You are on page 1of 31

Abstract

Purpose – To predict the real behavior of the full-scale model using a scale model, optimized

simulation should be achieved. In Reinforced Concrete (RC) models, scaling can be

substantially critical than in single-material models because of multiple reasons such as

insufficient bonding strength between small-diameter steel bars and concrete, and excessive

aggregate size. Overall, there is a shortfall of laboratory and field-testing studies on the

behavior of secant pile walls under lateral and axial loads. Accordingly, the main objective of

this study was to investigate the validity and the performance of the 1/10th scaled RC secant pile

wall under the influence of different types of loading.

Design/methodology/approach– The structural performance of the examined models was

evaluated using two types of tests (bending and axial compression). A self-compacting concrete

mix has been suggested, which provides the best concrete mix workability and appropriate

compressive strength.

Findings– Under axial and bending loads, the failure modes were typical, where the plain and

reinforced concrete piles worked in tandem to support the load throughout the loading process,

even when they failed. The experimental results were relatively consistent with some empirical

equations for calculating the modulus of elasticity and the critical buckling load, which

confirms the validity of the presented model.

Originality/value– According to the analysis and verification of the experimental tests, the

proposed 1/10th scaled RC secant pile model can be utilized for future laboratory purposes,

especially in the field of geotechnical engineering.

Keywords small-scale, secant pile wall, self-compacting concrete.

Paper type Research paper

1
1. Introduction
Field testing, full-scale tests, or prototypes are used to depict the real behavior of

various engineering materials or structures, as well as to verify design tools (Chakrabarti,

2005). However, the utilization of such methods requires a relatively high cost and

demands certain circumstances that are often difficult to achieve, especially in issues

related to the soil-structure interaction (Laefer and Erkal, 2016). Therefore, many

researchers rely on small-scale models to assemble data and to characterize the

performance of the elements under investigation. Although the utilization of small-scale

models is regarded as an effective approach in scientific research, there are potential

negative consequences due to the scale effect problem. For instance, when the features of

the materials are not scaled appropriately, the elements may not act in a way that reflects

the real behavior (Hou and Wu, 1997). So, if a perfect (optimized) simulation is achieved,

the data collected from a small-scale model can be utilized to anticipate the behavior of a

full-scale model (Laefer and Erkal, 2016). As is well known, for physical modeling,

essential similarities must be satisfied to achieve (conduct) meaningful and useful scaling

and modeling (Sheng, 2019). For meaningful physical modeling, geometrical similarity

must be satisfied. If the physical modeling is made to be useful, the kinematic and dynamic

similarities must be partially or fully satisfied, essentially depending on the specific

problem (Sheng et al., 2014).

The geometrical similarity could be defined as all linear lengths of one element with a

constant scale factor in comparison to the corresponding linear lengths of the second

element (Sheng et al., 2014). Also, kinematic, and dynamic similarities mean that all

velocity and force vectors in the model must be orientated in the same direction as those in

2
the full-scale model, with the related magnitudes linked by a fixed scale factor (Wood,

2017).

The dimensional analysis aids in the fulfillment of the aforementioned requirements by

creating dimensionless sets, known as π-sets (Van Geem, 2007). It consists of dimensional

variables for both the small-scale and the full-scale models. The identification of model

parameters is made easier by conforming the dimensionless sets for small-scale and full-

scale models. Buckingham was the first to suggest this approach, which is known as the

"Buckingham π theorem" (Buckingham, 1914). Moreover, several researchers contributed

have to the development of this approach (Langhaar, 1962; Vignaux, and Scott, 1999;

Sterrett, 2017).

Generally, scaling can be more crucial in the reinforced concrete models than in

single-material models. This is due to a variety of factors, such as weak bonding strength

between small-diameter steel bars and concrete, excessive aggregate size, and unsuitable

strength levels (Lu et al., 1999; Ohtaki, 2000). Table 1 summarizes the main reported

characteristics of the reinforced concrete scale models in recent worldwide studies.

3
*
Table 1 Summary of reported main characteristics of scaled reinforced concrete models in worldwide studies.
Concrete
Geometrical Overall model Steel bar Steel yield
compressive
Type of scale (model dimension diameter stress (fy)
Authors Types of tests strength (fc)
elements to
L1 L2 L3
prototype) mm MPa MPa
mm mm Mm
Earthquake RC multistory
Lu (2002) 1/5.5 2370 366 600 6 30 430
simulation tests structures
Quasi-static
** 4.2
Kim et tests and Rectangular RC
1/5 400 80 60 (threaded 9.5 to 40.8 400
al. (2009) pseudo column
rod)
dynamic tests
** Concrete faced
Liu et 1-g large-scale
rock-fill dams 1/214 1500 4000 800 - 0.03 -
al. (2014) shaking table
(CFRDs
4 mm,
Liao et al. Shear, tension, 327 1000 to
Secant pile wall 1/3 861 and 12 30 335
(2014) and bending (Dp) 3000
mm
0.7 to 1.4
Guoxing
Shaking table Subway (galvaniz
et al. 1/30 705 577 1216 7.5 -
tests structure ed steel
(2015)
wire)
**
Laefer 12.2
soil-structure RC building and
and Erkal 1/10 4.9 4.3 3.0 (Lead 3.45 19
experiments excavation wall
(2016) wires)
** 41‐floor
Shen
Shaking table building with a
and Qian 1/50 3.5 - - - 12.8 to 23.4 -
tests frame–core tube
(2019)
structure.
(*) Data from Laefer and Erkal (2016) has been used and updated.
(**) Material strength similitude was taken into consideration.

4
1.1Secant Pile Walls Background
Recently, with the rapid development of cities, the necessity of retaining systems

utilization has increased for constructing high-rise buildings, tunnels, and subways

(Alavinezhad and Shahir, 2020). Retaining systems could be addressed as the structural

elements that resist lateral earth pressures and are used to retain two different earth levels

(Chehadeh, 2015). To support excavations, many types of walls are utilized. Commonly

utilized wall types involve gravity, cantilever, or embedded walls; sheet pile walls

(SHPW); diaphragm walls (DW), secant pile walls (SPW), and tangent pile walls (TPW)

(Azzam and Elwakil, 2017). There are two popular piling methods according to their

construction techniques which are precast driven piles, and bored cast-in-situ piles (Luo et

al., 2019). Continuous flight auger (CFA) pile, also known as auger cast pile, is the most

common type of bored pile currently in use (Hameedi et al., 2021). Driven piles are

categorized according to the materials used in their construction (Dolati and Mehrabi,

2021). Recently, prestressed-precast concrete pile (PPCP) is the frequently utilized type of

driven pile (Khedmatgozar Dolati and Mehrabi, 2021).

The performance of retaining walls is reflected by the magnitude and distribution of

lateral wall deformation and vertical ground movement behind the wall (Zahmatkesh, and

Choobbasti, 2015). According to Moormann (2004), Wang et al. (2010), Zhang et al.

(2018), Zhao et al. (2019), Ying et al. (2020), Amari and Houhou (2021), and El-Nimr et

al. (2022), the lateral wall deformations and ground surface settlements are linked to

multiple factors such as (1) soil type and condition; (2) excavation geometry; (3) basal

heave stability; (4) earth pressures; (5) type of retaining wall and its construction material;

(6) wall stiffness/rigidity (7) spacing and stiffness of the struts/supports; (8) construction

method; (9) time effects; (10) workmanship; (11) soil stiffness; (12) adjacent loads; (13)

5
embedded depth; and (14) soil consolidation, creep, and preloading.

The secant pile walls are created by made of a line of cast-in-place overlapping piles.

Liao et al. (2014) reported that there are two types of cast-in-place secant pile walls,

according to the design requirements. The first type, known as the Reinforced-Reinforced-

Reinforced (RRR), or hard/hard type, basically consists of Reinforced Concrete Piles

(RCPs) constructed successively, as demonstrated in Figure 1a. The second type, known as

the Plain-Reinforced-Plain (PRP) type or hard/firm type, is composed of Reinforced

Concrete Piles (RCPs) and Plain Concrete Piles (PCPs) placed alternately, as shown in

Figure 1b. In the first stage of construction, the drilled and poured piles are known as

female piles. In the second stage of construction, the subsequent piles are known as male

piles, and they cut into the female piles before the female piles' concrete coagulates.

1st stage 1st stage


(female piles) (female piles)

2nd stage
2nd stage
(male piles)
(male piles)

(a) (b)
Figure 1 Secant pile types according to Liao et al. (2014): (a) RRR-type, or hard/hard
type (b) PRP-type, or hard/firm type.
For designing or modeling, the secant pile wall is usually simplified as an equivalent

continuous wall, and the intersection between secant pile faces is regarded as fully bonded

(Finno et al. 2002; Altuntas et al. 2009; Ramadan and Meguid, 2020). The secant pile walls

have been studied in a few case histories (Wong et al., 1997; Long, 2001; Moormann,

2004; Cetin, 2016; Zhang et al., 2018). For field testing of secant pile walls, limited studies

have been carried out such as Finno and Bryson (2002), Mohamad et al. (2011).

Regarding laboratory tests, Liao et al. (2014) examined the interface and the

interaction between secant piles under tension, shear, and bending loads. The geometrical,
6
and the material scales (model to prototype) were defined as CL =1/3, and CE = 1,

respectively. They concluded that the early strength of primary piles, the time intervals of

bonding (TIB), and the effects of bonding quality could be a source of wall stiffness

reduction. In general, their proposed dimensions are relatively large, and the length of the

wall was not included in the process of similarity. Therefore, the use of these models in the

laboratory with the presence of soil requires a very large space and high cost.

Besides supporting the excavation, the piled wall may be intended for other purposes,

like being part of the permanent foundation of a structure (Azzam and Elwakil, 2017;

Elzain and Dafalla, 2018). Few researchers have considered using piled retaining walls to

support both vertical and lateral earth loads (Underwood and Greenlee, 2010; Azzam and

Elwakil, 2017; Sylvain et al., 2017). However, this concept has not reached the final design

stage due to the lack of experimental studies that accept them as bearing elements.

To sum up, there is a lack of laboratory studies on the behavior of the secant pile walls

under lateral and axial loads. There hasn't been a single laboratory study on the behavior of

the interaction between soil and secant pile wall. As a result, a scale model of reinforced

concrete is required for achieving meaningful and useful scaling and modeling.

Accordingly, this paper aims to study the structural behavior of the 1/10th scaled RC

model. Initially, a concrete mix has been designed to achieve the best possible concrete

mix workability and appropriate compressive strength (fc). Finally, check the behavior of

the tested models under the influence of bending and axial loads.

2. Setup of The Model Test


2.1Design of the Similarity Ratio
According to Brown et al. (2007), the recommended length and diameter for CFA

secant pile walls range from 10 to 18 m, and 0.35 to 0.6 m, respectively. The overlapping

ratio between the piles can be taken 20% of the pile diameter (Liao et al., 2014). So, the
7
geometric characteristics of the utilized prototype for the modeling were popular secant

pile walls with a diameter of 600 mm, an overlapping width of 120 mm, and a length of

15000 mm.

Generally, the model's dimensions were decreased by 1/10 based on the actual test

conditions and the feasibility of the test operation. Where the model should be as small as

possible to facilitate its use and potential application in soil research, as long as this does

not influence its structural behavior.

The similarity ratios between physical variables were derived by Buckingham π theory

(Li et al., 2020; Perumalsamy and Ranganathan, 2021). Table 2 illustrates the similarity

law and the utilized similarity ratios for the proposed models. The geometrical and

material scales (model to prototype) were defined as CL = 1/10, CE = 1.0, respectively.

Accordingly, the model piles' diameter and the overlapping width between the secant piles

should be 60 and 12 mm, respectively. The scale reinforced concrete piles (RCPs) were

prepared to have a similar steel ratio as the prototype, which was determined as 4%. For

steel arrangement, four longitudinal steel bars (threaded rods) of 6 mm in diameter were

used, and a stirrup bar of 4 mm in diameter was spaced 100 mm apart.

Table 2 Similarity law and utilized similarity ratios for proposed models.
Parameters Similarity law Similarity constants (Prototype: model)
Geometry L CL 10
Displacement Y CY=CL 10
Elasticity modulus E CE 1
Flexural stiffness EI CEI= CECL4 104
Stress σ Cσ = CE 1
Strain ε Cε 1
Poisson ratio μ Cμ 1
Density ρ Cρ 1
Mass m Cm = CρCL3 103
Concentrated load F CF = CσCL2 102
Linear load w CL = CσCL2 10
Moment M CM = CσCL3 103

8
2.2 Material Composition
2.2.1 Samples Mold
To ensure the quality of the concrete models, the test samples were created in a

cylindrical steel framework that had been formed with detachable curved metal strips (4

mm in thickness) connected by screw bolts (Figure 2). The concrete was placed by

installing the steel framework vertically and pouring the concrete from the top. The single

reinforced pile (SRP) was poured into a hollow plastic pipe with an inner diameter of 60

mm.

(a)

(b)

Figure 2 Concrete samples making: (a) Steel framework for concrete casting, (b) Secant
pile wall model.
2.2.2 Concrete mix
Self-Compacting Concrete (SCC) has been deliberately selected to be able to fill every

corner of the framework and encapsulate all reinforcements only under the influence of

gravitational forces, without segregation or bleeding. SCC is especially beneficial in

situations when pouring is problematic, such as small-scale reinforced concrete elements

or complicated work forms. For achieving the best possible concrete mix workability and

9
appropriate compressive strength (fc), several concrete mixes have been attempted. Figure

3 shows some defective samples that were excluded due to the inadequacy of the concrete

mix for the small concrete section. As shown in Figure 3, some defects appeared, such as

the phenomenon of honeycombs in concrete or the crushing of the concrete surface during

the disassembly of the metal framework. The quality of the mixtures was judged during the

mixing and pouring phase (workability, and bleeding), the removal of the metal framework

phase (quality of concrete surface), and the compressive strength determination phase.

Figure 3 Defective samples due to the inadequacy of the concrete mix.

At the age of 28 days, the cubic compressive strength (fc) of the model concrete was

approximately 45 MPa for both PCPs and RCPs. Table 3 shows the components of the

selected concrete mix.

Table 3 Concrete mix components.


No Constituents, (unit) Quantity
.
1 10 mm nominal sized crushed basalt aggregate, (Kg) 770
2 1.00 mm nominal sized sand, (Kg) 770
3 Water (W), (Kg) 175
4 Cement (C), (Kg) 520
5 Silica Fume (S.F), (Kg) 27
6 Normal Plasticizers, (Kg) 17
7 Water–Cement ratio (W/C) 0.33

10
2.2.3 Reinforcing Steel

Reinforcing bars with a diameter of 6 mm are not usually available among the

standard deformed steel bars for a reinforced concrete section. Thus, a round threaded rod

was used for the test model with the same material properties as the standard deformed

steel bars. Previous researchers have experimentally validated the use of threaded rods in

small-scale models while taking reinforcing bond strength into account (Kim et al., 1988,

2009). For the stirrups, smooth bars with a diameter of 4 mm were used. Figure 4 presents

the reinforcement of the model.

(b)

(a)

(c)
Figure 4 Reinforcement of the test models (a) Utilized circular stirrups, (b) Threaded
rod, and (c) Setup for reinforcement.
2.3 Loading Patterns and Monitoring
For each test (bending and axial compression), the loading pattern and monitoring

system were set as illustrated in Figures 5 and 6. All the testing data were automatically

recorded using a laptop connected to a UCAM-550A strain data logger instrument.

For the bending test, one end of the sample was supported with hinge support and the

other with rolling support. The vertical load P was applied in the middle span and

11
transmitted to the sample via two smooth steel rolls situated one-third and two-thirds of the

span away from one end (Figure 5). The specimen loading was force controlled. Deflection

at the center of the model was measured with the help of an LVDT as shown in Figure 5.

Vertical Load

Hinged Support Roller Support

LVDT

(a)
Hydraulic Jack

Load Cell

Steel Rolls

Load Transfer Beam

Hinged Support Rolling Support

LVDT Data Logger

(b)
Figure 5 Four-point loading test: (a) Sketch of bending loading (b) Bending loading in
the laboratory.

For the axial compression test, two-pin supports (top and bottom) were used to allow

360° rotation as shown in Figure 6. The axial load was applied axially by the load cell,

where the loading was force-controlled (Figure 6). The model's mid-span deflections were

measured with loading increment using an LVDT, as illustrated in Figure 6.

12
Axial Load

Hydraulic Jack

Load Cell

LVDT
Mid-span
LVDT
deflection

Pin
Supports

Before loading After loading

(a)

(b)
Figure 6 Axial compression test (a) Sketch of axial compression loading (b)
Axialloading in the laboratory.
Table 4 shows the testing program for the current study. The first step in the

experimental program was the selection and design of appropriate concrete mix. Also,

Figure 7 shows the cross-section dimension of the tested models.

Table 4 Testing program.


Test name Samples Length Piles configuration
B1 Three piles (PCP-RCP-PCP)
Bending B2 Three piles (RCP-PCP-RCP)
B3 1.50 m Singular reinforced pile (SRP)
C1 Three piles (PCP-RCP-PCP)
Axial compression
C2 Three piles (RCP-PCP-RCP)

13
(a) (b) (c)
Figure 7 Samples cross-section dimension (a) Samples C1, and B1, (b) Samples C2,
and B2, and (c) Sample B3.
3. Testing Results and Discussion
All samples were loaded to failure under a monotonic static load with regular

increments to guarantee the uniformity of the data acquired from bending and axial

compression tests.

3.1 Bending Test Results


As illustrated in Table 4, three samples were prepared for the four-point loading,

including one sample of the PRP type, one sample of the RPR type, and one sample of a

singular reinforced pile (SRP).

3.1.1 Progressive Failure and Crack Observations


Figure 8 depicts observations on the cracks formation and the failure mechanism of the

tested models under bending loading. The cracks' evolution can be summarized in the

following critical steps. Firstly, no cracks were found either in the plain or the reinforced

concrete piles at the beginning of the loading, and the entire section was in a state of

elasticity. Also, in this stage, the plain and reinforced concrete piles worked in tandem to

support the load throughout the loading process. Secondly, cracks began to appear only in

the plain concrete piles, and the section started to behave in an elastoplastic state (Figure

8b). Thirdly, with increasing the loading, the cracks in the plain concrete piles widened,

and other cracks began to appear in the reinforced concrete piles (Figure 8c). Fourthly, the

lower part of the plain concrete pile failed, which means that the reinforced pile will bear

the majority of the load, and this resulted in relatively significant cracks in the reinforced
14
pile. Finally, the bending stress in the compression zone was greater than the compression

strength of concrete due to the small size of the tested model, which led to the failure of the

compression zone (Figure 8d).

For comparison, sample B3 (single-reinforced pile) failed in the same way as the other

samples (B1, and B2). Furthermore, the cracking load of sample B3 was significantly

smaller than the other samples, and its deflection developed much faster than them.

(a) Before loading

RCP
Cracks
PCP

(b) Initial cracks on the PCP

Cracks

(c) Cracks began to appear on the RCP


PCP

Compression zone

(d) Failure
Figure 8 Cracks of theincompression
observation the bending zone
test (Sample B1).
For the cross-sectional direction of the samples B1 and B2, it can be concluded that the
15
growth of cracks passed through four basic stages. Figure 9 summarizes these stages for

samples B1 and B2.

Stage (1) Stage (2) Stage (3) Stage (4)


No cracks Initial cracks on PCP Cracks extend to RCP Cracks widen until failure

Cracks
Cracks Cracks
(a)

Cracks
Cracks Cracks
Stage (1) Stage (2) (b) Stage (3) Stage (4)
No cracks Initial cracks on PCP Cracks extend to RCP Cracks widen until failure

Figure 9 Sketch of cracks observation for the cross-sectional direction (a) Sample B1,

(b) Sample B2.

Figure 10 illustrates the cracks' evolution in the longitudinal direction of the samples

under bending loading. At a low level of loading, small cracks occur in the zone of

maximum bending moment (approximately in the middle third of the span as shown in

(Figure 10b)). The first flexural crack appeared on the PCPs of samples B1, and B2 at loads

2.8 kN, and 2.9 kN, respectively. For RCPs, the first flexural cracks appeared on samples

B1, B2, B3 at loads 4.9 kN, 6.8kN, and 1.9 kN, respectively (Table 5). By increasing the

applied load, it appears that the cracks are evenly distributed across the sample's span

length (Figure 10c). When the final flexural cracks were achieved, the applied load was

considerably reduced and the vertical deflection suddenly increased. The applied ultimate

loads at flexural failure were 9.9 kN, 12.2 kN, and 7.9 kN for samples B1, B2, and B3,

respectively (Table 5). At a high level of loading, a compressive fracture of the concrete

occurred just below the point loads and in the middle of the span (Figure 10d). Also, the

cracks' evolution of the presented models agreed well with the results of the normal RC

beams which studied by Kristiawan et al. (2017).

16
P/2 P/2 (a) Stage (1)
No cracks

P/2 P/2 (b) Stage (2)


The 1st flexural crack
Increasing the applied load

Cracks

P/2 P/2 (c) Stage (3)


The final flexural cracks

Cracks

P/2 Compressive P/2 (d) Stage (4)


fracture The final flexural cracks

Cracks

Figure 10 Sketch of cracks observation in the longitudinal direction of the tested samples.

Figure 11 represents the load-deflection curves (middle point for Samples B1, B2, and B3) of the

tested models, which corresponded to the previous cracks’ observations. Additionally, under

bending loads, the failure modes were typical, where the plain and reinforced concrete piles

worked in tandem to support the load throughout the loading process, even when they failed.

Where the ultimate loads of samples B2 (PRP-type), and B3 (RPR-type) were about 26%, and

55%, respectively, higher than that of sample B1 (single reinforced pile), as presented in Table

5. Table 5 illustrates the details of the critical and ultimate stages throughout the bending

loading process. It can be noticed that the displacement ductility factors (R) were 8.8, 9.6, and

4.3 for samples B1, B2, and B3, respectively.

17
14
Elastic stage Elasto-plastic stage
12 Failure stage

10

B1
8
Load, kN

B2
B3
6

Elastic stage Elasto-plastic stage Failure stage


0
0 20 40 60 80 100 120
Deflection, mm
Figure 11 Load-deflection curves of bending tests (mid-span deflection, samples B1,
B2, and B3).

Table 5 Cracking and ultimate loads of samples in bending test.

Initial cracks stage Failure stage For comparison


P −P3 δ
Sample Pcr (kN) δPcr δRcr Pu δu Lc = ui R= u
s (kN) P3 δ cr
PCPs RCPs mm mm mm % Ratio
20.
B1 2.8 4.9 6.9 9.9 61 25 8.8
1
25.
B2 2.9 6.8 7.1 12.2 68 55 9.6
0
B3 - 1.9 - 12.0 7.9 52 0.0 4.3
Where, Pcr: cracking load (initial cracks), Pu: ultimate load, δPcr & δR: deflection at first observed
crack on the PCP, and RCP, respectively, δu: deflection at failure, Lc : increasing ratio in bending
load, R: the displacement ductility factor according to Tawfik et al. (2015).

3.1.2 Flexural stiffness


The flexural stiffness of the primary secant pile wall model was determined by

conducting a simple bending test with four-point loading. According to Perumalsamy and

Ranganathan (2021), flexure stiffness (EI) was calculated by using Equation 1.

P∗a
EI = (3 l 2−4 a2) (1)
24∗Deflection

Where E is the Young's modulus (N/mm2), I is the area moment of inertia of the cross-
18
section (mm4), P is the applied load (N), a is the distance between load point and the

closest support (mm), and l is the length between the two supports (mm).

The apparent modulus of elasticity (Ea) was also calculated by Christoforo et al.

(2012). It can be obtained from Equation 2.

Ea I =¿ ¿ (2)

Where Ea I flexure stiffness (N.mm2), P50% and P10% are the loads (N) corresponding to

10% and 50% of the ultimate applied load (Pu), respectively, while δ50% and δ10% are the

displacements (mm) corresponding to 10% and 50% of the ultimate applied load (Pu),

respectively.

Table 6 illustrates the calculated modulus of elasticity values according to Equation

(2), the Egyptian Code of Practice (ECP) method (ECP203, 2018), and the American

Concrete Institute method (ACI 318-11). Based on Equation (2), the elasticity modulus of

the tested models was calculated, and its values were 2.4x104 N/mm2, 3.0x104 N/mm2, and

2.0x104 N/mm2 for samples B1, B2, and B3, respectively. It is noticeable that there are some

differences between the experimental and the empirical results. Where the empirical

equations did not consider the percentage of reinforcement in the model's section, the

configuration of the steel bars, as well as the geometric properties of the section, and

depended only on the compressive strength of cubes or cylinders. According to Table 6,

the average value of elasticity modulus for samples B1 and B2 was found to be extremely

close to the values computed using the empirical equations, especially the method of the

American Concrete Institute (ACI 318-11). Also, the average value of the elasticity

modulus for samples B1, B2, and B3 is approximately 10% less than the value calculated

using the equation of the American Concrete Institute (ACI 318-11). As a result, it is

reasonable to conclude that the average values of the experimental results are compatible

19
with the results of the empirical equations.

Table 6 Modulus of elasticity for tested samples using various methods.

EEC −Ecal E AC −E cal


P10% P50% δ10% δ50% I Ecal EEC EAC
E EC E AC
No.

N N mm mm mm4 N/mm2 N/mm2 N/mm2 % %


B 100 500 20. 1.0x10
4.4 6 2.4x104 17.2 14.3
1 0 0 6
B 130 650 19. 1.0x10 2.9*10
2.7 6 3.0x104 4 2.8*104 -3.4 -7.1
2 0 0 9
B 370 27. 6.3x10
740 4.5 5 2.0x104 31.0 28.5
3 0 1
Average (Samples B1, B2, and B3) ≈2.5x104 13.8 10.7
≈2.75x10
Average (Samples B1, and B2) 4 5.1 1.8
Where Ecal: calculated modulus of elasticity according to Christoforo et al. (2012), Fcu: cube
compressive strength, Fcy: cylinder compressive strength (Fcy≈0.80Fcu), EEC = 4400 √ Fcu (ECP203,
2018), EAC= 4700 √ Fcy (ACI 318-11).

Figure 12 demonstrates the variability of flexural rigidity (EI) with the applied load,

based on Equation 1 and the experimental findings of bending tests (for samples B1, B2,

and B3). As illustrated in Figure 12, the initial flexural stiffness of samples B1 and B2 was

substantially greater than that of sample B3. Moreover, the flexural stiffness of samples B1

and B2 reduced significantly with load increment, but the stiffness of sample B3 reduced

more gently. The presence of plain concrete piles may be related to the varying behavior of

the samples stated previously. Liao et al. (2014) revealed the same conclusion.

40
Flexural Rigidity (EI), kN.m2 * 10-3

35 B3

30 B1
B2
25

20

15

10
0 2 4 6 8 10 12
Load, kN

20
Figure 12 Flexural rigidity-load curves (samples B1, B2, and B3).

3.2Axial Compression Test Results


As shown in Figure (6a), if the wall is fully aligned, the axial compressive load P can

be increased until internal stress reaches the compressive strength of the material.

Interestingly, the column is likely to collapse before reaching this load due to buckling as

shown in Figure (6a) on the right (Paspuleti, 2005). Buckling in the structural parts may be

disastrous if it happens during regular use. The structural item can no longer sustain even a

fraction of the applied vertical load when it begins to deform (Mott and Untener, 2017). In

general, buckling-related pile instability is among the most common triggers of

superstructure collapse (Basha, 2021).

For the axial compression tests, two samples were created 1.50 m long (C1, and C2), as

indicated in Table 4. The results of the axial compression tests were presented in Figure

13.

As illustrated in Figure 13, the lateral deformations (mid-span point) of the two

samples were divided into three stages. Firstly, there's the initial stage, in which the applied

vertical load raised (approximately 0- 0.65Pf) while the sample's lateral deformation

increased gently. Secondly, the stage of steady growth, in which the lateral deformation of

the sample increased rapidly, and the axial vertical load approached the ultimate load

(approximately 0.65-1.0Pf (. When the maximum load was reached, the curves switched

their direction from rising to falling. Thirdly, the failure stage, in which the applied load

significantly reduced, and the lateral deformation suddenly increased. It was also

confirmed that the plain and reinforced concrete piles in the samples C1 and C2 deformed

consistently and worked together to maintain the load throughout the loading procedure,

even when they failed. In comparison, with a constant loading rate for both samples, the

21
lateral deformations of sample C1 were substantially greater and evolved much faster than

sample C2. Also, the ultimate load of sample C2 was greater by 10% more than the load of

sample C1. This means that the difference in the distribution of steel bars within the section

affects its axial capacity. Figure 14 shows the axial loading steps failure for sample C1.

90
80
70
60
Axial load, Kn

50
40
C
30 1
20 C
2
10
0
0 5 10 15 20 25 30 35
Mid-span deflection, mm

Figure 13 Axial load–lateral deflection curves (middle point, Samples C1, and C2).

According to Euler's formula, the critical load (Pcr) is the compressive load at which

the tested element will suddenly bend or buckle. It is given by the following formula (Yoo

and Lee, 2011):

π 2 EI
Pcr = 2 (3)
(kL)
Where Pcr is the critical load (N), E is the modulus of elasticity (N/mm2), I is the

moment of inertia (mm4), L is the length of the element (mm), and K is the column

effective length factor (for pin-pin ends k=1.0).

Table 7 compares the values of the observed buckling load (Pf) of the tested samples

and the calculated critical load from Euler's equation. As shown in Table 7, the buckling

22
load from the experimental results is approximately 13% and 23% less than the calculated

load from Euler's equation for samples C1 and C2, respectively. Thus, the actual buckling

load increases with the increase in the number of reinforced concrete piles in the tested

sample. This may explain the presence of some differences between the theoretical and the

actual loads, as it is expected that the theoretical and the experimental results will be very

close in the case of reinforcing the entire concrete section. The difference in the results

may also be due to the initial eccentricity of the axial load, geometrical defects,

imperfections with the specimens (including pre-buckling), and crack generation, as

mentioned by Kotsmid et al. (2015), and Kashyap et al. (2018). However, the failure

patterns were typical which demonstrated the compatibility of the performance of the

concrete mix and reinforcing steel for the presented models.

(b) End
(a) Start of loading Figure 14ofAxial
loading (c) Tension
loading steps failure side (d) Compression side
(sample C1).
23
Table 7 The values of the buckling load (Pf) and the calculated critical load from
Euler's equation (Pcr).
Pcr =
Reduction
I E L π 2 EI Pf
Sampl ratio
K (kL)2
e
(Pcr-Pf)/Pcr
mm4 N/mm2 mm kN kN
(%)
80.5
C1 14.20
6 4 1.0 0
1.0x10 2.75x10 1700 93.82
0 71.4
C2 23.89
0
Where, Pf: buckling load, Pcr: critical load, E: the modulus of elasticity from bending tests, I:
the moment of inertia, L: the length of the element, and K: the column effective length factor (for
pin-pin ends k=1).

4. Conclusions
The main objective of this study is to investigate the validity and the performance of

1/10th scaled reinforced concrete secant pile wall under the influence of different types of

loading and then determine the possibility of using the presented models for future

laboratory purposes. The structural performance of the examined models was evaluated

using two types of tests (bending and axial compression). The key findings of this paper

can be summarized as follows:

1) The suggested self-compacting concrete mix was able to fill every corner of

the small framework and envelop all reinforcements steel, without segregation

or bleeding, and to achieve the best possible concrete mix workability and an

appropriate compressive strength (fc = 45 MPa).

2) Under axial and bending loads, the failure modes were typical, where the plain

and reinforced concrete piles worked in tandem to support the load throughout

the loading process, even when they failed. This demonstrates the

compatibility of performance and homogeneity of the concrete mix and

24
reinforcing steel.

3) According to the results of the bending tests, the bending load capacity is

determined by a composite section with the participation of the entire section

of reinforced concrete piles and a proportion of the section of plain concrete

piles. Steel reinforcement is also a significant factor regarding the bending load

capacity. Where the ultimate loads of samples B2 (PRP-type), and B3 (RPR-

type) were about 26%, and 55%, respectively, higher than that of sample B1

(single reinforced pile).

4) In axial compression tests, the lateral deformation of sample C1 (PRP-type) was

relatively larger and developed faster than the sample C2 (RPR-type).

Furthermore, the ultimate axial load of sample C2 was about 10% greater than

the load of sample C1. Therefore, the difference in pile distribution within the

model test section affects its performance of axial load capacity.

5) The laboratory results were relatively consistent with some empirical equations

for calculating the modulus of elasticity and the critical buckling load, which

confirms the validity of the presented model. The highlighted differences

between the theoretical and actual experimental loads may be due to the

geometrical defects or the loading procedures.

References

Alavinezhad, S. P., & Shahir, H. (2020), “Determination of apparent earth


pressure diagram for anchored walls in c–φ soil with surcharge”, World Journal of
Engineering.

Altuntas, C., Persaud, D., & Poeppel, A. R. (2009), “Secant pile wall design and
construction in Manhattan, New York”, In Contemporary Topics in Ground
Modification, Problem Soils, and Geo-Support (pp. 105-112).

25
Amari, T., & Houhou, M. N. (2021), “3D numerical analysis of pile response due
to soil movements induced by an adjacent deep excavation. World Journal of
Engineering.

Azzam, W. R., & Elwakil, A. Z. (2017), “Performance of axially loaded-piled


retaining wall: Experimental and numerical analysis”, International Journal of
Geomechanics, 17(2), 04016049.

Basha, A. M. (2021), “Post Buckling Behavior of Slender Piles Partially Embedded


in Sand Soil under Axial Load”, KSCE Journal of Civil Engineering, 25(3), 757-767.

Brown, D. A., Dapp, S. D., Thompson, W. R., & Lazarte, C. A. (2007), “Design
and construction of continuous flight auger piles (No. FHWA-HIF-07-03)”, United
States, Federal Highway Administration. Office of Technology Applications.

Buckingham, E. (1914), “On physically similar systems; illustrations of the use of


dimensional equations”, Physical review, 4(4), 345.

Cetin, D. (2016), “Performance of soilnailed and anchored walls based on field


monitoring data in different soil conditions in Istanbul”, Acta Geotechnica Slovenica,
13(1), 49-63.

Chakrabarti, S. (2005), “Handbook of Offshore Engineering (2-volume set)”,


Elsevier.

Chehadeh, A. S. (2015), “Analysis and design of circular shafts using finite element
method (Doctoral dissertation)”.

Christoforo, A. L., Rocco, F. A. L., Morales, E. A. M., Panzera, T. H., &


Borges, P. H. (2012), “Numerical evaluation of longitudinal modulus of elasticity of
Eucalyptus grandis timber beams”, International Journal of Agriculture and Forestry,
2(1), 166-170.

Dolati, S. S. K., & Mehrabi, A. (2021, April), “Review of available systems and
materials for splicing prestressed-precast concrete piles”, In Structures (Vol. 30, pp. 850-
865). Elsevier.

E.C.P. 203-2018. (2018), “Egyptian code for design and construction of reinforced
concrete structures”, Housing and Building National Research Center, Ministry of

26
Housing, Utilities and Urban Planning, Cairo.

El-Nimr, M. T., Basha, A. M., Abo-Raya, M. M. & Zakaria, M. H. (2022), “


General deformation behavior of deep excavation support systems: A review”, Global
Journal of Engineering and Technology Advances, 10(01), pp. 039–057. doi:
10.30574/gjeta.2022.10.1.0181.

Elzain, M. I. Y., & Dafalla, M. (2018, July), “Utilizing Secant Pile Walls as
Retaining Structures and Bridge Abutments”, In Civil Infrastructures Confronting
Severe Weathers and Climate Changes Conference (pp. 72-77). Springer, Cham.

Finno, R. J., & Bryson, L. S. (2002), “Response of building adjacent to stiff


excavation support system in soft clay”, Journal of performance of constructed facilities,
16(1), 10-20.

Guoxing, C., Su, C., Xi, Z., Xiuli, D., Chengzhi, Q. I., & Zhihua, W. (2015),
“Shaking-table tests and numerical simulations on a subway structure in soft soil”, Soil
Dynamics and Earthquake Engineering, 76, 13-28.

Hameedi, M. K., Fattah, M. Y., & Abd Al-Kaream, K. W. (2021), “FIELD


STUDY ON SOFT SOIL IMPROVEMENT USING CONTINUOUS FLIGHT AUGER
(CFA) PILES”, GEOMATE Journal, 21(86), 159-166.

Hou, T. Y., & Wu, X. H. (1997), “A multiscale finite element method for elliptic
problems in composite materials and porous media”, Journal of computational physics,
134(1), 169-189.

Kashyap, S. K., Kumar, S., Mallick, M., Singh, R. P., & Verma, M. (2018), “A
comparative study between experimental and theoretical buckling load for hollow steel
column”, International Journal of Engineering, Science and Technology, 10(3), 27-33.

Khedmatgozar Dolati, S. S., & Mehrabi, A. (2021), “Alternative Systems and


Materials for Splicing Prestressed-Precast Concrete Piles”, Transportation Research
Record, 03611981211052949.

Kim, N. S., Lee, J. H., & Chang, S. P. (2009), “Equivalent multi-phase similitude
law for pseudodynamic test on small scale reinforced concrete models”, Engineering
Structures, 31(4), 834-846.

27
Kim, W., El-Attar, A., & White, R. N. (1988), “Small-scale modeling techniques
for reinforced concrete structures subjected to seismic loads”, National Center for
Earthquake Engineering Research.

Kotsmid, S., Minárik, M., & Beno, P. (2015), “Comparison of the experimental
and theoretical critical buckling force at the rod of a defined shape”, Technics
Technologies Education Management, 10(2), 139-143.

Kristiawan, S., Supriyadi, A., Sangadji, S., & Santosa, D. (2017), “Cracking
behaviour and its effect on the deflection of patched-reinforced concrete beam under
flexural loading”, In MATEC Web of Conferences (Vol. 138, p. 02021). EDP Sciences.

Laefer, D. F., & Erkal, A. (2016), “Selection, production, and testing of scaled
reinforced concrete models and their components”, Construction and Building
Materials, 126, 398-409.

Langhaar, H. L. (1962), “Dimensional analysis and theory of models (No. 530.8


L35)”.

Li, N., Men, Y., Yuan, L., Gao, H., Li, J., & Wang, B. (2020), “Study on the
mechanical characteristic of micropiles supporting landslide under step-loadings”,
Geotechnical and Geological Engineering, 38(3), 2761-2771.

Liao, S. M., Li, W. L., Fan, Y. Y., Sun, X., & Shi, Z. H. (2014), “Model test on
lateral loading performance of secant pile walls”, Journal of Performance of
Constructed Facilities, 28(2), 391-401.

Liu, J., Liu, F., Kong, X., & Yu, L. (2014), “Large-scale shaking table model tests
of aseismic measures for concrete faced rock-fill dams”, Soil Dynamics and Earthquake
Engineering, 61, 152-163.

Long, M. (2001), “Database for retaining wall and ground movements due to deep
excavations”, Journal of Geotechnical and Geoenvironmental Engineering, 127(3), 203-
224.

Lu, Y. (2002), “Comparative study of seismic behavior of multistory reinforced


concrete framed structures”, Journal of structural engineering, 128(2), 169-178.

Lu, Y., Vintzileou, E., Zhang, G. F., & Tassios, T. P. (1999), “Reinforced

28
concrete scaled columns under cyclic actions”, Soil Dynamics and Earthquake
Engineering, 18(2), 151-167.

Luo, W., Sandanayake, M., & Zhang, G. (2019), “Direct and indirect carbon
emissions in foundation construction–Two case studies of driven precast and cast-in-situ
piles”, Journal of Cleaner Production, 211, 1517-1526.

Mohamad, H., Soga, K., Pellew, A., & Bennett, P. J. (2011), “Performance
monitoring of a secant-piled wall using distributed fiber optic strain sensing”, Journal of
Geotechnical and Geoenvironmental Engineering, 137(12), 1236-1243.

Moormann, C. (2004), “Analysis of wall and ground movements due to deep


excavations in soft soil based on a new worldwide database”, Soils and foundations,
44(1), 87-98.

Mott, R. L., & Untener, J. A. (2017), “Applied Strength of Materials, SI Units


Version. CRC Press”.

Ohtaki, T. (2000), “An experimental study on scale effects in shear failure of


reinforced concrete columns”, Strain, 500(500), 36-D13.

Paspuleti, S. (2005), “Mechanical and thermal buckling of thin films”, University of


Missouri-Columbia (Master's thesis).

Perumalsamy, K., & Ranganathan, S. (2021), “Experimental Study on Pipe Sheet


Pile Wall Berthing Structure”, Indian Geotechnical Journal, 1-13.

Ramadan, M. I., & Meguid, M. (2020), “Behavior of cantilever secant pile wall
supporting excavation in sandy soil considering pile-pile interaction”, Arabian Journal
of Geosciences, 13(12), 1-13.

Shen, C., & Qian, D. (2019), “Dynamic characteristics and seismic response of
frame–core tube structures, considering soil–structure interactions”, The Structural
Design of Tall and Special Buildings, 28(3), e1575.

Sheng, W. (2019), “Wave energy conversion and hydrodynamics modelling


technologies: A review”, Renewable and Sustainable Energy Reviews, 109, 482-498.

Sheng, W., Alcorn, R., & Lewis, T. (2014), “Physical modelling of wave energy
converters”, Ocean Engineering, 84, 29-36.
29
Sylvain, M. B., Pando, M. A., Whelan, M. J., Rice, C. D., Ogunro, V. O., Park,
Y., & Koch, T. (2017), “Case History of a Full Scale Axial Load Test of Sheet Piles”, In
Geotechnical Frontiers 2017 (pp. 355-365).

Standard, A. A. (2011, August), “Building Code Requirements for Structural


Concrete (ACI 318-11)”, In American Concrete Institute.

Sterrett, S. G. (2017), “Physically Similar Systems-A History of the Concept”, In


Springer handbook of model-based science (pp. 377-411). Springer, Cham.

Tawfik, A. S., Badr, M. R., & ElZanaty, A. (2014), “Behavior and ductility of
high strength reinforced concrete frames”, HBRC Journal, 10(2), 215-221.

Underwood, C. A., & Greenlee, R. M. (2010), “Steel sheet pile used as permanent
foundation and retention systems-design and construction”, In Earth Retention
Conference 3 (pp. 129-136).

Van Geem, K. M., Žajdlík, R., Reyniers, M. F., & Marin, G. B. (2007),
“Dimensional analysis for scaling up and down steam cracking coils”, Chemical
Engineering Journal, 134(1-3), 3-10.

Vignaux, V. A., & Scott, J. L. (1999), “Theory & methods: Simplifying regression
models using dimensional analysis”, Australian & New Zealand Journal of Statistics,
41(1), 31-41.

Wong, I. H., Poh, T. Y., & Chuah, H. L. (1997), “Performance of excavations for
depressed expressway in Singapore”, Journal of Geotechnical and Geoenvironmental
Engineering, 123(7), 617-625.

Wood, D. M. (2017), “Geotechnical modelling”, CRC press.

Ying, H. W., Cheng, K., Zhang, L. S., Ou, C. Y., & Yang, Y. W. (2020),
“Evaluation of excavation-induced movements through case histories in Hangzhou”,
Engineering Computations.

Yoo, C. H., & Lee, S. (2011), “Stability of structures: principles and applications”,
Elsevier.

Zahmatkesh, A., & Choobbasti, A. J. (2015), “Evaluation of wall deflections and


ground surface settlements in deep excavations”, Arabian Journal of Geosciences, 8(5),
30
3055-3063.

Zhang, W. G., Goh, A. T. C., Goh, K. H., Chew, O. Y. S., Zhou, D., & Zhang,
R. (2018), “Performance of braced excavation in residual soil with groundwater
drawdown”, Underground Space, 3(2), 150-165.

31

You might also like