermal strain due to the constraint of the surrounding soil
where soil–pile interface friction exists. This explains why, for a
given applied working load, the strain mobilized near the ground
surface was greater than at depth. Apparently, under any working
load, the pile strains mobilized at all depths were almost reversible
during the four heating–cooling cycles. This seems to imply that
the soil–pile interaction may be largely thermoelastic. Indeed, upon
Fig. 7. Maintained axial load thermal cyclic tests for model pile Type 1
heating or cooling of silica sand, only very limited thermoplastic
at 1g. All dimensions are in millimeters in model scale.
response has previously been observed (Ng et al. 2015). It has been

© ASCE 04020011-7 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2020, 146(4): 04020011


Downloaded from ascelibrary.org by Ahsanullah University of Science and Technology on 12/28/21. Copyright ASCE. For personal use only; all rights reserved.

Fig. 8. Time histories of (a) applied vertical load; (b) pile surface temperature; and (c) thermally induced axial strain of the model pile Type 1 when
subjected to cyclic pile thermal loading under maintained working loads (WLs), all expressed in prototype scale following N ¼ 35 and using the 1g
scaling factors in Table 1.

recently advocated that thermoelasticity may be sufficient to cap- soil beneath the pile toe, as consistently found in the numerical pre-
ture the behavior of RC energy piles installed in sand (Rotta Loria diction made by Zhao et al. (2017).
and Laloui 2017; Zhao et al. 2017). The strain results suggest that
the plastic pile head settlement observed under high working load
in Fig. 9 was associated with the continual plastic straining of the Summary and Conclusions
This study presented a new type of model concrete that is suitable
for realistically modeling the thermomechanical behavior of energy
piles and their interaction with the surrounding soil. The new model
concrete is a thermally enhanced mortar. It is a mixture of plaster,
water, and sand to geometrically scale the aggregate and copper
powder to improve the thermal properties of the mixture. It was
revealed that 6% of copper powder (by volume) is an optimal value
for the new model concrete, matching the coefficient of thermal
expansion and thermal conductivity of field concrete, while not
detrimentally affecting the mechanical properties.
The model concrete was used to construct two small-scale RC
energy piles of different dimensions and scales to examine their
axial and lateral thermomechanical behavior in soils. Besides model
reinforcement, silicone pipes arranged in U-shaped loops were em-
bedded in each pile to allow water of different temperatures to be
circulated through the pile. Controlled pile heating–cooling could
hence be applied to the surrounding soil, allowing pipe–concrete–
soil heat transfer efficiency to be evaluated. This is a unique feature
Fig. 9. Pile head displacement when subjected to cyclic pile thermal of the new model concrete that most existing small-scale piles mod-
loading under maintained working loads (WLs), all expressed in eled by, for example, aluminum alloy, cannot achieve. Given the
prototype scale following N ¼ 35 and using the 1g scaling factors
range of pile temperature tested (5°C–50°C) and 6% of copper pow-
in Table 1.
der content added, key axial and flexural stiffnesses of the model

© ASCE 04020011-8 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2020, 146(4): 04020011


RC piles were not temperature sensitive, which is consistent with Esmaely, A. 2013. KSU_RC: Moment curvature, force deflection and in-
field RC. teraction analysis of RC members. Manhattan, KS: Civil Engineering
As revealed from the 1g tests, temperature-induced cyclic pile Dept., Kansas State Univ.
head settlement (ratcheting) for piles under vertical maintained Faizal, M., A. Bouazza, J. S. McCartney, and C. Haberfield. 2018. “Axial
load, as typically observed in field energy pile systems for build- and radial thermal responses of an energy pile under a 6-storey residen-
ings, was reproduced in the laboratory when the new type of model tial building.” Can. Geotech. J. 56 (999): 1–15. https://doi.org/10.1139
/cgj-2018-0246.
RC pile was heated and cooled consecutively in a uniform soil de-
Goode, J. C., III, and J. S. McCartney. 2015. “Centrifuge modeling of
posit. Consistent with the findings of most of the field testing and
end-restraint effects in energy foundations.” J. Geotech. Geoenviron.
finite-element modeling reported in the literature, the pile head Eng. 141 (8): 04015034. https://doi.org/10.1061/(ASCE)GT.1943-5606
ratcheting measured increased with an increase in the working load
Downloaded from ascelibrary.org by Ahsanullah University of Science and Technology on 12/28/21. Copyright ASCE. For personal use only; all rights reserved.

.0001333.
applied on the pile head. The range of thermally induced pile axial GSHP (Ground Source Heat Pump Association). 2012. Thermal pile
strain is reasonably close to the existing field data. design, installation and materials standards. Milton Keynes, UK:
GSHP.
Hayward, T., A. Lees, W. Powrie, D. J. Richards, and J. Smethurst. 2000.
Acknowledgments Centrifuge modelling of a cutting slope stabilised by discrete piles.
Rep. 471. Berkshire, UK: Transport Research Laboratory.
The first author (RZ) acknowledges the studentship provided by the Iai, S., T. Tobita, and T. Nakahara. 2005. “Generalised scaling relations for
Chinese Scholarship Council. The second (AL) and third (DV) dynamic centrifuge tests.” Géotechnique 55 (5): 355–362. https://doi
authors thank the research expenditures provided by Energy Tech- .org/10.1680/geot.2005.55.5.355.
nology Partnership (ETP), Scottish Road Research Board (SRRB) Jeong, S., B. Kim, J. Won, and J. Lee. 2003. “Uncoupled analysis of sta-
from Transport Scotland, the EPSRC Doctoral Training Award as bilizing piles in weathered slopes.” Comput. Geotech. 30 (8): 671–682.
well as the Scottish Funding Council (SFC). https://doi.org/10.1016/j.compgeo.2003.07.002.
Knappett, J. A., M. J. Brown, L. Shields, A. H. Al-Defae, and M. Loli.
2018. “Variability of small scale model reinforced concrete and impli-
cations for geotechnical centrifuge testing.” In Proc., 9th Int. Conf.
References on Physical Modelling in Geotechnics (ICPMG 2018), 241–246.
Boca Raton, FL: CRC Press.
Al-Defae, A. H., and J. A. Knappett. 2014. “Centrifuge modeling of the
Knappett, J. A., K. Caucis, and M. J. Brown. 2016. “CHD pile perfor-
seismic performance of pile-reinforced slopes.” J. Geotech. Geoen-
viron. Eng. 140 (6): 04014014. https://doi.org/10.1061/(ASCE)GT mance: Part II—Numerical modelling.” Proc. Inst. Civ. Eng. Geotech.
.1943-5606.0001105. Eng. 169 (5): 436–454. https://doi.org/10.1680/jgeen.15.00132.
Bažant, Z. P. 2005. Scaling of structural strength. Oxford, UK: Elsevier Knappett, J. A., and R. F. Craig. 2019. Craig’s soil mechanics. 9th ed.
Butterworth-Heinemann. London: Spon.
Belgin, C. M., and S. Sener. 2008. “Size effect on failure of over reinforced Knappett, J. A., C. Reid, S. Kinmond, and K. O’Reilly. 2011. “Small-scale
concrete beams.” Eng. Fract. Mech. 75 (8): 2308–2319. https://doi.org modeling of reinforced concrete structural elements for use in a geo-
/10.1016/j.engfracmech.2007.09.006. technical centrifuge.” J. Struct. Eng. 137 (11): 1263–1271. https://doi
Bourne-Webb, P. J., B. Amatya, K. Soga, T. Amis, C. Davidson, and P. .org/10.1061/(ASCE)ST.1943-541X.0000371.
Payne. 2009. “Energy pile test at Lambeth College, London: Geotech- Loli, M., J. A. Knappett, M. J. Brown, I. Anastasopoulos, and G. Gazetas.
nical and thermodynamic aspects of pile response to heat cycles.” Géo- 2014. “Centrifuge modeling of rocking-isolated inelastic RC bridge
technique 59 (3): 237–248. https://doi.org/10.1680/geot.2009.59.3.237. piers.” Earthquake Eng. Struct. Dyn. 43 (15): 2341–2359. https://doi
Brandl, H. 2006. “Energy foundations and other thermo-active ground .org/10.1002/eqe.2451.
structures.” Géotechnique 56 (2): 81–122. https://doi.org/10.1680/geot Loveridge, F., and W. Powrie. 2013. “Pile heat exchangers: Thermal behav-
.2006.56.2.81. iour and interactions.” Proc. Inst. Civ. Eng. Geotech. Eng. 166 (2):
Broms, B. B. 1964. “Lateral resistance of piles in cohesionless soils.” J. Soil 178–196. https://doi.org/10.1680/geng.11.00042.
Mech. Found. Div. 90 (3): 123–156. Mander, J. B., M. J. N. Priestley, and R. Park. 1988. “Theoretical stress-
BSI (British Standards Institution). 1990. Methods for determining thermal strain model for confined concrete.” J. Struct. Eng. 114 (8): 1804–1826.
insulating properties. Tests for thermal transmittance and conductance. https://doi.org/10.1061/(ASCE)0733-9445(1988)114:8(1804).
Calibrated hot-box method. BS 874-3.2. London: BSI. Marcotte, D., and P. Pasquier. 2008. “On the estimation of thermal resis-
BSI (British Standards Institution). 2009a. Testing hardened concrete. tance in borehole thermal conductivity test.” Renewable Energy 33 (11):
Compressive strength of test specimens. EN 12390-3. London: BSI. 2407–2415. https://doi.org/10.1016/j.renene.2008.01.021.
BSI (British Standards Institution). 2009b. Testing hardened concrete. McCartney, J. S., and J. E. Rosenberg. 2011. “Impact of heat exchange on
Flexural strength of test specimens. EN 12390-5. London: BSI. side shear in thermo-active foundations.” In Proc., Geo-Frontiers 2011:
CEN (European Committee for Standardization). 2004a. Design of
Advances in Geotechnical Engineering, 488–498. Reston, VA: ASCE.
concrete structures. Part 1-1: General rules and rules for buildings.
Mimouni, T., and L. Laloui. 2015. “Behaviour of a group of energy piles.”
Eurocode 2. Brussels, Belgium: CEN.
Can. Geotech. J. 52 (12): 1913–1929. https://doi.org/10.1139/cgj-2014
CEN (European Committee for Standardization). 2004b. Geotechnical
-0403.
design—Part 1: General rules. Eurocode 7. Brussels, Belgium: CEN.
Murphy, K. D., and J. S. McCartney. 2015. “Seasonal response of energy
Chen, D., and J. S. McCartney. 2017. “Parameters for load transfer analy-
sis of energy piles in uniform nonplastic soils.” Int. J. Geomech. foundations during building operation.” Geotech. Geol. Eng. 33 (2):
17 (7): 04016159. https://doi.org/10.1061/(ASCE)GM.1943-5622 343–356. https://doi.org/10.1007/s10706-014-9802-3.
.0000873. National Instruments. 2016. “Measuring strain with strain gages.” Accessed
Di Donna, A., A. Ferrari, and L. Laloui. 2016. “Experimental investigations May 25, 2016. http://www.ni.com/white-paper/3642/en/.
of the soil–concrete interface: Physical mechanisms, cyclic mobiliza- Ng, C. W. W., C. Shi, A. Gunawan, L. Laloui, and H. L. Liu. 2015.
tion, and behaviour at different temperatures.” Can. Geotech. J. 53 (4): “Centrifuge modelling of heating effects on energy pile performance
659–672. https://doi.org/10.1139/cgj-2015-0294. in saturated sand.” Can. Geotech. J. 52 (8): 1045–1057. https://doi
Elzeiny, R., M. T. Suleiman, M. Abu Qamar, S. Xiao, and M. Al-Khawaja. .org/10.1139/cgj-2014-0301.
2018. “Axial pull-out response of a small-scale concrete pile subjected Pasten, C., and J. C. Santamarina. 2014. “Thermally induced long-term dis-
to cyclic thermal loading in sand.” In Proc., Int. Foundations Conf. and placement of thermoactive piles.” J. Geotech. Geoenviron. Eng. 140 (5):
Equipment Expo 2018, 709–714. Reston, VA: ASCE. 06014003. https://doi.org/10.1061/(ASCE)GT.1943-5606.0001092.

© ASCE 04020011-9 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2020, 146(4): 04020011


Ramadan, M. I., S. D. Butt, and R. Popescu. 2013. “Offshore anchor piles Takeuchi, M., M. Hiramoto, N. Kumagai, N. Yamazaki, A. Kodaira, and
under mooring forces: Numerical modeling.” Can. Geotech. J. 50 (2): K. Sugiyama. 1993. “Material properties of concrete and steel bars at
189–199. https://doi.org/10.1139/cgj-2012-0251. elevated temperatures.” In Proc., 12th SMiRT Int. Conf., 133–138.
Randolph, M. F., and C. P. Wroth. 1978. “Analysis of deformation of Stuttgart, Germany: V. H. Stuttgart.
vertically loaded piles.” J. Geotech. Eng. Div. 104 (12): 1465–1488. Xiao, S., M. T. Suleiman, R. Elzeiny, C. Naito, S. Neti, and M. Al-Khawaja.
Rotta Loria, A. F., A. Di Donna, and L. Laloui. 2015. “Numerical study on 2018. “Effect of temperature and radial displacement cycles on soil–
the suitability of centrifuge testing for capturing the thermal-induced concrete interface properties using modified thermal borehole shear
mechanical behavior of energy piles.” J. Geotech. Geoenviron. Eng. test.” J. Geotech. Geoenviron. Eng. 144 (7): 04018036. https://doi.org
141 (10): 04015042. https://doi.org/10.1061/(ASCE)GT.1943-5606 /10.1061/(ASCE)GT.1943-5606.0001892.
.0001318. Yavari, N., A. M. Tang, J. M. Pereira, and G. Hassen. 2014. “Experimental
Rotta Loria, A. F., and L. Laloui. 2017. “Thermally induced group effects study on the mechanical behaviour of a heat exchanger pile using physi-
Downloaded from ascelibrary.org by Ahsanullah University of Science and Technology on 12/28/21. Copyright ASCE. For personal use only; all rights reserved.

among energy piles.” Géotechnique 67 (5): 374–393. https://doi.org/10 cal modeling.” Acta Geotech. 9 (3): 385–398. https://doi.org/10.1007
.1680/jgeot.16.P.039. /s11440-014-0310-7.
Stewart, M. A., and J. S. McCartney. 2014. “Centrifuge modeling of soil- Zhao, R., A. K. Leung, and I. Anastasopoulos. 2017. “Numerical investigation
structure interaction in energy foundations.” J. Geotech. Geoenviron. of the capacity and settlement of a thermo-active pile in saturated sand.” In
Eng. 140 (4): 04013044. https://doi.org/10.1061/(ASCE)GT.1943 Proc., 70th Canadian Geotechnical Conf. and the 12th Joint CGS/IAH-
-5606.0001061. CNC Groundwater Conf. Ottawa: Canadian Geotechnical Society.

© ASCE 04020011-10 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2020, 146(4): 04020011

You might also like