1 s2.0 S2214509524002936 Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 28

Case Studies in Construction Materials 20 (2024) e03142

Contents lists available at ScienceDirect

Case Studies in Construction Materials


journal homepage: www.elsevier.com/locate/cscm

Experimental and numerical research on shear performance of


GFRP bar reinforced seawater sea-sand concrete deep beams
without stirrups
Zhiquan Xing a, b, Yao Zhu a, *, Yongbo Shao c, d, Enlin Ma d, Kwok-Fai Chung b, e, f,
Yu Chen a, b, **
a
College of Civil Engineering, Fuzhou University, Fuzhou 350116, China
b
International and Hong Kong, Macao and Taiwan Joint Laboratory of Structural Engineering, Fuzhou University, Fuzhou 350108, China
c
School of Civil Engineering and Geomatics, Southwest Petroleum University, Chengdu 610500, China
d
School of Architecture and Civil Engineering, Xihua University, Chengdu 610039, China
e
Chinese National Engineering Research Centre for Steel Construction (Hong Kong Branch), The Hong Kong Polytechnic University, Hong Kong,
China
f
Department of Civil and Environmental Engineering, The Hong Kong Polytechnic University, Hong Kong, China

A R T I C L E I N F O A B S T R A C T

Keywords: Using glass fibre-reinforced polymer (GFRP) bars to reinforce seawater sea-sand concrete
Seawater sea-sand concrete (SWSSC) (SWSSC) is a feasible way to replace traditional concrete structures. Thus, this paper aims to
GFRP bar understand the shear response of GFRP bar-reinforced SWSSC (GFRP-SWSSC) deep beams.
Seashell content
Experimental and numerical programs were carried out on four-point shear tests of GFRP-SWSSC
Shear performance
FEM
deep beams without stirrups. Seventy specimens were tested to investigate the effects of key
parameters on shear responses, including concrete categories, seashell content, section heights,
and GFRP bar diameter. The test results indicated that the cracking strength of GFRP-SWSSC deep
beams was slightly higher than ordinary concrete deep beams. The increased section height of
GFRP-SWSSC deep beams and the decreased shell content remarkably enhanced the stiffness and
shear ultimate strength. The corresponding finite element model (FEM) of GFRP-SWSSC speci­
mens was established and validated by comparison with test results. Further, three guidelines
predictions for the shear strength of GFRP-SWSSC beams were too conservative. The new design
formulae derived from modified tension-compression theory were put forward to evaluate the
shear strength of GFRP-SWSSC deep beams, and the comparisons demonstrated that the proposed
design formulae achieved sufficient accurate predictions for practical engineering. Based on
research on shear performance, the hybrid GFRP-SWSSC structure is a feasible solution to
resource shortages, which provides a promising application prospect in marine engineering.

* Corresponding author.
** Corresponding author at: International and Hong Kong, Macao and Taiwan Joint Laboratory of Structural Engineering, Fuzhou University,
Fuzhou 350108, China.
E-mail addresses: ZhiquanXing@yeah.net, 220525041@fzu.edu.cn (Z. Xing), skinyao@163.com (Y. Zhu), yuchen@fzu.edu.cn (Y. Chen).

https://doi.org/10.1016/j.cscm.2024.e03142
Received 25 January 2024; Received in revised form 28 March 2024; Accepted 7 April 2024
Available online 9 April 2024
2214-5095/© 2024 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142

1. Introduction

Nomenclature

A Cross-section area of concrete beam


AG Cross-section area of GFRP bars
b Width of concrete beam
D Diameter of GFRP bar
H Height of concrete beam
h0 Effective height of concrete beam
L Longitudinal length of concrete beam
Le Effective span length
SC Seashell content in sea-sand
Ec Elastic modulus of concrete
EG Elastic modulus of GFRP bar
f c’ Concrete cylinder compressive strength
fcu Concrete cubic compressive strength
ft Concrete cylinder tensile strength
fG Tensile strength of GFRP bar
ki Stiffness of GFRP-SWSSC beams
Fc Strength of the compression rod
Ft Strength of the tension rod
Fn Strength of the tension rod
P Applied load
V Shear load
Vcr Cracking load
Vu Shear ultimate load
V0.8 80% of shear ultimate load
Vcr Cracking load
Vu Shear ultimate load
V0.8 80% of shear ultimate load
DC Ductility coefficient
Δu Displacement at shear ultimate load
Δy Displacement at cracking load
Δ0.8 Displacement at 80% of shear ultimate load.
ε Vertical strain
φ Reduction factor
γ Seashell coefficient
ρG Longitudinal reinforcement ratio
nG Ratio of EG to Ec

In recent years, the over-exploitation of primary sand and gravel resources has led to the shortage of river sand resources and a
series of ecological imbalances. Considering the abundant seawater and sea-sand resources, research on seawater sea-sand concrete
(SWSSC) has raised more academic interest [1–3]. However, the chloride ions in unwashed sea-sand would induce accelerated
corrosion effects on the internal steel bars [4]. The desalination of washing sea-sand would result in extra construction costs and large
amounts of fresh water, making it hard to apply in marine construction [5–7]. Scholars have developed one possible way to replace
steel in concrete by applying fiber-reinforced polymer (FRP) [8–14]. FRP materials are widely used in engineering construction owing
to their high strength-weight ratio and excellent corrosion resistance. Research has proved that chloride ions do not have accelerated
corrosion effects on the durability of FRP [15–17]. The hybrid structure of FRP-SWSSC is regarded as a feasible composite in marine
engineering. Some literature pointed out that compared with other types of FRP materials, GFRP showed many benefits, such as its
lower cost and good ductility [18,19].
With the development of engineering technology, there are more and more deep beam problems that should consider the effect of
shear deformation, such as the case of large structural cross-section size relative to the span, higher-order vibration of beams, local
height bearing, and foundation settlement of beams with elastic foundations, etc. Applying the primary beam theory analysis will lead
to problems with small calculated deflections and insufficient calculation accuracy. In order to solve these problems, deep beams are
necessary for further research. In the meantime, the shear capacity of seawater sea-sand engineered cementitious composite beams
without stirrups is improved by 6.37%–73.68% when compared with seawater sea-sand concrete beams with stirrups [20]. Accord­
ingly, FRP bars reinforced SWSS beams may possess satisfactory cost-effectiveness. Unfortunately, there is limited research on the

2
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142

GFRP-SWSS beams under shear loading. Reinforced concrete (RC) beams may fail in shear due to poor design and incorrect con­
struction. Shear damage is known to be more brittle than bending damage. Concrete in the shear span of RC beams is in the
tension-shear (point A), pure shear (point B) and compression-shear (point C) stress state, as displayed in Fig. 1[21]. The
compression-shear characteristics of seawater sea-sand engineered cementitious composites with various fibre volume fractions were
studied in the literature [22]. Additionally, Liao et al. [23] explored the tension-shear characteristics of seawater sea-sand engineered
cementitious composites with various water/binder ratios. The above results indicate that seawater sea sand-engineered cementitious
composites have higher breaking strength and ductility than mortar due to fibre bridging action. Engineered cementitious composite
beams reinforced with GFRP bars show higher shear capacity than high-strength concrete beams [24]. As a result, it is necessary to
investigate the shear performance of GFRP-SWSSC deep beams.
Considerable studies on the mechanical properties of SWSSC have been conducted. Zhang et al. [18] pointed out that sea sand’s
seashell contents were generally higher than in river sand. It should be noted that the seashell content in different sea areas is different.
Wegian et al. [25] reported an investigation on the effect of seawater used in concrete on the compressive strength and flexural
strength of SWSSC. It was concluded that the axial compressive strength and other seawater sea-sand concrete were increased at early
14 days compared to ordinary concrete, and the strength decreased with the increase of exposure time. At the same time, some re­
searchers found that the strength of SWSSC was increased with the ascending exposure time [26,27]. Through experimental studies,
Yang et al. [28] explored the effect of seashell content ratio on the long-term mechanical properties of concrete. The results indicated
that when seashell was used to replace 10% fine aggregate, the long-term strength of the concrete was similar to that of ordinary
concrete. However, the long-term strength was lower when the replacement ratio was up to 20%. Qu et al. [29] studied the effects of
glass fibre and blast-furnace slag on the mechanical properties of seawater sea-sand cementitious mortar. Experimental results showed
that the reaction between chloride ions in seawater, sea sand, and cement particles enhanced the cementitious mortar’s compressive
strength and that adding glass fibre could improve flexural strength. Li et al. [30] investigated the effects of seawater and distilled
water on the mechanical properties of concrete and found that seawater could accelerate the hydration of ordinary portland cement
(OPC), resulting in a remarkable increase in compressive strength at the early stage. The authors stated that using seawater in concrete
could produce prominent benefits to the economy and the environment. Xing et al. [31] carried out an experimental study on the axial
compressive properties of pultruded GFRP columns. In addition, some researchers [32–38] investigated the behaviour of SWSSC
columns, and FRP-reinforced concrete columns, respectively. Zhou et al. [39] and Yang et al. [40] experimentally investigated the
SWSSC columns externally wrapped with CFRP and proposed a design model to predict the ultimate strength of CFRP-reinforced
SWSSC columns. Wei et al. [41] proposed a hybrid column composed of wrapped FRP, steel tube and filled SWSSC and developed
a general model for evaluating the ultimate strength of the hybrid column. It was concluded from the comparison that FRP-confined
columns showed better deformation capacity than carbon-FRP, and FRP layers have a more significant effect on the ultimate strength
of the columns than steel tube thickness. Yang et al. [42] further developed a stress-strain model to accurately predict the axial
behaviour of CFRP partially confined the SWSSC columns under the base of existing research and found that the proposed model could
be applied in not only CFRP fully confined SWSSC columns but also partially FRP confined columns.
Otherwise, there is some literature about FRP-reinforced SWSSC beams’ behaviour [43–45]. Zhang et al. [43] conducted experi­
mental investigations on forty-eight pultruded GFRP tubes filled with SWSSC to explore the effects of cross-section size and GFRP
thickness on the flexural behaviour of beams. Dong et al. [44,45] conducted a relevant study on the shear behaviour of
BFRP-reinforced SWSSC beams. Results showed that reinforcements could effectively enhance the shear capacity of the hybrid beams.
The authors also proposed a new hybrid beam type consisting of an SWSSC beam, an FRP bar, and a UHPC shell. Experimental results
indicated that the cracking load of the hybrid beam was improved by approximately 70%.
Very limited studies on the shear behaviour of GFRP-SWSSC composites were conducted, let alone studies on GFRP-SWSSC deep
beams with seashells inside. This paper investigates the feasibility of combining SWSSC reinforced with GFRP bar as an alternative to
reinforced concrete structures. The effects of five seashell content, section heights, and GFRP bar diameter on the shear performance of
GFRP-SWSSC deep beams are analyzed. Three existing guidelines for GFRP-reinforced concrete beams are utilized to evaluate the
shear strength of GFRP-SWSSC deep beams. Design approaches are proposed based on the Strut-and-tie model and compared with
experiments. The corresponding FEM of GFRP-SWSSC deep beams were established to simulate the shear behaviour of GFRP-SWSSC
specimens.

Fig. 1. Stress states of concrete material in beam [21].

3
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142

Table 1
Details of test specimens.
Specimens H/mm b/mm D/mm As/mm2 h0/mm SC Shear-span ratio Reinforcement ratio /% Concrete categories

SH100-D8–1 100 100 8 100.48 80 0.00% 1.88 1.00% SWSSC


SH100-D8–2 100 100 8 100.48 80 3.57% 1.88 1.00% SWSSC
SH100-D8–3 100 100 8 100.48 80 7.14% 1.88 1.00% SWSSC
SH100-D8–4 100 100 8 100.48 80 10.71% 1.88 1.00% SWSSC
SH100-D8–5 100 100 8 100.48 80 14.29% 1.88 1.00% SWSSC
SH100-D10–1 100 100 10 157.00 80 0.00% 1.88 1.57% SWSSC
SH100-D10–2 100 100 10 157.00 80 3.57% 1.88 1.57% SWSSC
SH100-D10–3 100 100 10 157.00 80 7.14% 1.88 1.57% SWSSC
SH100-D10–4 100 100 10 157.00 80 10.71% 1.88 1.57% SWSSC
SH100-D10–5 100 100 10 157.00 80 14.29% 1.88 1.57% SWSSC
SH100-D12–1 100 100 12 226.08 80 0.00% 1.88 2.26% SWSSC
SH100-D12–2 100 100 12 226.08 80 3.57% 1.88 2.26% SWSSC
SH100-D12–3 100 100 12 226.08 80 7.14% 1.88 2.26% SWSSC
SH100-D12–4 100 100 12 226.08 80 10.71% 1.88 2.26% SWSSC
SH100-D12–5 100 100 12 226.08 80 14.29% 1.88 2.26% SWSSC
SH100-D14–1 100 100 14 307.72 80 0.00% 1.88 3.08% SWSSC
SH100-D14–2 100 100 14 307.72 80 3.57% 1.88 3.08% SWSSC
SH100-D14–3 100 100 14 307.72 80 7.14% 1.88 3.08% SWSSC
SH100-D14–4 100 100 14 307.72 80 10.71% 1.88 3.08% SWSSC
SH100-D14–5 100 100 14 307.72 80 14.29% 1.88 3.08% SWSSC
SH125-D8–1 125 100 8 100.48 105 0.00% 1.43 0.80% SWSSC
SH125-D8–2 125 100 8 100.48 105 3.57% 1.43 0.80% SWSSC
SH125-D8–3 125 100 8 100.48 105 7.14% 1.43 0.80% SWSSC
SH125-D8–4 125 100 8 100.48 105 10.71% 1.43 0.80% SWSSC
SH125-D8–5 125 100 8 157.00 105 14.29% 1.43 0.80% SWSSC
SH125-D10–1 125 100 10 157.00 105 0.00% 1.43 1.26% SWSSC
SH125-D10–2 125 100 10 157.00 105 3.57% 1.43 1.26% SWSSC
SH125-D10–3 125 100 10 157.00 105 7.14% 1.43 1.26% SWSSC
SH125-D10–4 125 100 10 157.00 105 10.71% 1.43 1.26% SWSSC
SH125-D10–5 125 100 10 157.00 105 14.29% 1.43 1.26% SWSSC
SH125-D12–1 125 100 12 226.08 105 0.00% 1.43 1.81% SWSSC
SH125-D12–2 125 100 12 226.08 105 3.57% 1.43 1.81% SWSSC
SH125-D12–3 125 100 12 226.08 105 7.14% 1.43 1.81% SWSSC
SH125-D12–4 125 100 12 226.08 105 10.71% 1.43 1.81% SWSSC
SH125-D12–5 125 100 12 226.08 105 14.29% 1.43 1.81% SWSSC
SH125-D14–1 125 100 14 307.72 105 0.00% 1.43 2.46% SWSSC
SH125-D14–2 125 100 14 307.72 105 3.57% 1.43 2.46% SWSSC
SH125-D14–3 125 100 14 307.72 105 7.14% 1.43 2.46% SWSSC
SH125-D14–4 125 100 14 307.72 105 10.71% 1.43 2.46% SWSSC
SH125-D14–5 125 100 14 307.72 105 14.29% 1.43 2.46% SWSSC
SH150-D8–1 150 100 8 100.48 130 0.00% 1.15 0.67% SWSSC
SH150-D8–2 150 100 8 100.48 130 3.57% 1.15 0.67% SWSSC
SH150-D8–3 150 100 8 100.48 130 7.14% 1.15 0.67% SWSSC
SH150-D8–4 150 100 8 100.48 130 10.71% 1.15 0.67% SWSSC
SH150-D8–5 150 100 8 100.48 130 14.29% 1.15 0.67% SWSSC
SH150-D10–1 150 100 10 157.00 130 0.00% 1.15 1.05% SWSSC
SH150-D10–2 150 100 10 157.00 130 3.57% 1.15 1.05% SWSSC
SH150-D10–3 150 100 10 157.00 130 7.14% 1.15 1.05% SWSSC
SH150-D10–4 150 100 10 157.00 130 10.71% 1.15 1.05% SWSSC
SH150-D10–5 150 100 10 157.00 130 14.29% 1.15 1.05% SWSSC
SH150-D12–1 150 100 12 226.08 130 0.00% 1.15 1.51% SWSSC
SH150-D12–2 150 100 12 226.08 130 3.57% 1.15 1.51% SWSSC
SH150-D12–3 150 100 12 226.08 130 7.14% 1.15 1.51% SWSSC
SH150-D12–4 150 100 12 226.08 130 10.71% 1.15 1.51% SWSSC
SH150-D12–5 150 100 12 226.08 130 14.29% 1.15 1.51% SWSSC
SH150-D14–1 150 100 14 307.72 130 0.00% 1.15 2.05% SWSSC
SH150-D14–2 150 100 14 307.72 130 3.57% 1.15 2.05% SWSSC
SH150-D14–3 150 100 14 307.72 130 7.14% 1.15 2.05% SWSSC
SH150-D14–4 150 100 14 307.72 130 10.71% 1.15 2.05% SWSSC
SH150-D14–5 150 100 14 307.72 130 14.29% 1.15 2.05% SWSSC
RH100-D10–1 100 100 10 157.00 80 0.00% 1.88 1.57% RWRSC
RH100-D10–2 100 100 10 157.00 80 3.57% 1.88 1.57% RWRSC
RH100-D10–3 100 100 10 157.00 80 7.14% 1.88 1.57% RWRSC
RH100-D10–4 100 100 10 157.00 80 10.71% 1.88 1.57% RWRSC
RH100-D10–5 100 100 10 157.00 80 14.29% 1.88 1.57% RWRSC
RH150-D10–1 150 100 10 157.00 130 0.00% 1.15 1.05% RWRSC
RH150-D10–2 150 100 10 157.00 130 3.57% 1.15 1.05% RWRSC
RH150-D10–3 150 100 10 157.00 130 7.14% 1.15 1.05% RWRSC
RH150-D10–4 150 100 10 157.00 130 10.71% 1.15 1.05% RWRSC
RH150-D10–5 150 100 10 157.00 130 14.29% 1.15 1.05% RWRSC

4
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142

2. Experimental programs

The four-point loading test programs contained two categories depending on the type of concrete used. The two categories were
identified: S series and R series. The S series was comprised of sixty GFRP-SWSSC beams. Moreover, the R series consisted of ten river-
water river-sand concrete (RWRSC) beams reinforced with GFRP bars as control groups. Seventy concrete beams reinforced with GFRP
bars were loaded until failure.

2.1. Test specimens

The experiments were conducted on seventy specimens with the same width (b) of 100 mm, a longitudinal length (L) of 500 mm,
effective span length (Le) of 400 mm, and thickness of the protective layer of concrete (C) is designed to be 20 mm. The cross-section
dimensions (b × H × L) of GFRP-reinforced SWSSC specimens are provided in Table 1. The shear-span ratio is the ratio of shear-span to
effective section height (h0). All specimens were comprised of concrete and GFRP reinforcement, as shown in Fig. 2. Two GFRP bars
with the same diameter were arranged in a tension zone to explore the shear response of SWSSC specimens under four-point bending.
Since GFRP bars are not easy to bend, the bottom GFRP bars in the test beam are thickened at the end for equivalent anchorage. Five
seashell contents (0%, 3.57%, 7.14%, 10.71% and 14.29%), labelled as SC in the tables, were set to investigate seashell content’s effect
on sea-sand’s structural behaviour of SWSSC beams. All specimens were cured under the same condition, thereby reducing experi­
mental variation.
The nomenclature of all specimens is related to the fabrication of specimens. For instance, the nomenclature "SH100-D8–1" is
defined as follows.

● The first notation, “S” denotes that the concrete used is seawater sea-sand concrete. If the first expression is “R”, it denotes the
concrete used is ordinary concrete comprising river-water river-sand concrete.
● The second expression, “H100”, indicates the cross-section height is 100 mm. If the expression is “H125”, “125” denotes that the
cross-section height is 125 mm. If the expression is “H150”, “150” denotes that the cross-section height is 150 mm.
● The following expression, “D8”, indicates that the diameter of longitudinal GFRP reinforcement is 8 mm. If the expression is “D10”,
“10” denotes that longitudinal GFRP reinforcement is 10 mm. If the expression is “D12”, “12” denotes that longitudinal GFRP
reinforcement is 12 mm. If the expression is “D14”, “14”denotes that longitudinal GFRP reinforcement is 14 mm.
● The last part, “1” means no sea seashell content in concrete. If the part is “2”, it denotes that the seashell content in the sea-sand is
3.57%. If the part is “3”, it denotes that the seashell content in the sea-sand is 7.14%. If the part is “4”, it denotes the seashell content
in the sea-sand is 10.71%. If the part is “5”, it denotes the seashell content in the sea-sand is 14.29%.

2.2. Materials

The designed concrete with a compressive strength of 30 MPa was chosen for GFRP bar-reinforced RC specimens. The S series
comprised seawater, sea-sand, cement, seashells, fly ash and coarse aggregate. The sea-sand provided by Fujian manufacturer was used
after screening. The R series contained river water, sand, cement, fly ash and coarse aggregate. Tables 2–3 show five mixed proportions
of SWSSC and RWRSC used to fabricate beams. Fifteen cylinder cubes of seawater sea-sand concrete with five seashell ratio and fifteen
cylinder cubes of river water river-sand concrete were cured under the same condition [46]. The compressive test results are shown in
Table 4. The proposed seawater sea-sand concrete mix achieved a compressive strength of 32 MPa with desired workability in the
28-day term. The GFRP bars were chosen as longitudinal reinforcements of SWSSC beams. GFRP bars with four diameters, 8 mm
(Am1=0.5 cm2), 10 mm (Am2=0.73 cm2), 12 mm (Am3=1.03 cm2) and 14 mm (Am4=1.34 cm2), were used in this study. The manu­
facturer provided the materials with an ultimate tensile strength of 980 GPa, elastic modulus of 52 GPa, and Poisson’s ratio of 0.3.

Fig. 2. Geometry of GFRP reinforced SWSSC beams.

5
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142

Table 2
Mix proportion of SWSSC.
Concrete categories Mix of per cubic meter

Seawater /kg Cement /kg Sea-sand /kg Coarse aggregate /kg Seashell /kg Fly ash /kg

SWSSC-R0 175 296 765 1057 0 78


SWSSC-R1 175 296 738 1057 27 78
SWSSC-R2 175 296 710 1057 55 78
SWSSC-R3 175 296 683 1057 82 78
SWSSC-R4 175 296 656 1057 109 78

Table 3
Mix proportion of RWRSC.
Concrete categories Mix of per cubic meter

River-water /kg Cement /kg River-sand /kg Coarse Seashell /kg Fly ash /kg
aggregate /kg

RWRSC -R0 175 296 765 1057 0 78


RWRSC -R1 175 296 738 1057 27 78
RWRSC -R2 175 296 710 1057 55 78
RWRSC -R3 175 296 683 1057 82 78
RWRSC -R4 175 296 656 1057 109 78

Table 4
Test results of concrete cubes.
Concrete categories Test coupon Cubic compressive strength (MPa) Average value (MPa)

SWSSC-1 NO.1 35.15 33.87


NO.2 34.87
NO.3 31.59
SWSSC-2 NO.1 34.12 32.85
NO.2 33.01
NO.3 31.42
SWSSC-3 NO.1 30.09 32.05
NO.2 34.10
NO.3 31.96
SWSSC-4 NO.1 32.57 30.96
NO.2 30.29
NO.3 30.02
SWSSC-5 NO.1 28.11 29.35
NO.2 30.83
NO.3 29.12

Fig. 3. Testing instrumentation.

6
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142

2.3. Instrumentation and loading procedure

A MTS testing machine with 100 kN capacity is employed to perform a supported shear test on the behaviour of SWSSC beams
reinforced with GFRP bars, as demonstrated in Fig. 3(a). All the beams were supported, and the rotation and lateral deflection were not
restrained. The test beam was loaded monotonically at the rate of 0.2 mm/min until the failure of the test beam. The photo of testing
instrumentation is shown in Fig. 3(b). According to the observed loading process of specimens, the cracking load was recorded. Fig. 4
presents the arrangement of strain gauges. In order to study the strain variation of concrete on the side of the specimen during loading
and identify the cracking load value of the specimen, five strain gauges (T1-T5) were pasted at the corresponding position on the right
side of the tested beam. The values of strain gauges under different loads were measured by the DH3816 system.

3. Results and discussions

Experimental results of GFRP-SWSSC deep beams are presented in Table 5. Results are shown in terms of cracking strength(Vcr),
shear ultimate strength (Vu), stiffness (ki), ductility coefficient (DC), λ and obtained shear load-displacement curves. This paper ex­
presses the shear strength (V) of GFRP-SWSSC deep beams as follows: V=P/2. λ is refers to the ratio of GFRP-SWSSC with seashell deep
beams to GFRP-SWSSC deep beams without seashells. DC can be defined as DC=Δu/Δy, where Δu represents the displacement at Vu and
Δy represents the displacement at Vcr. The stiffness is ki=V0.8/Δ0.8, where Δ0.8 represents the displacement at V0.8.

3.1. Crack patterns and failure modes

The deep beams show higher capacity than the beams, which is attributed to its Strut-and-tie force system. According to the
observed loading process, the crack development process of GFRP-reinforced SWSSC beams and GFRP-reinforced RWRSC beams were
similar. The general cracking behaviour is described below, depending upon the observed failure process of GFRP-SWSSC specimens
shown in Fig. 5.
Based on the observed loading process, the crack development process of GFRP-reinforced SWSSC beams and GFRP-reinforced
RWRSC beams were similar. All specimens experienced three stages: cracking, critical fracture and ultimate failure. At the begin­
ning of loading, the specimen was in the elastic stage, and there was no crack on the concrete surface. The first flexural crack occurred
in the maximum moment region. The flexural crack developed vertically owing to the absence of shear stress. Several tiny vertical
cracks were formed at the shear span during the loading process. The formed crack extended towards the loading and supporting
points. At this time, it was considered that the specimen reached an inclined cracking state. As the load approached the shear ultimate
capacity of the tested beam, the diagonal crack penetrated rapidly towards the loading point until failure.
Moreover, The diagonal crack that resulted in the final sudden failure was defined as a critical diagonal crack. The phenomenon can
be explained by the fact that a diagonal crack reduced the effective height of the shear-compression area, resulting in concrete damage
in the shear-compression region and the failure of GFRP-reinforced SWSSC beams. According to the literature[47], such a failure mode
was defined as shear-compression failure (SCF). In addition, another cracking pattern was recorded during the loading process. The
primary difference was that a wider, deeper flexural crack extended and crossed the whole beam height. Subsequently, the diagonal
crack was formed rapidly towards the loading point.
The development of flexural crack resulted in premature shear failure and a large loss of shear strength. It should be noted that the
phenomenon occurred in tested beams reinforced with 14 mm GFRP bars. When the longitudinal GFRP reinforcement exceeded a
certain range, the failure mode of specimens changed to flexural-shear failure (FSF), and the crack at the mid-span was wider and
deeper. Such a premature failure may be due to the decrease of the effective shear area of the inclined section owing to the config­
uration of longitudinal reinforcement. On the other hand, the lower elastic modulus compared to the tensile strength of GFRP bars
caused the penetrating crack [48]. Interestingly, anchorage failure (FSWAF) was observed in some specimens with 14 mm GFRP bars.
The configuration of GFRP bars reduced the effective cross-section area of concrete, resulting in the aggregate interlock force around
the supporting points.
Fig. 6 provides part of the specimens after cracking. A few cracks occurred on the surface of the beams for specimens with a section

Fig. 4. Locations of strain gauges.

7
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142

Table 5
Test results of specimens.
Specimens H/mm D/mm SC Vcr/kN Vu/kN Δu/mm V0.8/kN DC Ki ξ Failure
modes

SH100-D8–1 100 8 0.00% 12.0 25.45 3.27 20.36 1.49 9.25 1.00 SCF
SH100-D8–2 100 8 3.57% 11.3 21.52 3.14 17.22 1.54 8.16 0.85 SCF
SH100-D8–3 100 8 7.14% 10.9 17.35 2.19 13.88 1.49 9.99 0.68 SCF
SH100-D8–4 100 8 10.71% 10.7 17.30 2.96 13.84 1.35 6.53 0.68 SCF
SH100-D8–5 100 8 14.29% 8.7 17.93 3.08 14.34 1.9 8.80 0.70 SCF
SH100-D10–1 100 10 0.00% 12.5 30.23 4.28 24.18 1.6 9.13 1.00 SCF
SH100-D10–2 100 10 3.57% 11.6 27.13 3.81 21.70 1.66 9.48 0.88 SCF
SH100-D10–3 100 10 7.14% 9.8 24.70 2.75 19.76 1.4 10.03 0.82 SCF
SH100-D10–4 100 10 10.71% 11.0 19.11 3.07 15.29 2.05 9.99 0.64 SCF
SH100-D10–5 100 10 14.29% 10.5 17.50 2.53 14.00 1.82 10.29 0.57 SCF
SH100-D12–1 100 12 0.00% 14.0 36.56 3.69 29.25 1.46 11.70 1.00 SCF
SH100-D12–2 100 12 3.57% 13.8 30.41 3.45 24.33 1.31 9.36 0.85 SCF
SH100-D12–3 100 12 7.14% 13.0 25.92 3.44 20.74 1.73 10.47 0.73 SCF
SH100-D12–4 100 12 10.71% 12.7 20.18 4.11 16.14 1.98 8.07 0.57 SCF
SH100-D12–5 100 12 14.29% 12.1 19.34 2.19 15.47 1.22 9.98 0.53 SCF
SH100-D14–1 100 14 0.00% 10.1 24.92 5.12 19.94 1.53 5.54 1.00 FSWAF
SH100-D14–2 100 14 3.57% 9.9 25.02 7.26 20.02 1.85 4.99 1.00 FSWAF
SH100-D14–3 100 14 7.14% 9.0 20.94 3.58 16.75 1.43 4.79 0.84 FSWAF
SH100-D14–4 100 14 10.71% 7.9 18.75 6.47 15.00 1.25 2.90 0.75 FSWAF
SH100-D14–5 100 14 14.29% 7.5 17.75 8.56 14.20 1.29 2.15 0.71 FSWAF
SH125-D8–1 125 8 0.00% 18.5 41.33 4.45 33.06 1.73 12.87 1.00 SCF
SH125-D8–2 125 8 3.57% 17.7 40.15 4.61 32.12 1.48 10.36 0.99 FSF
SH125-D8–3 125 8 7.14% 16.3 34.28 4.45 27.42 1.32 9.05 0.83 SCF
SH125-D8–4 125 8 10.71% 14.3 27.56 3.41 22.05 1.4 9.42 0.64 SCF
SH125-D8–5 125 8 14.29% 14.1 28.56 3.87 22.85 1.53 9.68 0.70 SCF
SH125-D10–1 125 10 0.00% 23.3 46.23 3.85 36.98 1.26 12.05 1.00 SCF
SH125-D10–2 125 10 3.57% 21.4 40.66 4.15 32.53 1.49 11.74 0.87 SCF
SH125-D10–3 125 10 7.14% 18.6 36.84 3.42 29.47 1.5 12.87 0.78 SCF
SH125-D10–4 125 10 10.71% 18.9 31.46 2.79 25.17 1.24 11.24 0.67 SCF
SH125-D10–5 125 10 14.29% 17.0 30.36 3.41 24.29 1.44 10.38 0.65 SCF
SH125-D12–1 125 12 0.00% 23.6 50.91 3.96 40.73 1.35 14.49 1.00 SCF
SH125-D12–2 125 12 3.57% 20.4 45.13 3.02 36.10 1.34 16.05 0.91 SCF
SH125-D12–3 125 12 7.14% 20.0 41.41 3.73 33.13 1.38 14.10 0.80 SCF
SH125-D12–4 125 12 10.71% 19.2 35.63 3.35 28.50 1.38 11.93 0.76 SCF
SH125-D12–5 125 12 14.29% 18.7 34.89 3.23 27.91 1.29 11.88 0.75 SCF
SH125-D14–1 125 14 0.00% 15.5 29.53 3.13 23.62 1.26 9.19 1.00 FSF
SH125-D14–2 125 14 3.57% 14.7 29.49 6.19 23.59 1.99 6.98 1.00 FSF
SH125-D14–3 125 14 7.14% 13.0 23.52 5.71 18.82 1.35 4.46 0.80 FSWAF
SH125-D14–4 125 14 10.71% 11.8 20.35 3.01 16.28 1.19 6.43 0.69 FSWAF
SH125-D14–5 125 14 14.29% 9.7 21.39 5.87 17.11 1.67 4.75 0.72 FSWAF
SH150-D8–1 150 8 0.00% 29.0 52.08 3.38 41.66 1.42 17.65 1.00 SCF
SH150-D8–2 150 8 3.57% 27.3 46.53 3.34 37.22 1.22 14.95 0.89 SCF
SH150-D8–3 150 8 7.14% 26.1 42.21 3.36 33.77 1.58 14.62 0.84 SCF
SH150-D8–4 150 8 10.71% 24.6 37.98 3.09 30.38 1.27 14.07 0.77 SCF
SH150-D8–5 150 8 14.29% 20.3 35.49 3.41 28.39 1.31 14.64 0.68 SCF
SH150-D10–1 150 10 0.00% 31.8 57.49 2.87 45.99 1.21 19.32 1.00 SCF
SH150-D10–2 150 10 3.57% 27.0 53.92 3.42 43.14 1.43 18.28 0.94 SCF
SH150-D10–3 150 10 7.14% 27.9 49.11 3.19 39.29 1.4 17.38 0.85 SCF
SH150-D10–4 150 10 10.71% 28.3 46.14 3.26 36.91 1.18 15.07 0.81 SCF
SH150-D10–5 150 10 14.29% 28.0 40.00 3.12 32.00 1.28 14.75 0.83 SCF
SH150-D12–1 150 12 0.00% 36.9 59.75 2.49 47.80 1.26 24.26 1.00 SCF
SH150-D12–2 150 12 3.57% 34.2 59.13 4.03 47.30 1.37 16.26 0.99 SCF
SH150-D12–3 150 12 7.14% 32.7 56.68 2.87 45.34 1.34 21.09 0.96 FSF
SH150-D12–4 150 12 10.71% 31.5 50.98 2.6 40.78 1.5 21.47 0.96 SCF
SH150-D12–5 150 12 14.29% 31.0 49.36 3.04 39.49 1.32 17.09 0.84 SCF
SH150-D14–1 150 14 0.00% 18.7 45.65 3.91 36.52 1.24 10.90 1.00 FSF
SH150-D14–2 150 14 3.57% 21.0 42.19 14.01 33.75 3.24 7.50 1.16 FSF
SH150-D14–3 150 14 7.14% 20.9 40.87 7.99 32.70 1.82 8.60 1.12 FSWAF
SH150-D14–4 150 14 10.71% 18.4 37.35 5.06 29.88 1.33 6.57 1.02 FSF
SH150-D14–5 150 14 14.29% 18.0 38.20 4.29 30.56 1.37 6.02 1.05 FSWAF
RH100-D10–1 100 10 0.00% 12.8 30.82 4.28 24.66 1.37 8.01 1.00 SCF
RH100-D10–2 100 10 3.57% 11.0 27.81 3.82 22.25 1.41 8.12 0.92 SCF
RH100-D10–3 100 10 7.14% 11.3 25.19 2.77 20.15 1.39 9.42 0.82 SCF
RH100-D10–4 100 10 10.71% 10.6 20.35 2.65 16.28 1.66 9.58 0.66 SCF
RH100-D10–5 100 10 14.29% 9.8 19.98 2.81 15.98 1.35 7.80 0.65 SCF
RH150-D10–1 150 10 0.00% 32.6 56.63 3.32 45.30 1.27 16.90 1.00 SCF
RH150-D10–2 150 10 3.57% 31.0 55.73 3.47 44.58 1.32 17.21 0.98 SCF
(continued on next page)

8
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142

Table 5 (continued )
Specimens H/mm D/mm SC Vcr/kN Vu/kN Δu/mm V0.8/kN DC Ki ξ Failure
modes

RH150-D10–3 150 10 7.14% 29.9 48.22 2.64 38.58 1.28 18.64 0.85 SCF
RH150-D10–4 150 10 10.71% 23.1 42.95 2.67 34.36 1.32 17.35 0.76 SCF
RH150-D10–5 150 10 14.29% 29.0 40.58 3.31 32.46 1.34 13.25 0.72 SCF

Fig. 5. Typical crack patterns.

Fig. 6. Part of specimens after cracking.

Fig. 7. Comparison of SWSSC and RWRSC specimen load deflection curves.

9
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142

height of 150 mm. However, many tiny cracks appeared on the surface of specimens with a section height of 100 mm, resulting in the
beams’ premature failure. The cracks in the surface of the beams reduced the effective compressive concrete area under a four-point
load. Furthermore, the specimens with lower seashell content exhibited fewer cracks. In comparison, the specimens with higher
seashell content showed wider cracks accompanied by obvious brittle failure.

3.2. Shear load-displacement curves

Fig. 7 compares the shear load-deflection curves of the hoopless test beams of river water river sand and seawater sea-sand con­
crete. The vertical axis represents the applied shear force (V), and the transverse axis represents the displacement (Δ). The shear load-
deflection curves of the GFRP-reinforced SWSSC deep beams are not significantly different from those of the GFRP-reinforced RWRSC
beams. This is because the differences in the mechanical properties of the SWSSC and the RWRSC are relatively small, and the cubic
compressive strengths and elastic modulus of the two are similar. Hence, the two shear load-deflection curves are the same. This proves
the feasibility of replacing river water river sand concrete (RWRSC) with seawater sea-sand concrete (SWSSC) for GFRP-reinforced
concrete beams.
Three groups of shear load-displacement curves of GFRP-reinforced SWSSC beams without stirrup under design parameters are
presented in Fig. 8. The curves of specimens exhibited almost linear behaviour until failure. It should be noted that the load-
displacement curves had no obvious elastoplastic stage before the specimens failed, indicating that the failure modes of specimens
were brittle. The single linear stress-strain curve and the large tensile strength of the GFRP bar could explain it. Thus, the plastic yield
stage did not appear in GFRP-reinforced SWSSC beams. The crack continued to expand until the beam failed with obvious brittleness. It
could be found from Fig. 8(a) that the stiffness increased with the increase of section height. When other conditions remained constant,
only increasing the section height would increase the angle of the diagonal crack. This was attributed to the contribution of the
principal tensile stress in the shear critical region of SWSSC beams. In addition, it could be found from Fig. 8(b) that the stiffness of
specimens with different GFRP longitudinal reinforcement ratios was the same before cracking.
GFRP-reinforced SWSSC beams showed better stiffness within a certain reinforcement ratio degree with increased GFRP bar
diameter. Meanwhile, the specimens reinforced with 14 mm GFRP bars exhibited the lowest stiffness. For a certain protective layer
thickness, the relative protective layer thickness (C/D) of concrete decreases as the diameter of GFRP reinforcement increases, and the
bond between concrete and longitudinal reinforcement is weakened. Configuration of 14 mm diameter GFRP bar test beam, before the
specimen cracking, by the concrete and longitudinal reinforcement to bear the vertical load; when the specimen cracking, the lon­
gitudinal bond between the concrete and longitudinal bar slippage occurs, resulting in a wide and long crack near the middle of the

Fig. 8. Shear load-displacement curves of specimens.

10
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142

span, most of the concrete out of work, at this time, the vertical load is mainly borne by the longitudinal bar of the GFRP, longitudinal
reinforcement only bear the bending moment, does not bear the shear force. Due to the relatively small modulus of elasticity of GFRP
reinforcement, the flexural stiffness of the member is reduced, so the deflection of the test beam suddenly increases after cracking.
When the deflection reaches a certain limit, the concrete starts to bear the shear force. When the load value reaches the critical crack of
the member, the member finally undergoes bond-shear damage because the shear force in the shear span area exceeds the shear

Fig. 9. Vertical strain distribution curves at the midspan.

11
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142

Fig. 10. Effects of parameters on the cracking strength.

12
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142

strength of the concrete. The critical crack load of the member is almost the same as the ultimate load value. Therefore, when
increasing the diameter of GFRP longitudinal reinforcement, the relative protective layer thickness should be considered. When the
relative protective layer thickness (C/D) is <1.5, the specimen will not undergo shear damage but may undergo bond-shear
compression damage, which is more brittle and obvious. Furthermore, it could be observed from Fig. 8(c) that the stiffness of spec­
imens decreased with the ascending of seashell content in general. This was because the addition of seashells in sea-sand led to the
formation and expansion of cracks.

3.3. Strain distribution curves at the mid-span

Fig. 9 provides the shear load-strain curves at the mid-span of GFRP-SWSSC deep beams. Under the same loading, the strains of the
GFRP-SWSSC specimens were the same as those of the R series. It revealed that the effect of SWSSC on hybrid GFRP-SWSSC deep beams
could be neglected. According to Fig. 9, it could be found that the zero strain value was obtained from 3/4 h to 5/6 h of cross-section
position. This observation indicated that the neutral axis of the cross-section of GFRP-reinforced SWSSC beams was removed upward
since the concrete in the tensile area was cracked. The neutral axis moved upward for specimens with a section height of 150 mm
compared to specimens with a 100 mm section height. Moreover, the neutral axis moved downward for specimens reinforced with
12 mm GFRP bars, compared to specimens with 10 mm GFRP bars. The observation demonstrated that the shear strength could be
enhanced by increasing section height and reducing the longitudinal reinforcement ratio of beams. In addition, the strain curves
remained linear before the applied load reached the cracking load. However, the strain developed non-uniformly after the concrete in
the tensile region cracked, indicating that the plane-section assumption was ineffective after the concrete cracked.

Fig. 11. Effect of section-height on shear strength of specimens.

13
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142
Fig. 12. Effect of GFRP bar diameter on shear strength of specimens.
14
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142

3.4. Effect of section height

The shear load-displacement curves of tested beams with different section heights were plotted in Fig. 8(a). It could be seen from
Fig. 8(a) and Table 5 that the stiffness of tested GFRP-reinforced SWSSC beams was remarkably enhanced with the increase of section
height. Moreover, the beams with a section height of 150 mm were roughly 87% higher ki than those of specimens with a section
height of 100 mm. This was because when the shear-span length remained unchanged, the test beam’s shear-span ratio decreased with
the effective section height increase. The large improvement can be attributed to the reduced shear-span ratio’s contribution to the
specimens’ enhanced shear resistance. The enhanced stiffness also indicated that the effect of the shear-span ratio on the shear strength
of GFRP-reinforced SWSSC beams was significant.
Fig. 10(a) shows a cracking load plot against different section heights. According to Table 5 and the figure, increasing the section
height of tested beams could significantly enhance the specimen’s cracking load. The effect of aggregate interlock increased with the
effective section area of tested specimens. On the other hand, the stiffness of the tested beam was improved with the increase of the
effective height, which indirectly enhanced the crack resistance of the beam, resulting in the improvement of the cracking strength of
GFRP-SWSSC deep beams.
Fig. 11 shows the effect of different section heights of tested beams on the ultimate strength of the shear of the specimens. As
expected, increasing section height considerably enhanced the ultimate strength of the shear. Specifically, the shear ultimate strength
of specimens with 150 mm section height remarkably increased by approximately 105% compared to 100 mm section height. Such a
large enhancement was attributed to the effect of section height on the arch action of GFRP-SWSSC deep beams. The shear strength for
tested beams with a smaller shear-span ratio was mainly provided by the arch action, which was significantly affected by section height
[48].
Moreover, the shear ultimate strength of specimens with 150 mm section height increased by approximately 37% compared to that

Fig. 13. Effect of shell content on cracking strength and shear strength.

15
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142

of 125 mm section height, while the comparison between 100 mm and 125 mm section height resulted in a 49% difference in terms of
shear ultimate strength. The comparison implied that the effect of section height on the arch action of GFRP-reinforced SWSSC deep
beams without stirrups reinforcement was significant. Furthermore, as the GFRP bar diameter increased, section-height’s effect on
shear ultimate strength was more considerable. When specimens were reinforced with an 8 mm GFRP bar, the comparison of shear
ultimate capacity between 125 mm and 150 mm section height resulted in a 26% difference, while the difference increased to 35% and
36% for specimens reinforced with 10 mm and 12 mm GFRP bar, respectively. These observations implied that deep beams with
higher section height had better shear resistance and deformation capacity for GFRP-reinforced SWSSC structures. As the section
height increased, the section height enhancement on shear resistance was significant.

3.5. Effect of GFRP bar diameter

Fig. 10(b) provides the effect of cracking load versus GFRP bar diameter curves for GFRP-reinforced SWSSC beams without stirrup.
It can be seen from Fig. 10(b) and Table 5 that the influence of GFRP bar diameter on the cracking load of specimens is obvious. Tested
beams with 12 mm GFRP bars possessed the best cracking load and deformation capacity. The beams with 12 mm GFRP bars showed
roughly 27% higher cracking strength than those of specimens with 8 mm. This enhancement could be attributed to the fact that with
the increase of longitudinal reinforcement ratio, the height of the diagonal crack in the specimen is better controlled, which enhances
the shear resistance due to the contribution of the effective concrete area. However, the beams with 14 mm GFRP bars showed about
38% lower cracking strength than those of specimens with 12 mm. In particular, the 14 mm GFRP bar specimens showed better
ductility.
Fig. 12 shows the effect of GFRP bar diameter on the shear ultimate strength of specimens. Increasing GFRP bar diameter represents
the ascending longitudinal reinforcement ratio. The shear ultimate strength of specimens with a 12 mm GFRP bar remarkably
increased by approximately 27% compared to that of 8 mm, while the shear ultimate strength of specimens with a 14 mm GFRP bar
was 31% lower than that of 12 mm. This observation demonstrated that an increase of longitudinal GFRP reinforcement ratio up to
17% resulted in a 36% reduction of shear strength compared to a 14% longitudinal GFRP reinforcement ratio. The reduced effective
cross-section area resulted in vertical cracks around the loading points, thereby reducing the strength of GFRP-SWSSC deep beams.
Within a certain reinforcement range, the shear resistance of tested specimens increased with the longitudinal reinforcement ratio.
When the ratio is up to 2.05%, the specimens’ stiffness obviously decreased after the concrete was cracked. This is because the elastic
modulus of GFRP reinforcement is relatively small. With the increase in reinforcement ratio, the configuration of the GFRP bar reduced
the effective area. A wider and longer crack first developed at mid-span, resulting in an obvious reduction of shear strength of GFRP-
SWSSC deep beams. Further, some researchers demonstrated that the bonding strength descended with the ascending GFRP diameter
[49]. Further studies on the effect of bonding behaviour on the shear strength of GFRP-reinforced SWSSC beams are suggested.

3.6. Effect of seashell content

Fig. 13 reflects the pattern of the effect of shell content on the cracking load and ultimate shear capacity of GFRP-reinforced
seawater sea-sand concrete deep beams without hoops. From Fig. 13 and Table 5, it can be seen that for the overall trend, the
cracking load decreases with the increase of shell content due to the presence of shells altering the interfacial mechanical properties of
the concrete and decreasing the bond and tensile strength of the concrete.
At the same time, the ultimate shear strength of specimens generally decreases with the ascending seashell content. The shear
ultimate strength of specimens with 14.29% seashell content was approximately 26% lower than the corresponding specimens. This
can be explained by the fact that seashells in concrete reduced the fluidity of concrete, resulting in uneven distribution in the pouring
process. In addition, it could be found that as the seashell content (SC) was increased, the effect of seashell content on the ultimate
shear strength of tested beams was diminished. For example, as the specimen had 3.57% seashell content, the shear ultimate strength
decreased by 10% compared to the 7.14% seashell content. At the same time, The decrease was reduced to 4%, while seashell content
ranged from 10.71% to 14.29%. It could be noticed that when seashell contentment reached 10.71%, the effect of increasing seashell
content on shear resistance was limited for specimens with section heights of 100 mm and 125 mm.
The change in the stiffness of the GFRP-SWSSC deep beams is not very significant with a small difference in shell content. However,
as the shell content continues to increase, the stiffness of the GFRP-SWSSC deep beams begins to decrease significantly and their
maximum deflection decreases. This is due to the fact that the pore structure in the concrete changes due to the increase in shell
content, and the presence of shells changes the interfacial mechanical properties of the concrete and reduces the bond properties,
modulus of elasticity, and tensile strength of the concrete.There is no obvious yield plateau for the GFRP-SWSSC deep beams, and there
is an overall decreasing trend in the ultimate bearing capacity. The shells in the sea sand adversely affected the shear capacity of the
concrete beams, and the concrete at the support of the GFRP-SWSSC deep beams with large shell content was more likely to be crushed
and dislodged.The ultimate shear strength of the GFRP-SWSSC deep beams decreased with the increase of shell content, and the ul­
timate shear strength of the GFRP-SWSSC deep beams with 14.29% shell content was about 1.5 times lower than that of the shell-free
GFRP-SWSSC deep beams. SWSSC deep beams by about 29.86%. The reason for this is that the shells resemble needle-like particles,
which are not conducive to the fluidity of the concrete mixture, resulting in a lack of uniform distribution during the pouring process,
altering the pore structure of the concrete, and the shells are more brittle than the sea sand. Shells increase the porosity of concrete,
thus reducing the cracking load and ultimate shear capacity of concrete beams with diagonal cross-section. The effect of shell content
on the ultimate shear strength of GFRP-SWSSC deep beams decreases with the increase of shell content in sea sand. When the shell
content of GFRP-SWSSC deep beams ranged from 3.57% to 10.71%, the decrease of shear ultimate strength was 8–12%, and when the

16
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142

shell content of GFRP-SWSSC deep beams was 14.29%, the shear ultimate strength was only 3% lower than that of GFRP-SWSSC deep
beams with shell content of 10.71%. Moreover, when the shell content reaches 10.71%, the effect of increasing shell content on the
shear strength is limited for the GFRP-SWSSC deep beams with cross-section heights of 100 mm and 125 mm.

4. Design formula

Shear performance is an important part of GFRP-reinforced SWSSC beams. Compared with RC beams, the shear performance of
GFRP reinforced concrete differs from that of RC beams due to the low elastic modulus and axial stiffness of GFRP. The conventional
shear guidelines for RC deep beams are unavailable for GFRP-reinforced concrete deep beams. However, no design guideline for
evaluating the shear capacity of GFRP-reinforced SWSSC deep beams is proposed, let alone SWSSC beams with seashells. It should be
noted that SWSSC and RWRSC have similar mechanical axial performance with a 28-day. Thus, three current shear guidelines for FRP-
reinforced concrete are utilized to calculate the shear capacity of GFRP-reinforced SWSSC concrete deep beams. Also, the tension-
compression model calculates the shear strength of GFRP-SWSSC deep beams. The shear capacity formula for GFRP-SWSSC deep
beams without stirrups is proposed considering the seashell strength-reduction coefficient.

4.1. Existing guidelines for shear strength of GFRP-SWSSC deep beams without stirrups

It should be noted that three shear guidelines for the evaluation of GFRP-reinforced SWSSC concrete deep beams are based on the
assumption that the maximum stress in the concrete in the tensile zone can reach the tensile strength of concrete at a certain position in
the beam shear span. American guideline (ACI 440.1R-06) [50] mainly considers the effect of concrete strength and axial stiffness of
GFRP longitudinal reinforcement on the shear strength of FRP reinforced concrete as follows.
√̅̅̅̅ (√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ )
Vc = 0.4 f ′c bho 2ρG nG + (ρG nG )2 − ρG nG (1)

ЕG
nG = (2)
Еc

where fc’ is the concrete cylinder compressive strength, b is the width of cross-section, h0 is the effective height of specimen, ρG is the
longitudinal reinforcement ratio, EG and Ec are the elastic modulus of GFRP and concrete, respectively, nG is the ratio of EG to Ec, fcu is
the cubic compressive strength of concrete, fc’ is defined as 0.79fcu [46].
Based on the modified pressure-field theory, the Canada guideline (CSA S806–12) [51] simplifies the following shear calculation
formula.

(3)
1
Vc = 0.05km kr ks ka (fc′)3 bdv
√̅̅̅̅̅̅̅̅
Vh0 750 2.5VG h0
(4)
1
km = ≤ 1.0, kr = 1 + (EG ρG )3 , ks = ≤ 1.0, 1.0 ≤ ka = ≤ 2.5
M 450 + h0 MG

where V and M are the shear force and bending moment of specimen, respectively, b is the width of cross-section, dv is the effective
shear height, which could be simply taken as dv = 0.9h0 , for simply-supported beams subjected to concentrated force, it can be taken as
1
λ = Vh 0
М .
Chinese guideline (GB 50608–2010) [52] on the shear behaviour of FRP reinforced concrete beam shows a similar form with ACI

Table 6
Comparison of shear strength between predictions and experiments of GFRP reinforced SWSSC specimens.
Specimens Veu/kN GB 50608–2010 ACI 440.1R-06 CSA S806-12 Predictions

VCBu/kN VCBu/Veu VAu/kN VAu/Veu VCSu/kN VCSu/Veu Vpu Vpu/Veu

SH100-D8–1 25.5 19.4 0.76 15.5 0.61 10.3 0.40 24.84 0.97
SH100-D10–1 30.2 20.6 0.68 16.4 0.54 11.7 0.39 25.06 0.83
SH100-D12–1 36.6 21.3 0.58 17.0 0.46 13.1 0.36 24.89 0.68
SH100-D14–1 24.9 21.7 0.87 17.4 0.70 14.4 0.58 19.08 0.77
SH125-D8–1 41.3 24.7 0.60 19.7 0.48 19.0 0.46 35.78 0.87
SH125-D10–1 46.2 26.3 0.57 21.0 0.45 21.7 0.47 36.45 0.79
SH125-D12–1 50.9 27.4 0.54 21.9 0.43 24.2 0.47 37.05 0.73
SH125-D14–1 29.5 28.1 0.95 22.4 0.76 26.6 0.90 27.99 0.95
SH150-D8–1 52.1 29.6 0.57 23.6 0.45 30.7 0.59 46.71 0.90
SH150-D10–1 57.5 31.8 0.55 25.4 0.44 35.0 0.61 47.84 0.83
SH150-D12–1 59.8 33.2 0.56 26.6 0.44 39.0 0.65 48.90 0.82
SH150-D14–1 45.7 34.3 0.75 27.4 0.60 42.8 0.94 35.22 0.77
Mean - 0.67 - 0.53 - 0.57 - 0.84
Vari 0.13 0.11 0.18 0.08

17
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142

440.0R-06.
(√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ )
( )2
Vc = 0.86ft bho 2ρf nf + ρf nf − ρf nf (5)

where ft is the concrete cylinder tensile strength, b is the width of cross-section, h0 is the effective height of specimen, ρf is the lon­
gitudinal reinforcement ratio, nf is the ratio of EG to Ec.
The above three shear guidelines predict the ultimate shear strength of GFRP-SWSSC deep beams. The experimental results versus
the predicted calculation results are summarized in Table 6 and presented in Fig. 16. The average values of Vu/Veu are 0.67, 0.53 and
0.57, respectively. The results demonstrate that the GB 50608–2010 prediction is more accurate regarding the shear ultimate strength.
However, the predictions of three shear guidelines are too conservative and are not available for predicting the shear ultimate strength
of GFRP-SWSSC deep beams. Thus, a new shear ultimate strength calculation formula for GFRP-SWSSC deep beams without stirrups is
proposed.

4.2. Design formulae for shear strength of GFRP-SWSSC deep beams without stirrups

The deep beam mainly passes the compressive stress generated by the diagonal compression bar instead of the shear stress to
transfer the load to the support. The transmission way, the arching effect, is composed of longitudinal bars acting as tension rods and
concrete between shear-inclined cracks acting as compression rods. In this paper, the shear capacity of RWRSC deep beams with GFRP
reinforcements is calculated and analyzed according to the tension-compression bar model, as presented in Fig. 14. Based on the force
mechanism of GFRP-SWSSC deep beams, the following assumptions are made to simplify the study: (a) the concrete only undertakes
compression force, and the bottom longitudinal reinforcement undertakes all tensile forces; (b) the centroid of the tension rod and
compression rod intersects with the concentrated load at the node; (c) it is considered that the member reaches the failure load when
the tension rod yields or the compression rod fails. The complex stress flow in this model is equivalent to a rod element under uniaxial
force, in which the resultant force of compressive stress is the compression bar, the resultant force of tensile stress is the tension rod,
and the intersection of the tension rod and the compression bar is the node area. In other words, the specimens’ ultimate shear strength
depends on the tension rod, compression rod, and node area strength.
Under the symmetrical concentrated load, the specimen is composed of one horizontal tension rod (HT), one horizontal
compression rod (HC), two symmetrical diagonal compression rods (DC) and four nodes. Only the contribution of the longitudinal
GFRP reinforcement on the strength of the tension rod is considered. The compression rod’s strength is mainly due to the concrete’s
contribution to the compression rod area. As for the node area, the influence coefficient of effective compressive strength (βn) is taken
to decide the compressive strength of the node area. The design formulae achieved through calculation analysis are expressed as
follows.
Fu ≤ ηFnr (6)

Ft = FG AG (7)

Fc = fec Ac , fec = 0.85fc′ (8)

Fn = fen An , fen = 0.85βn fc′ (9)

where Fu is the force acting on a rod or joint area, η is strength reduction coefficient taken as 0.75, Fnr is the nominal strength of rod, Ft
is the strength of the tension rod, fG is the tensile strength of GFRP bar, AG is the cross-section area of GFRP bars, Fc is the strength of the
compression rod, fec is the effective strength of concrete, Ac is the cross-section area of the compression rod, Fn is the compressive
strength of the node area, fen is the effective strength of concrete in node area, An is the cross-section area of the node area, βn is the
influence coefficient of effective compressive strength of concrete in node area, taking into account of the influence of tension rods on
the bearing capacity of node area, βn=1.0 is for the node without tension rod, βn=0.8 is for the node crossed with the tension rod.
Fig. 15 provides the calculation diagram of the specimen model derived from the Strut-and-tie system. The CT node is the N2 node,
and the CC node is the N1 node. Based on the equilibrium conditions and assumptions above, the design formula could be expressed as

Fig. 14. Mechanical diagram of the shear model for GFRP reinforced SWSSC deep beams.

18
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142

Fig. 15. Calculation diagram of GFRP reinforced SWSSC deep beams.

follows.
V = FDC sin θ (10)

FHT
FDC = (11)
cos θ

Fig. 16. Comparison of shear strength between predictions and experimental results of specimens.

19
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142

1 ha
θ = tan− (12)
a

a1 + a2
ha = h − (13)
2

Table 7
Comparison of shear strength between predictions from proposed design approaches and experiments of GFRP reinforced SWSSC specimens.
Specimens H/mm D/mm SC(%) Vpu Veu Vpu / Veu

SH100-D8–1 100 8 0.00 24.8 25.5 0.97


SH100-D8–2 100 8 3.57 22.6 21.5 1.05
SH100-D8–3 100 8 7.14 18.8 17.4 1.08
SH100-D8–4 100 8 10.71 17.7 17.3 1.03
SH100-D8–5 100 8 14.29 17.2 17.9 0.96
SH100-D10–1 100 10 0.00 25.1 30.2 0.83
SH100-D10–2 100 10 3.57 22.8 27.1 0.84
SH100-D10–3 100 10 7.14 19.0 24.7 0.77
SH100-D10–4 100 10 10.71 17.9 19.1 0.94
SH100-D10–5 100 10 14.29 17.4 17.5 0.99
SH100-D12–1 100 12 0.00 24.9 36.6 0.68
SH100-D12–2 100 12 3.57 22.6 30.4 0.74
SH100-D12–3 100 12 7.14 18.9 25.9 0.73
SH100-D12–4 100 12 10.71 17.8 20.2 0.88
SH100-D12–5 100 12 14.29 17.2 19.3 0.89
SH100-D14–1 100 14 0.00 19.1 24.9 0.77
SH100-D14–2 100 14 3.57 17.3 25.0 0.69
SH100-D14–3 100 14 7.14 14.5 20.9 0.69
SH100-D14–4 100 14 10.71 13.6 18.8 0.72
SH100-D14–5 100 14 14.29 13.2 17.8 0.74
SH125-D8–1 125 8 0.00 35.8 41.3 0.87
SH125-D8–2 125 8 3.57 32.5 40.2 0.81
SH125-D8–3 125 8 7.14 27.1 34.3 0.79
SH125-D8–4 125 8 10.71 25.5 27.6 0.93
SH125-D8–5 125 8 14.29 24.8 28.6 0.87
SH125-D10–1 125 10 0.00 36.4 46.2 0.79
SH125-D10–2 125 10 3.57 33.1 40.7 0.81
SH125-D10–3 125 10 7.14 27.6 36.8 0.75
SH125-D10–4 125 10 10.71 26.0 31.5 0.83
SH125-D10–5 125 10 14.29 25.3 30.4 0.83
SH125-D12–1 125 12 0.00 37.1 50.9 0.73
SH125-D12–2 125 12 3.57 33.6 45.1 0.75
SH125-D12–3 125 12 7.14 28.1 41.4 0.68
SH125-D12–4 125 12 10.71 26.5 35.6 0.74
SH125-D12–5 125 12 14.29 25.7 34.9 0.74
SH125-D14–1 125 14 0.00 28.0 29.5 0.95
SH125-D14–2 125 14 3.57 25.4 29.5 0.86
SH125-D14–3 125 14 7.14 21.2 23.5 0.90
SH125-D14–4 125 14 10.71 20.0 20.4 0.98
SH125-D14–5 125 14 14.29 19.4 21.4 0.91
SH150-D8–1 150 8 0.00 46.7 52.1 0.90
SH150-D8–2 150 8 3.57 42.4 46.5 0.91
SH150-D8–3 150 8 7.14 35.4 42.2 0.84
SH150-D8–4 150 8 10.71 33.4 38.0 0.88
SH150-D8–5 150 8 14.29 32.4 35.5 0.91
SH150-D10–1 150 10 0.00 47.8 57.5 0.83
SH150-D10–2 150 10 3.57 43.4 53.9 0.81
SH150-D10–3 150 10 7.14 36.3 49.1 0.74
SH150-D10–4 150 10 10.71 34.2 46.1 0.74
SH150-D10–5 150 10 14.29 33.1 40.0 0.83
SH150-D12–1 150 12 0.00 48.9 59.8 0.82
SH150-D12–2 150 12 3.57 44.4 59.1 0.75
SH150-D12–3 150 12 7.14 37.1 56.7 0.65
SH150-D12–4 150 12 10.71 34.9 51.0 0.68
SH150-D12–5 150 12 14.29 33.9 49.4 0.69
SH150-D14–1 150 14 0.00 35.2 45.7 0.77
SH150-D14–2 150 14 3.57 32.0 42.2 0.76
SH150-D14–3 150 14 7.14 26.7 40.9 0.65
SH150-D14–4 150 14 10.71 25.2 37.4 0.67
SH150-D14–5 150 14 14.29 24.4 38.2 0.64
Mean 0.82
Vari 0.09

20
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142

a1 = 0.8a2 (14)

a2 = 2c + D (15)

where V is the supporting force, FDC is the strength of the DC rod, FHT is the strength of the HT rod, θ is the included angle between DC
rod and HT rod, ha is the vertical distance between tension rod and compression rod, a1 and a2 are the height of N1 node and N2 node,
respectively, c is the thickness of concrete-cover.
Based on the geometry relationships, the calculated formula for the compression rod width can be expressed as follows:
wN1 = a1 cos θ + t sin θ (16)

wN2 = a2 cos θ + t sin θ (17)

wHC = a1 (18)

where wN1 and wN2 are the height of N1 and N2 ends of DC rod, respectively, wHC is the width of HC rod.
The stress limit conditions of the compression node and node area will be followed. If the stress limit conditions were met, the shear
capacity of the tested beam could be obtained. Otherwise, the corresponding value shall be modified and then replaced into Eq. (9) for
calculation, and then the stress limit conditions should be checked again until the conditions are met. The test beam’s shear capacity
calculated by the model could be obtained.
The design formula fully considers the tensile strength of the GFRP bar. It should be noted that specimens with a GFRP bar thickness
of 14 mm showed premature failure. Considering insufficient strength, a strength reduction factor φ is taken for specimens (t=14 mm)
to consider the effect of premature failure on the shear strength of GFRP-reinforced SWSSC deep beams. The design formulation could
be expressed through fitting by SPSS software as follows.
V ,c = φVc (19)

H
φ = 0.938 − 0.022 (20)
D

in which φ is the reduction factor.


The comparison of the shear strength of specimens between estimated values using the proposed model and experimental results is
summarized in Table 6 and shown in Fig. 16(d). The mean and variance values of Vpu/Veu were 0.84 and 0.08, respectively. The
comparison indicated that the proposed model produced more accurate predictions.
Considering the effect of seashells in sand, the design calculation for GFRP-reinforced SWSSC beams was used to evaluate the shear
strength of GFRP-reinforced SWSSC beams with seashells. The design formulae for specimens of seashells obtained through SPSS are
expressed as follows.
Vpus = γVpu (21)

(22)
1.287− 0.588
γ = 1.14e S

Fig. 17. Comparison of shear strength between design predictions and experimental results of GFRP reinforced SWSSC without stirrups.

21
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142

Fig. 18. Shear load-displacement curves of GFRP reinforced SWSSC without stirrups between numerical and experimental results.

22
Z. Xing et al.
23

Case Studies in Construction Materials 20 (2024) e03142


Fig. 19. Failure modes of GFRP reinforced SWSSC without stirrups between numerical and experimental results.
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142

in which, Vpus is the shear strength of the specimen with seashell, γ is the seashell coefficient, Vpu is the shear strength of the specimen
without seashell, SC is the seashell content in sand.
The comparison of shear strength of GFRP-reinforced SWSSC beams with seashell between design formulation and experiments is
presented in Table 7 and Fig. 17. The mean value and the variance value of Vpu/Veu were 0.82 and 0.09, respectively. It is demonstrated
that a good agreement between predictions and experiments was achieved regarding shear strength. The data proves the effectiveness
of the proposed model in predicting the shear strength of GFRP-reinforced SWSSC deep beams. It is worth noting that shells in sand
decreased the shear strength of GFRP-SWSSC deep beams without stirrup.

5. Numerical modeling

A finite element model using the general-purpose program ABAQUS was simulated to evaluate the shear performance of GFRP-
SWSSC deep beams without stirrups. The objective of numerical results was to verify against experiments.

5.1. Materials

This paper selected the Concrete Damage Plasticity (CDP) model available in ABAQUS software as the concrete model to simulate
the nonlinear SWSSC behaviour. Referring to the research of Cui [53], the constitutive relationship of the SWSSC was given in Eq. (23).
The mechanical properties of SWSSC in the reference were similar to the experimental program in this study. Therefore, the
stress-strain model proposed by Cui was adopted.
{
2.38x − 1.77x2 + 0.39x3 (0 ≤ x ≤ 1)
y= [ ]− 1 (23)
x 2.38x(x − 1)2 (x > 1)

Where y and x were defined as the ratio of σ to fc and ε to ε0, respectively; σ , fc, ε0 and ε were the stress, ultimate stress strain and
ultimate strain corresponding to ultimate stress of SWSSC, respectively.
The linearly elastic model was used to simulate the GFRP bar under tension. GFRP bar was a linear and brittle material. The
expression for the stress-strain model of the GFRP bar was given in Eq. (24). After the stress of GFRP reached the fracture stress, the
fracture of GFRP occurred, and the stress value immediately decreased to 0. If the force were applied to the GFRP-reinforced SWSSC
specimens, the local stress concentration would easily appear, resulting in the non-convergence of numerical results. Thus, the rigid
element was introduced to apply concentrated force to the specimens. The elastic modulus and Poisson ratio values were 1000 GPa and
0.3, respectively.
σG = EG εG (0 ≤ εG ≤ εGu ) (24)

Where σG, EG, εG and εGu were the peak stress, elastic modulus, strain and peak strain of GFRP bar, respectively.

5.2. Establishment of FEM

A three-dimensional solid element (C3D8R) was used to model the SWSSC and steel plate. To achieve the balance between accuracy
and efficiency, the element mesh dimension of 5 mm was divided for the SWSSC section. The two-node linear truss element (T3D2) was
used to model the GFRP bar. The steel plate at supports and loading points was connected to the surface of SWSSC by using tie
constraint. Additionally, the surfaces of steel plates located in supports and loading points were coupled with reference points. The
specimen was loaded by applying displacement-controlled load through the reference points. Furthermore, the SWSSC specimen
simulated in this paper was a supported beam. In order to simulate the experimental boundary conditions, a hinge restraint was
applied at the centre of the steel plate located at both ends of the specimen bottom. The specimen could only translate longitudinally
along the beam and rotate itself.

5.3. Validation against experimental results

A total of 60 GFRP-reinforced SWSSC specimens were established in a numerical study. To verify the accuracy of the FE models, the
numerical results were compared with the data obtained from experiments. Fig. 18 compares the shear load-displacement curves of
GFRP bar-reinforced SWSSC specimens between numerical and experimental results. It can be seen from Fig. 18 that the shear load-
displacement curves obtained from FE models have a similar tendency to those of experiments. Fig. 19 presents the comparison of
failure modes of GFRP bar-reinforced SWSSC specimens between numerical and experimental results. The maximum diagonal crack in
numerical results was nearly consistent with the experimental results. Table 8 compares ultimate shear strength between numerical
and experimental results. The ultimate shear strength obtained from numerical study and experiments matched well with the mean
value (VMu / Veu=0.98). A good agreement was achieved between numerical and experimental results.

6. Conclusions

This paper conducted experimental and numerical research on the feasibility of replacing ordinary concrete with GFRP-reinforced

24
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142

Table 8
Comparison of shear strength between numerical and experimental results of GFRP reinforced SWSSC specimens.
Specimens H/mm D/mm SC(%) Veu/kN VMu/kN Error/% VMu/Veu

SH100-D8–1 100 8 0.00 25.45 25.65 0.79% 1.008


SH100-D8–2 100 8 3.57 21.52 20.92 2.79% 0.972
SH100-D8–3 100 8 7.14 17.35 17.38 0.17% 1.002
SH100-D8–4 100 8 10.71 17.30 16.83 2.72% 0.973
SH100-D8–5 100 8 14.29 17.93 17.88 0.28% 0.997
SH100-D10–1 100 10 0.00 30.23 31.67 4.76% 1.048
SH100-D10–2 100 10 3.57 27.13 29.65 9.27% 1.093
SH100-D10–3 100 10 7.14 24.70 27.05 9.51% 1.095
SH100-D10–4 100 10 10.71 19.11 21.25 11.21% 1.112
SH100-D10–5 100 10 14.29 17.50 18.79 7.36% 1.074
SH100-D12–1 100 12 0.00 36.56 37.08 1.42% 1.014
SH100-D12–2 100 12 3.57 30.41 31.15 2.43% 1.024
SH100-D12–3 100 12 7.14 25.92 25.69 0.89% 0.991
SH100-D12–4 100 12 10.71 20.18 20.14 0.20% 0.998
SH100-D12–5 100 12 14.29 19.34 19.08 1.34% 0.987
SH100-D14–1 100 14 0.00 24.92 32.57 30.69% 1.307
SH100-D14–2 100 14 3.57 25.02 31.66 26.52% 1.265
SH100-D14–3 100 14 7.14 20.94 27.29 30.32% 1.303
SH100-D14–4 100 14 10.71 18.75 23.04 22.88% 1.229
SH100-D14–5 100 14 14.29 17.75 19.75 11.28% 1.113
SH125-D8–1 125 8 0.00 41.33 41.56 0.56% 1.006
SH125-D8–2 125 8 3.57 40.15 39.44 1.77% 0.982
SH125-D8–3 125 8 7.14 34.28 33.42 2.51% 0.975
SH125-D8–4 125 8 10.71 27.56 29.31 6.35% 1.063
SH125-D8–5 125 8 14.29 28.56 27.21 4.73% 0.953
SH125-D10–1 125 10 0.00 46.23 45.88 0.76% 0.992
SH125-D10–2 125 10 3.57 40.66 41.73 2.63% 1.026
SH125-D10–3 125 10 7.14 36.84 35.25 4.32% 0.957
SH125-D10–4 125 10 10.71 31.46 32.57 3.53% 1.035
SH125-D10–5 125 10 14.29 30.36 28.48 6.19% 0.938
SH125-D12–1 125 12 0.00 50.91 48.84 4.07% 0.959
SH125-D12–2 125 12 3.57 45.13 45.49 0.80% 1.008
SH125-D12–3 125 12 7.14 41.41 41.69 0.68% 1.007
SH125-D12–4 125 12 10.71 35.63 35.80 0.48% 1.005
SH125-D12–5 125 12 14.29 34.89 33.27 4.64% 0.954
SH125-D14–1 125 14 0.00 29.53 38.87 31.62% 1.316
SH125-D14–2 125 14 3.57 29.49 36.85 24.96% 1.250
SH125-D14–3 125 14 7.14 23.52 30.18 28.32% 1.283
SH125-D14–4 125 14 10.71 20.35 25.94 27.49% 1.275
SH125-D14–5 125 14 14.29 21.39 23.12 8.11% 1.081
SH150-D8–1 150 8 0.00 52.08 50.57 2.90% 0.971
SH150-D8–2 150 8 3.57 46.53 45.61 1.98% 0.980
SH150-D8–3 150 8 7.14 42.21 43.05 1.99% 1.020
SH150-D8–4 150 8 10.71 37.98 39.02 2.74% 1.027
SH150-D8–5 150 8 14.29 35.49 34.02 4.14% 0.959
SH150-D10–1 150 10 0.00 57.49 58.82 2.31% 1.023
SH150-D10–2 150 10 3.57 53.92 54.38 0.85% 1.009
SH150-D10–3 150 10 7.14 49.11 48.16 1.93% 0.981
SH150-D10–4 150 10 10.71 46.14 44.25 4.10% 0.959
SH150-D10–5 150 10 14.29 40.00 38.40 4.00% 0.960
SH150-D12–1 150 12 0.00 59.75 60.92 1.96% 1.020
SH150-D12–2 150 12 3.57 59.13 59.01 0.20% 0.998
SH150-D12–3 150 12 7.14 56.68 54.95 3.05% 0.969
SH150-D12–4 150 12 10.71 50.98 50.50 0.94% 0.991
SH150-D12–5 150 12 14.29 49.36 48.68 1.38% 0.986
SH150-D14–1 150 14 0.00 45.65 56.62 24.02% 1.240
SH150-D14–2 150 14 3.57 42.19 53.45 26.68% 1.267
SH150-D14–3 150 14 7.14 40.87 50.03 22.41% 1.224
SH150-D14–4 150 14 10.71 37.35 46.88 25.53% 1.255
SH150-D14–5 150 14 14.29 38.20 42.98 12.52% 1.125
Mean 1.061
Vari 0.114

25
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142

SWSSC beams. Attention was paid to the crack patterns and failure modes, shear load-displacement curves, strain distribution curves
and effects of parameters on shear strength. The mechanical performance of GFRP-reinforced SWSSC without stirrups was compared
with the reference RWRSC specimens. The design formulae were put forward to evaluate the shear strength of GFRP-SWSSC deep
beams. The following conclusions could be summarized from the investigations above.

1. The cracking strength of GFRP-SWSSC deep beams was slightly higher than ordinary concrete deep beams. On the other hand,
adding seashells to sea-sand reduced the shear strength of SWSSC beams. The percentage decrease of GFRP-reinforced SWSSC
beams with 14.29% seashell content was up to 47% in terms of shear ultimate strength.
2. All GFRP-reinforced SWSSC beams exhibited similar failure characteristics, including the formation of vertical cracks and diagonal
cracks. However, the vertical crack crossed the beams section for SWSSC beams reinforced with 14 mm GFRP bars.
3. The effect of section height on the shear strength and deformation capacity of the cracking strength of GFRP-SWSSC deep beams
was significant. The stiffness and shear ultimate strength of GFRP-SWSSC deep beams were remarkably enhanced with the increase
of section height of SWSSC.
4. The shell content negatively influenced the stiffness and ultimate strength of GFRP-SWSSC deep beams. The ultimate strength of
GFRP-SWSSC deep beams decreased with the ascending of seashell content, especially when the shear-span ratio is 1.88.
5. The existing guidelines all underestimated the shear strength of GFRP-SWSSC deep beams. The proposed design formulae based on
modified tension-compression theory could accurately predict the shear strength of GFRP-SWSSC deep beams.
6. Good agreements were achieved by comparing numerical and experimental results of GFRP-reinforced SWSSC specimens, indi­
cating that the numerical model could correctly predict the shear performance of GFRP-SWSSC with seashell specimens.

CRediT authorship contribution statement

Yao Zhu: Conceptualization, Formal analysis. Yu Chen: Funding acquisition, Project administration, Software, Writing – review &
editing. Kwok-Fai Chung: Project administration. Yongbo Shao: Investigation. Enlin Ma: Data curation. Zhiquan Xing: Writing –
review & editing.

Declaration of Competing Interest

The authors strictly comply with ethical standards of research and publishing paper. This study was funded by the National Natural
Science Foundation of China (No. 52078138), Guiding Project of Fujian Province (No. 2021Y0003), Fifth Batch of Science and
Technology Plan Project of Housing and Urban-Rural Construction Industry of Fujian Province in 2022 (No.2022-K-187; 2022-K-298;
2022-K-303), and Second Batch of Science and Technology Plan Projects of Housing and Urban-Rural Construction Industry in Fujian
Province in 2023 (No.2023-K-65; 2023-K-71). The authors declared that they have no conflicts of interest to this work. We declare that
we do not have any commercial or associative interest that represents a conflict of interest in connection with the work submitted. We
declare that we have no financial and personal relationships with other people or organizations that can inappropriately influence our
work, there is no professional or other personal interest of any nature or kind in any product, service and/or company that could be
construed as influencing the position presented in, or the review of, the manuscript entitled, “Experimental and numerical research on
shear performance of GFRP bar reinforced seawater sea-sand concrete deep beams without stirrups”.

Data availability

Data will be made available on request.

Acknowledgements

The authors gratefully acknowledged the support by National Natural Science Foundation of China (No. 52078138), Guiding
Project of Fujian Province (No. 2021Y0003), Fifth Batch of Science and Technology Plan Project of Housing and Urban-Rural Con­
struction Industry of Fujian Province in 2022 (No.2022-K-187; 2022-K-298; 2022-K-303), and Second Batch of Science and Technology
Plan Projects of Housing and Urban-Rural Construction Industry in Fujian Province in 2023 (No.2023-K-65; 2023-K-71). Special
thanks to Wonstar (Shanghai) Technology Development Co., Ltd. for sponsoring all the fibre-reinforced composites involved in this
study.The authors are particularly grateful to Yuhang Fang for his support and assistance throughout the experimental study and
numerical simulations.

References

[1] J.Z. Xiao, C.B. Qiang, A. Nanni, et al., Use of sea-sand and seawater in concrete construction: Current status and future opportunities, Compos. Struct. 155 (11)
(2017) 1101–1111.
[2] Z.M. Ma, W. Li, H.X. Wu, et al., Chloride permeability of concrete mixed with activity recycled powder obtained from C&D waste, Constr. Build. Mater. 199 (2)
(2019) 652–663.
[3] J. Limeira, L. Agulló, M. Etxeberria, Dredged marine sand as a new source for construction materials, Mater. De. ConstruccióN. 62 (2012) 7–24.
[4] R. Nie, Y. Chen, Z. Xing, et al., Finite element analysis of deterioration of axial compression behavior of corroded steel-reinforced concrete middle-length
columns, Rev. Adv. Mater. Sci. 63 (1) (2024) 20230184.

26
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142

[5] C.F. Liang, H.W. Ma, W.Q. Pan, et al., Chloride permeability and the caused steel corrosion in the concrete with carbonated recycled aggregate, Constr. Build.
Mater. 218 (2019) 506–518.
[6] P. Du, A.Y. Sarah, C. Keyou, et al., Study of the influence of seawater and sea-sand on the mechanical and microstructural properties of concrete, J. Build. Eng.
42 (10) (2021) 103006.
[7] W. Liu, H. Cui, Z. Dong, et al., Carbonation of concrete made with dredged marine sand and its effect on chloride binding, Constr. Build. Mater. 120 (2016) 1–9.
[8] X.K. Guo, C.S. Xiong, Z.Q. Jin, et al., A review on mechanical properties of FRP bars subjected to seawater sea sand concrete environmental effects, J. Build. Eng.
58 (2022), 105-038.
[9] Z.Y. Wang, Z.Y. Li, J.Q. Yang, et al., Axial compressive and seismic performance of GFRP wrapped square RC columns with different scales, J. Build. Eng. 65
(2022) 105–342.
[10] Z.W. Shan, K. Liang, L.J. Chen, Bond behavior of helically wound FRP bars with different surface characteristics in fiber-reinforced concrete, J. Build. Eng. 65
(2023) 105–504.
[11] V.C. Vui, B.V. Huy, H.D. Luan, et al., Experimental behavior of fire-exposed reinforced concrete slabs without and with FRP retrofitting, J. Build. Eng. 51 (2022)
104–315.
[12] A. Junaid, A. Shehroze, Y. Tao, et al., Analytical investigation on the load-moment interaction behavior of the FRP reinforced geopolymer concrete filled FRP
tube circular columns, J. Build. Eng. 42 (2021) 102–818.
[13] T. Kartheek, T. Venkat Das, An experimental investigation on concrete with replacement of treated sea sand as fine aggregate, Mater. Today.: Proc. 27 (1) (2020)
1017–1023.
[14] O. Kayali, M.N. Haque, B. Zhu, Some characteristics of high strength fiber reinforced lightweight aggregate concrete, Cem. Concr. Compos. 25 (2) (2003)
207–213.
[15] J.M. Sousa, J.R. Correia, S. Cabral-Fonseca, et al., Effects of thermal cycles on the mechanical response of pultruded GFRP profiles used in civil engineering
applications, Compos. Struct. 116 (9-10)) (2014) 720–731.
[16] P. Feng, J. Wang, Y. Wang, D. Loughery, D. Niu, Effects of corrosive environments on properties of pultruded GFRP plates, Compos. Part B: Eng. 67 (12)) (2014)
427–433.
[17] A. Ahmed, S.C. Guo, Z.H. Zhang, et al., A review on durability of fiber reinforced polymer (FRP) bars reinforced seawater sea sand concrete, Constr. Build.
Mater. 256 (9)) (2020) 119484.
[18] Q.T. Zhang, J.Z. Xiao, Q.X. Liao, et al., Structural behavior of seawater sea-sand concrete shear wall reinforced with GFRP bars, Eng. Struct. 189 (2019)
458–470.
[19] V. Gribniak, A. Rimkus, L. Torres, et al., An experimental study on cracking and deformations of tensile concrete elements reinforced with multiple GFRP bars,
Compos. Struct. 201 (10)) (2018) 1477–1485.
[20] Q. Liao, J. Yu, T. Shi, et al., Seawater sea-sand engineered cementitious composites contribution to shear performance of beams reinforced with BFRP bars,
Compos. Struct. 320 (2023) 117183.
[21] Q. Liao, J. Yu, F. Dong, et al., FRP bars reinforced seawater sea-sand engineered cementitious composites beams with various salinities: Shear behaviors and cost
effectiveness. J. Build. Eng. 83 (2024) 108452.
[22] Q. Liao, Y. Su, J. Yu, et al., Compression-shear performance and failure criteria of seawater sea-sand engineered cementitious composites with polyethylene
fibers, Constr. Build. Mater. 345 (2022) 128386.
[23] Q. Liao, J. Yu, T. Shi, et al., Mechanical behaviors and failure criteria of seawater sea-sand engineered cementitious composites under combined tension and
shea. J. Build. Eng. 54 (2022) 104552.
[24] V. Li, S. Wang, Flexural behaviors of glass fiber-reinforced polymer (GFRP) reinforced engineered cementitious composite beams, Acids Mater. J. 99 (1) (2002)
11–12.
[25] F.M. Wegian, Effect of seawater for mixing and curing on structural concrete, IES J. Part A: Civ. Struct. Eng. 3 (4) (2010) 235–243.
[26] J. Liu, R. An, Z.L. Jiang, et al., Effects of w/b ratio, fly ash, limestone calcined clay, seawater and sea-sand on workability, mechanical properties, drying
shrinkage behavior and micro-structural characteristics of concrete, Constr. Build. Mater. 321 (28)) (2022) 126333.
[27] D.L. Narver, Good concrete made with coral and water, Civ. Eng. 24 (1964) 654–658.
[28] E. Yang, M. Kim, H. Park, et al., Effect of partial replacement of sand with dry oyster shell on the long-term performance of concrete, Constr. Build. Mater. 24 (5)
(2010) 758–765.
[29] F.L. Qu, W.G. Li, Z. Tang, et al., Property degradation of seawater sea sand cementitious mortar with GGBFS and glass fiber subjected to elevated temperatures,
J. Mater. Res. Technol. 13 (7-8)) (2021) 366–384.
[30] P.R. Li, W.G. Li, T. Yu, et al., Investigation on early-age hydration, mechanical properties and microstructure of seawater sea sand cement mortar, Constr. Build.
Mater. 249 (7)) (2020) 118776.
[31] Z. Xing, Y. Zhu, Q. Liu, et al., Experimental study on axial compression behavior of square pultruded GFRP tubular stub column, Case Stud. Constr. Mater. 20
(2024) e03061.
[32] P.D. Li, T.Q. Yang, Q. Zeng, et al., Axial stress-strain behavior of carbon FRP-confined seawater sea-sand recycled aggregate concrete square columns with
different corner radii, Compos. Struct. 207 (1) (2019) 576–588.
[33] Z. Dong, G. Wu, X.L. Zhao, et al., Mechanical properties of discrete BFRP needles reinforced seawater sea-sand concrete-filled GFRP tubular stub columns,
Constr. Build. Mater. 244 (2020) 118330.
[34] Y. Zhang, Y. Wei, J. Bai, et al., A novel seawater and sea sand concrete filled FRP-carbon steel composite tube column: concept and behaviour, Compos. Struct.
246 (2020) 112421.
[35] T. Ozbakkaloglu, J.C. Lim, T. Vincent, FRP-confined concrete in circular sections: Review and assessment of stress–strain models, Eng. Struct. 49 (4)) (2013)
1068–1088.
[36] H.J. Liang, S. Li, Y.Y. Lu, et al., Electrochemical performance of corroded reinforced concrete columns strengthened with fiber reinforced polymer, Compos.
Struct. 207 (1)) (2019) 576–588.
[37] A. Aha, A. Hmm, B. Cec, et al., Evaluating the shear design equations of FRP-reinforced concrete beams without shear reinforcement, Eng. Struct. 235 (5) (2021)
112017.
[38] Y. Zhu, Y. Chen, K. He, et al., Flexural behavior of concrete-filled SHS and RHS aluminum alloy tubes strengthened with CFRP, Compos. Struct. 238 (4)) (2020)
111975.
[39] A. Zhou, R.Y. Qin, C.L. Chow, et al., Structural performance of FRP confined seawater concrete columns under chloride environment, Compos. Struct. 216 (7))
(2019) 12–19.
[40] J.L. Yang, J.Z. Wang, Z.R. Wang, Axial compressive behavior of partially CFRP confined seawater sea-sand concrete in circular columns-Part I: Experimental
study, Compos. Struct. 246 (8)) (2020) 112373.
[41] Y. Wei, J.W. Bai, Y.R. Zhang, et al., Compressive performance of high-strength seawater and sea sand concrete-filled circular FRP-steel composite tube columns,
Eng. Struct. 240 (8)) (2021) 112357.
[42] J.L. Yang, J.Z. Wang, Z.R. Wang, Axial compressive behavior of partially CFRP confined seawater sea-sand concrete in circular columns-Part II: A new analysis-
oriented model, Compos. Struct. 246 (8)) (2020) 112368.
[43] X.Y. Zhang, Z.J. Xu, Y. Zhu, et al., Behavior of square pultruded GFRP tube filled with seawater and sea sand concrete beam, Compos. Struct. 284 (3)) (2022)
115183.
[44] Z.Q. Dong, Y. Sun, H. Zhu, et al., Shear behavior of hybrid seawater sea-sand concrete short beams reinforced with BFRP reinforcements, Eng. Struct. 252 (2))
(2022) 113615.
[45] Z.Q. Dong, G. Wu, X.L. Zhao, et al., Behaviors of hybrid beams composed of seawater sea-sand concrete (SWSSC) and a prefabricated UHPC shell reinforced with
FRP bars, Constr. Build. Mater. 213 (2019) 32–42.

27
Z. Xing et al. Case Studies in Construction Materials 20 (2024) e03142

[46] Chinese Code. Code for Design of Concrete Structures. GB50010-2010, Beijing, China. (in Chinese), 2010.
[47] A.F. Ashour, I.F. Kara, Size effect on shear strength of FRP reinforced concrete beams, Compos.: Part B 60 (4) (2014) 612–620.
[48] G. Kani, The riddle of shear failure and its solution, J. Am. Concr. Inst. (1964).
[49] Z. Achillides, K. Pilakoutas, Bond behavior of fiber reinforced polymer bars under direct pullout conditions, J. Compos. Constr. 8 (2) (2004) 173–181.
[50] ACI Committee (ACI). ACI 440.1R-06: Guide for the Design and Construction of Structural Concrete Reinforced with FRP Bars. Farmington Hills: ACI
Committee, 2006.
[51] Canadian Standards Association (CSA). CSA S806-12: Design and Construction of Buildings Components with Fiber reinforced Polymers. Toronto: Canadian
Standards Association, 2012.
[52] Chinese Standard (CS). GB 50608-2010: Standard for construction engineering application of fiber reinforced composites, Beijing, China, 2003 (in Chinese).
[53] Cui M., Mao J.Z., Jia D.G., et al. Experimental Study on Mechanical Properties of Marine Sand and Seawater Concrete. 2014 International Conference on
Mechanics and Civil Engineering (icmce-14). 2014.

28

You might also like