You are on page 1of 22

Case Studies in Construction Materials 19 (2023) e02643

Contents lists available at ScienceDirect

Case Studies in Construction Materials


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

Case study

Strut-and-tie model and its applications in reinforced concrete


deep beams: A comprehensive review
Imad Shakir Abbood
Engineering Affairs Department, Sunni Endowment Diwan, Baghdad, Iraq

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

Keywords: Strut-and-tie model (STM), with its truss analogy, is yet a powerful tool for analyzing and
Reinforced concrete designing complex prestressed and reinforced concrete structures. Which represents a simplistic
Concrete structures method that predominantly provides conservative strength predictions. This article aims to
Strut-and-tie model
provide a comprehensive review on the STM and its applications in reinforced concrete (RC) deep
Truss model
Discontinuity regions
beams that are available in the open literature. Accordingly, a review of the STM technique is
Deep beams presented with a particular focus on the basic concept of STM, associated B- and D-regions,
Shear strength fundamentals of STM, historical development of STM, and its applications in RC deep beams. The
review also covers a wide range of STM aspects including: strut strength and its effectiveness
factor, node strength and its effectiveness factor, tie strength and minimum reinforcement, the
effect of tie anchorage, and strain energy minimization. Hence, essential formulas and parameters
of STM proposed by various design codes and researchers are reviewed and then summarized in
comprehensive tables. Consequently, from the reviewed literature, several challenges are high­
lighted. Findings and recommendations that are necessary to be considered in future works are
addressed. With the provided information, this review article serves as a useful reference for
structural engineers and researchers who are familiar with STM, as well as who wish to enter this
research field.

1. Introduction

The shear design of deep concrete structural members is a challenging issue that is still being debated. Since the 1960 s, various
empirical and analytical approaches have been proposed to evaluate the shear capacity of reinforced concrete (RC) deep beams. The
most commonly employed approach among those is the strut-and-tie model (STM) (see Fig. 1). Therefore, STM has emerged as an
effective and alternative design method for deep concrete structural members, those of high shear stresses. Although the STM is
applicable for all parts of concrete members, it is employed only for D-regions (discontinuity regions), where the distribution of strain
within a cross-section is significantly nonlinear [1,2]. Typical examples of such regions are deep beam [3], corbel [4], beam-column
joint [5], pile cap [6,7], confined masonry wall [8], and squat wall [9].
The STM is conceptualized based on lower-bound plasticity theorem and is strictly an equilibrium model. In which, the stresses are
calculated by satisfying the equilibrium and yield criterions, whilst neglecting the compatibility condition. The engineers are free to
choose the efficient STMs that satisfying the equilibrium condition [11]. Therefore, for any given structure, several STMs are possible
for application. STM idealizes the flow of forces in cracked concrete member via a truss connected by node regions, so that all internal
forces effects can be traced in complicated details without requiring separate shear or flexural models. Thus, the real stress fields can be

E-mail address: imadshakirabbood@gmail.com.

https://doi.org/10.1016/j.cscm.2023.e02643
Received 10 September 2023; Received in revised form 13 October 2023; Accepted 2 November 2023
Available online 3 November 2023
2214-5095/© 2023 The Author. 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/).
I.S. Abbood Case Studies in Construction Materials 19 (2023) e02643

described in a discrete manner [12–14].


Since the STM is classified as a lower-bound plastic method, theoretically, this indicates that the estimated ultimate capacity is
consistently equal to or below the actual collapse load. Which leads to a safe assessment regarding the ultimate load-carrying capacity
and applying such a method is reasonable [15]. For this reason, the STM is gaining growing tendency toward application. Nevertheless,
although the STM provides a rational solution, it is often termed an approximation procedure with uncertain accuracy, which makes
the designers uncomfortable [16–19]. Detailed information and schematic representation of the STM can be found in Section 3.
The general design procedures using the STM for prestressed and reinforced concrete structures are addressed by various codes of
practices such as AASHTO LRFD [20], ACI 318 [21], CSA A23.3 [22], CSA S806 [23], CSA S6 [24], NZS 3101 [25], DIN 1045 [26],
Model Code [27], AS 3600 [28], and Eurocode 2 [29]. In addition, various applications of the STM have been investigated by many
researchers [30–45]. However, in spite of these broad applications along with design procedures and its simple concept, the complexity
and uncertainty in shear strength prediction are still existing. Furthermore, there is no clear guidance or commentary to define the STM
geometry in practice. The major challenge is how to transfer a continuous domain of a structure to an STM. Consequently, this is a
difficult task for designers in selecting an appropriate STM, particularly for RC deep beams with different configurations and complex
loading conditions. Knowing that, the STM geometry changes with beam configurations and loading conditions (e.g., equality,
magnitude, symmetry, combination, etc.).
Only a few review articles on the applications of STM in structural engineering can be found in the open literature [46–48].
Therefore, the aim of this paper is to provide a comprehensive review on STM and its applications in RC deep beams that are available
in the literature. The paper attempts to present a review on the STM technique with a particular focus on the basic concept of STM,
associated B- and D-regions, STM fundamentals, STM historical development, and STM applications in RC deep beams. The review also
aims to illustrate the most significant challenges that face the structural engineers and designers. Hence, it intends to draw several
recommendations which are necessary to be considered in future works for improving the STM.

2. B-regions and D-regions

It is important first to define B- and D-regions before going further through the STM. Based on strain distribution in a cross-section,
a concrete structure is divided into two types of regions, each of which has a different design approach. Those regions where Euler-
Bernoulli’s theory (“plane sections remain plane”) is applicable, are known as beam or Bernoulli regions, or (B-regions). Whereas, the
other regions in which the Euler–Bernoulli theory is not applicable, are known as discontinued or disturbed regions, or (D-regions)
[49].
B-regions can be found in (plates, beams, frames, etc.), when the depth of cross-section is constant or changed gradually and the
load is distributed continuously (see Fig. 2). Hence, B-regions exhibit linear strain distribution within a cross-section. The internal state
of stresses can be simply derived using sectional effects (e.g., axial and shear forces, torsional and bending moments). These stresses are
calculated based on section properties (cross-section area and moment of inertia) as long as the section is uncracked [50]. Alterna­
tively, the truss model will be applied if tensile stresses are greater than tensile strength of concrete. In addition, code provisions such
as Eurocode 2 and ACI 318 have permitted other standard procedures that are experimentally verified [51].
D-regions are found in (deep beam, beam-column joint, corbels, pile cap, squat wall, etc.). D-regions exhibit significant nonlinear
strain distributions within the cross-section as a result of discontinuity induced by the sudden change in concentrated loads (static
discontinuity) or geometry (geometric discontinuity). The high nonlinearity of these regions challenges structural engineers to choose
an effective and accurate design method. Thus, among various methods, the STM has been considered an effective and simple method
to design such regions [52]. Fig. 2 and Fig. 3 show the typical D-regions in a concrete structure.
According to Schlaich et al. [53,54], for determining the separating sections between B- and D-regions, there is no much accuracy
required. These sections are assumed to be at (h) distance from the concentrated load or the geometric discontinuity, where (h) equals
the depth of the nearby B-region as shown in Fig. 3. This presumption is based on Saint-Venant’s principle. However, a concrete
structure such as deep beam is considered one entire D-region [55].

Fig. 1. Shear capacity models used for concrete deep beams [10].

2
I.S. Abbood Case Studies in Construction Materials 19 (2023) e02643

Fig. 2. B- and D-regions in a concrete frame [56].

Fig. 3. D-regions: geometric discontinuity (a-c); static discontinuity (d-f) [21].

Fig. 4. Different STM elements in a symmetrical RC deep beam [62], where; Cc is the diagonal compression forces, H is the horizontal forces, T is
the tension forces, and θs is the angle of diagonal strut with respect to horizontal axis.

3
I.S. Abbood Case Studies in Construction Materials 19 (2023) e02643

3. Overview of STM

This section provides a brief information about STM basic concept. Hence, the various STM elements are defined and explained.
Each element is briefly discussed in terms of its shape and function within the structural member. In addition, the historical devel­
opment of STM are also provided for reader.

3.1. Fundamentals of STM

STM is a design procedure for complex prestressed and reinforced concrete members, that simplifies the complicated states of
stresses to a collection of simple stress paths. STM visualizes a truss-like model in a concrete structure or its parts. It comprises three
types of elements: concrete struts as compression members, ties as tension members, and connecting nodes. Concrete around a node
can be defined as a nodal zone, which transfers the diagonal strut forces to other struts and ties. Fig. 4 shows the STM elements in a
simply-supported RC deep beam. The details of each element of STM are as follows:

3.1.1. Struts
Struts can be employed for multiple purposes (see the red dashed lines in Fig. 4). They are employed in the truss system as a
compressive element to resist the moment. They also function as diagonal strut, transferring shear forces to supports. The diagonal
strut is typically arranged parallel to the anticipated cracking direction. Struts are classified according to their shape in the STM,
namely, prism, bottle-shaped, and fan-shaped (see Fig. 5). The fan-shaped struts are used for the RC deep beam with uniform
distributed loads. The strut dimensions are affected by many factors: compression strut size, steel reinforcement distribution, and tie
anchorage [57,58].

3.1.2. Ties
Ties refer to the tension members (see solid lines “T” in Fig. 4). They comprise longitudinal and transverse reinforcements, as well
as concrete around the reinforcement. A crucial hint to observe in detailing STM is that sufficient anchorage for reinforcing must be
provided. On the condition that anchorage is not enough, then brittle failure may take place at the loading stage below the expected
ultimate capacity [52,59].

3.1.3. Nodes
Nodes are the intersection points of STM elements (struts and ties) with concentrated forces (see Figs. 4 and 6). At least three forces
should act on a node in order to satisfy the equilibrium condition. Nodes are classified based on the sign of these forces: C-C-C, C-C-T,
and C-T-T, in which C and T refer to compression and tension, respectively [60,61].

3.2. Historical development of STM

At the end of 19th century, during early development of design procedures for structural concrete, it was recognized that flexure
theories were insufficient to deal with regions exposed to significant shear stresses [48]. Hence, a truss model for slender beams
(B-regions) was developed firstly by Ritter (1899) [63] to promote the role of web reinforcements in improving the shear capacity. The
compression struts interact with stirrups and longitudinal reinforcements (compression and tension) to form a plane-truss analogy (45◦
truss model) (see Fig. 7). In 1902, Mörsch [64] attempted to refine Ritter model. Mörsch indicated that discrete diagonal forces in the
Ritter model can be ideally presented as a continuous field of diagonal compression as shown in Fig. 8.
Experimental investigations carried out by Talbot (1909) [65] indicated that truss model exhibited overestimations in strength.

Fig. 5. Common types of struts in deep beam [58].

4
I.S. Abbood Case Studies in Construction Materials 19 (2023) e02643

Fig. 6. Node designation [61].

Fig. 7. Ritter truss model (1899) [58].

This overestimation was occurred due to neglect of concrete tensile strength. In 1927, Richart [66] proposed more sophisticated truss
model, where the inclination of diagonal strut was allowed to differ from 45◦ . Rausch (1929) [67] expanded the plane-truss to a
space-truss and introduced the torsion-resisting mechanism for RC beam. The truss model was later refined and extended by Kupfer
(1964) [68], Leondardt (1965) [69], and others until Thurlimann et al. (1975, 1983), Mueller (1978) [70], Nielsen and Hoang (1984)
[71], and their coauthors derived its scientific basis for a logical approach by tracing the concept back to the plasticity theorem. In
1984, the model was adopted by CSA A23.3 as a standard method in the (“Design of Concrete Structures”) considering the compression
field theory (CFT).
After the investigations carried out by Marti (1985) [72], Collins and Mitchell (1986) [73], and Rogowsky and Macgregor (1986)
[74], the STM has been successfully derived and adopted to solve various RC problems. Accordingly, Schlaich et al. (1987) [54]
expanded the truss model to a consistent STM for application to all parts of a concrete structure and all concrete structures. Thereafter
in 1989, the STM was first incorporated by AASHTO in the (“Guide Specifications for Design and Construction of Segmental Concrete
Bridges”). Ramirez and Breen (1991) [75] revealed that a modified truss model with an adjustable inclination of diagonal struts may be
adopted to design prestressed and reinforced concrete beam. Since then, STM has been applied in designing prestressed concrete
members in various applications [76–85].
Parallel to the sectional method, the STM has been extensively adopted in designing deep beams [53,77,86–88]. Experimental
investigations conducted by Tan et al. (1997) [88] indicated that the STM was capable to evaluate the strength of RC deep beam under
combined loading. The strength prediction was found consistent and conservative. The method was considered more rational than
other empirical models. The STM and related procedure were summarized in the state-of-the-art report of ASCE-ACI 445 in 1998 [89].
Moreover, the STM procedure has been adopted for design by both AASHTO LRFD (1998) in the (Bridge Design Specifications) and ACI
318 (2002) in the (Building Code Requirements). Nowadays, the STM has been incorporated in many worldwide code provisions such
as CSA A23.3, CSA S806, CSA S6, NZS 3101, DIN 1045, Model Code, AS 3600, and Eurocode 2.

4. STM applications in RC deep beams

Hwang et al. [90] introduced a softened STM for RC deep beams that satisfies stress equilibrium, strain compatibility, and ma­
terial’s constitutive law (see Fig. 9). The load-carrying mechanisms in this model were governed by diagonal, horizontal, and vertical.
The softening effects of the concrete involved controlling the shear resistance. This STM type may be called “an analytical STM”.

Fig. 8. Mörsch truss model (1902) [58].

5
I.S. Abbood Case Studies in Construction Materials 19 (2023) e02643

However, this model is not suitable for practical design purposes because it was proposed with complicated formulations. Hwang and
Lee [91] presented a simplified softened STM for different types of discontinuance regions. Later, the softened STM was further
modified by Lu et al. [92] for deep beams loaded through a column stub.
Tan et al. [93,94] formulated analytical STMs for RC deep beams without/with web reinforcements. The stress field was assumed to
be distributed uniformly through the strut width and its length. Both models adopted Mohr-Coulomb failure criterion. Additionally,
several simplified STMs were proposed for RC deep beams by deriving explicit formulations, for example models of Matamoros and
Wong [95], Russo et al. [96], and Chetchotisak et al. [97]. These STMs estimate the shear capacity induced by diagonal strut, lon­
gitudinal bars, and web reinforcements. The softening factor was taken into account to determine the concrete strut strength, whereas
experimental data were used to calibrate the contribution of web reinforcements. These STMs may be called “semiempirical STMs”.
Tang and Tan [98] presented an interactive STM for deep beams. The shear failure in this model was governed by crushing of strut
and diagonal splitting. This model adopted Mohr-Coulomb failure criterion to describe the interaction between the two failure modes.
Furthermore, the contributions provided by concrete, longitudinal reinforcement, and web reinforcement to the splitting capacity
were supposed to contribute equally. The model of Tang and Tan [98] was modified later by Tan and Cheng [99] in order to better
capturing the size effect under shear. This modified model accounts for size effect by adjusting the strut geometry, as well as diameter
and spacing of shear reinforcements. The modified STM was proposed to estimate the size-dependent shear capacity for deep and
slender beams. The model was further modified by Zhang and Tan [100] with consideration of the effects of concrete cracking and
tensile strain on flexural steel.
Brown and Bayrak [101,102] developed semiempirical STMs for designing deep beams. These two models predict the member
capacity contributed by concrete strength, strut inclination, geometry of a bottle-shape strut, as well as web reinforcement. Zhang et al.
[103] proposed an STM to evaluate the behavior of unsymmetrically-loaded deep beams through an iterative procedure to identify the
top nodal zone depth, which required extensive computational effort. Another semiempirical STM was proposed by Yang and Ashour
[104] for deep beams. The model considered the size effect under shear based on fracture mechanics. The shear failure in this model
was governed by crushing of strut. Moreover, Lim and Hwang [105] introduced a modified macro STM-based approach to reflect the
shear behavior of deep beam. The model was verified against experimental tests and considered various variables such as identifying
shear elements compatible with force discontinuity, defining the strut size by considering the elastic behavior, the effect of loading
plate on force spreading, as well as selecting the proper failure modes.
For a better predicting of shear capacity of deep beams, so-called cracking STM was developed by Chen et al. [16,106]. The model
considered the STM concept combined with the influences of diagonal crack patterns and strain distributions in horizontal rein­
forcement within shear span as shown in Fig. 10. The model showed good agreement with experimental tests, but required complicated
calculations. Recently, Fan et al. [1] proposed an STM for unsymmetrically-loaded RC deep beam (see Fig. 11). The geometry of
compression nodal zones was determined by adopting the Mohr’s Circle and minimum strain energy criterions. Chetchotisak et al. [62]
developed a modified interactive STM for RC deep beams. The STM was developed based on two independent load-carrying mech­
anisms: inclined strut and truss. The mechanism of strut was modified from interactive STM. A new failure mode for concrete was
derived by implementing empirical constants into Mohr–Coulomb failure criterion.
In literature, deep beams were also strengthened by other means of strengthening such as fiber reinforced polymers (FRPs) [107]
and fiber reinforced concrete (FRC) [108] to enhance the shear capacity. Accordingly, there have been attempts on improving the STM
to design (FRPs)-strengthened concrete deep beam [109–120]. An STM model was proposed by Park and Aboutaha [119] to assess
shear capacity of RC deep beams strengthened by carbon fiber reinforced polymer (CFRP) sheets. The CFRP sheets were wrapped
around the complete beam depth. The CFRP strips were considered as additional tensile reinforcements (ties). Therefore, a reduction
factor of 0.75 to the nominal strength was used to integrate CFRP strips into the steel stirrups. While, Godat and Chaallal [109] used a
modified STM to predict the shear capacity of large-scale RC beams strengthened by bi-directional CFRP laminates wrapped around the
web. In this model, the tensile strength of CFRP wraps was determined from ACI-440 guidelines and then combined with tensile
strength of steel stirrups.

Fig. 9. Softened STM proposed by Hwang et al. [90].

6
I.S. Abbood Case Studies in Construction Materials 19 (2023) e02643

Fig. 10. Configuration of cracking STM for RC deep beam developed by Chen et al. [16].

Fig. 11. STM geometry for an unsymmetrically-loaded RC deep beam developed by Fan et al. [1].

Panjehpour et al. [118] conducted experimental tests and suggested a modified strut effectiveness factor for an STM. The model was
restricted to normal RC deep beam strengthened by unidirectional CFRP sheets applied by using the wet-lay methods. Recently,
Mustafa et al. [116] developed a simplified analytical STM for (FRPs)-strengthened RC deep beams. The model predicts the shear
capacity contributed by diagonal strut, longitudinal rebars, web reinforcements, span-to-depth ratio, and FRP composite. Perfect
bonding was assumed between FRP materials and concrete surface. The shear failure in this model was governed by crushing of strut,
diagonal splitting, and tensile rupture of FRP composite. The developed STM is applicable for simply-supported deep beam
strengthened by any type of FRP.
Similarly, there have also been attempts on improving the STM to design FRC deep beams [121–123]. Deng et al. [121] proposed a
modified STM for both RC and FRC deep beams applications (Fig. 12). The contribution of concrete tensile strength and longitudinal
reinforcements to the shear strength were modified in this model. The model was evaluated against experiments and showed good
agreement. Recently, Sagi et al. [123] proposed a simplified fiber reinforced cracking STM (FR-CSTM) for deep beams based on

7
I.S. Abbood Case Studies in Construction Materials 19 (2023) e02643

Fig. 12. A modified STM proposed by Deng et al. [121].

experiment observation and extending the existing cracking STM of Chen et al. [16]. The proposed model incorporated the effects of
discrete steel fibers and the extended length of loading plate based on some assumptions (see Fig. 13).
There are few applications of STM on evaluating the strength and behavior of RC deep beams under dynamic loading [124–127]
and cyclic loading [128–130] in the literature. Apart from aforementioned conventional design procedures, Perera et al. [12,131],
Hardjasaputra [132], Yavuz [133], Dhahir [134,135], Hanoon et al. [136], Zaborac et al. [137], and Silveira et al. [2] can be
mentioned as other researchers’ scholarly who adopted the optimization-based STM approach for designing RC deep beam. Also, it
should be noted that all previous researchers adopted the STM to evaluate the strength of concrete deep beams at ultimate capacity of
limit state. Yet, very few efforts have been made on modifying the STM to investigate the serviceability behavior of such members after
cracking [138–141]. To this end, the following sections are discussing the most challenging aspects of the STM.

5. Strength of struts

As known, the design of deep structural concrete members based on STM is relying on the lower-bound plasticity theorem. This
theory assumes that concrete and steel ties behave like perfect plastic materials and full redistribution of internal stresses occurs. The
properties and behavior of steel ties are well-known and the tensile strength can be predicted accurately. However, this assumption is
not true for concrete or members failing in compression. Therefore, in order to consider such an imperfect assumption, modification
factors are applied to adjust both dimension and strength of concrete members [48,142–144]. The effective strength of the concrete
strut fce is determined as:

fce = v f ˊc ; v ≤ 1 (1)

where v and f ˊc are the effectiveness factor of strut and concrete strength, respectively [48]. Typically, the design strength of a concrete
strut fcd is specified as:
fcd = Φ fce ; Φ ≤ 1 (2)

Fig. 13. Schematic representation of FR-CSTM proposed by Sagi et al. [123].

8
I.S. Abbood Case Studies in Construction Materials 19 (2023) e02643

in which Φ is a strength reduction factor applied to reflect the imperfection in strut.

5.1. Effectiveness factor of struts

The effectiveness factor (v) is called “efficiency factor”, also lately named as the “strength reduction factor”. It is applied to reflect
the reduction in concrete strength due to the softening effect of cracked concrete [145]. It is depending on the strut shape and presence
of transverse tensile strain in shear span. This factor also takes into account any further shortcoming in the STM. Research has shown
that the effectiveness factor value of strut depends generally on a wide range of variables such as steel reinforcement, beam size,
material properties, strut geometry, and structure loading [46,146–149] (see Fig. 14). Therefore, these variables should be considered
in calculating the effectiveness factor. The existing effectiveness factors of concrete struts proposed by various researchers and
standard codes are summarized Table 1.

5.2. Evaluation of existing effectiveness factors of struts

As seen in Table 1, different models of the effectiveness factor for struts have been proposed. In addition, several attempts have
been made to evaluate these models [42,119,144,158]. However, there has been a considerable debate about the value of these factors.
Park and Aboutaha [119] examined seven effectiveness factor models proposed by [20,21,150,151,153–155] against 17 RC deep
beams from literature. It was recognized that effectiveness factor models depending on concrete strength (e.g., Bergmeister et al. [150]
and MacGreger [154]) and those models depending on strut shape and concrete weight (e.g., ACI 318 [21]) showed a poor correlation
with experimental result. While, models based on strut angle (e.g., Vecchio and Collins [151] and Foster and Gilbert [153]) and those
models based on modified compression field theory (e.g., AASHTO [20]) demonstrated a good correlation with experiment result. On
the other hand, the model of Kaufman and Marti [155] overestimated the beam strength result. Fig. 15 demonstrates the statistical
analysis of the examined 7 effectiveness factor models. In conclusion, the LRFD’s [20] model was recommended for conservative
analysis since it delivered the smallest standard deviation. Furthermore, Foster and Gilbert’s [153] model was recommended for
strengthened beams to achieve the most accurate result since it delivered the smallest median value among all other models, while did
not produce negative differences.
Tuchscherer et al. [42] suggested an STM, and hence conducted a statistical comparison to evaluate the effectiveness factor of his
STM with other three models proposed by AASHTO LRFD [20], ACI 318 [21], and Model Code [27] against a dataset of 179 RC deep
beams. The statistical analysis of examined effectiveness factor models is shown in Fig. 16. Based on observation, the Model Code [27]
produced more accurate results than other models in predicting the capacity since it delivered the smallest median value. The model of
Tuchscherer et al. [42] was a significant improvement in predicting the capacity over AASHTO LRFD [20] and ACI 318′s [21] models,
while slightly less accurate than Model Code [27]. On the other hand, it was observed that AASHTO LRFD’s [20] model produced less
accurate result than ACI 318′s [21] model. This trend, however, was due to variations in triaxial confinement factors.
Recently, Ismail et al. [158] carried out a parametric investigation to evaluate eight effectiveness factors used in STM proposed by
codes [21,27,29] and researchers [72,73,151–153] against extensive datasets of experimental results from literature. Accordingly, 519

Fig. 14. Factors that affect the strength of strut [125].

9
I.S. Abbood Case Studies in Construction Materials 19 (2023) e02643

Table 1
Effectiveness factor models of concrete struts.
Model Effectiveness factor (v) Description

Marti [72] 0.6 for general use


Collins and Mitchell [73] 1/(0.8 + 170ε1 ) ≤ 1.0 12 ≤ f ˊc ≤ 35MPa
ε1 = εs + (εs − ε2 )cot2 θ
Schlaich et al. [54] 0.85 uncracked strut
0.68 cracks parallel to the strut (θ = 60 )

0.51 cracks skewed to the strut (θ = 45 )


0.34 minimum strength of strut (θ ≤ 30 )


√̅̅̅̅
Ramirez and Breen [75] 15 ≤ f ˊc ≤ 45 Mpa
2.5/ f ˊc
Alshegeir and Ramirez [77,87] 0.8–0.95 uncracked strut
0.75 cracks parallel to the strut (θ = 60 )

0.50 cracks skewed to the strut (θ = 45 )


0.2–0.25 minimum strength of strut (θ ≤ 30 )


√̅̅̅̅
Bergmeister et al. [150] 20 ≤ f ˊc ≤ 80 Mpa
0.5 + 1.25/ f ˊc
Vecchio and Collins [151] 1/(1 + kc kf )
[− ε ]0.80
1
kc = 0.35 − 0.28
ε2
√̅̅̅̅
kf = 0.1825 f ˊc ≥ 1.0
Warwick and Foster [152] 0.85 uncracked strut
0.81 cracks parallel to the strut (θ = 60 )

0.63 cracks skewed to the strut (θ = 45 )


0.45 minimum strength of strut (θ ≤ 30 )


Foster and Gilbert [153] 1 20 ≤ f ˊc ≤ 100 Mpa


a2
1.14 + (0.64 + f ˊc /570)(
z)
MacGregor [154] v1 v2
v1 = 1.00 uncracked strut
v1 = 0.80 cracked strut with transverse bars
v1 = 0.65 cracked strut without transverse bars
v1 = 0.60 severely cracked strut
√̅̅̅̅
v2 = 0.55 + 15/ f ˊc
Kaufmann and Marti [155] 1
1/3
(0.4 + 30ε1 )f ˊc
Zhang and Hsu [156] 5.8 1 0.9
√̅̅̅̅ √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ ≤ √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
f ˊ 1 + 400ε1
c
1 + 400ε1

Russo et al. [96]


Zwicky and Vogel [157] (1.8 − 38ε1 )f ˊc
(− 1/3)
0.85f ˊc
(− 1/3)
≤ v ≥ 1.6f ˊc
(− 1/3)

Brown and Bayrak [101] 0.89 without web reinforcement


√̅̅̅̅
a/d f ˊc
2.7 with web reinforcement
√̅̅̅̅
a/d f ˊc
Tuchscherer et al. [42] v = 0.45 ≤ 0.85 − f ˊc /138 ≤ 0.65 28 ≤ f ˊc ≤ 55 Mpa
AASHTO LRFD [20] 1/(0.8 + 170ε1 ) ≤ 0.85 ε1 = εs + (εs + 0.002)cot2 θ
ACI 318 [21] 0.85βs
βs = 1.00 horizontal strut
βs = 0.75 bottle-shape strut with bars
βs = 0.4λ bottle-shape strut w/o bars
λ = 1.00 normal weight concrete
λ = 0.85 sand-lightweight concrete
λ = 0.75 lightweight concrete
βs = 0.40 tensioned members
βs = 0.60 all other cases
CSA A23.3 [22] 1/(1.14 + 0.68cot2 θ) ≤ 0.85
CSA S806 [23] 1/(0.8 + 170ε1 ) ≤ 0.85 for FRP bars
ε1 = εf + (εf + 0.002)cot2 θ
CSA S6 [24] 1/(0.8 + 170ε1 ) ≤ 0.85
ε1 = εs + (εs + 0.002)cot2 θ
DIN 1045 [26] 1.00 η1 uncracked strut
0.75 η1 parallel to cracks
η1 = 1.00 normal weight concrete
η1 = 0.4 + 0.6(ρ/2200) lightweight concrete
Model Code [27] α(30/fck )1/3 ≥ 0.55 for general use
AS 3600 [28] 0.80 − fck /200 for general use
Eurocode 2 [29] 0.6(1–fck /250) strut with tensile stresses

10
I.S. Abbood Case Studies in Construction Materials 19 (2023) e02643

Fig. 15. Descriptive statistics of effectiveness factor models by differences (%) [119].

Fig. 16. Statistical analysis of effectiveness factor models by Tuchscherer et al. [42].

11
I.S. Abbood Case Studies in Construction Materials 19 (2023) e02643

RC deep beams without and with the contributions of web reinforcement were utilized to investigate the effect of effectiveness factor
on the STM in capturing the shear strength. The results are demonstrated in Figs. 17 and 18. The statistical analyses are summarized in
Figs. 19 and 20. Overall, strength prediction for beams with shear reinforcements were more conservative than those without shear
reinforcements. It was also concluded that, specifically for RC deep beams without web reinforcement, the effectiveness factors rec­
ommended by ACI 318, Eurocode 2, Model Code, and Marti produced very unsafe results. The justification is these models don’t really

Fig. 17. Effects of effectiveness factors on STM in predicting shear capacity for RC deep beams without shear reinforcement [158].

12
I.S. Abbood Case Studies in Construction Materials 19 (2023) e02643

Fig. 18. Effects of effectiveness factors on STM in predicting shear capacity for RC deep beams with shear reinforcement [158].

consider all of the essential contributing aspects (e.g., shear reinforcement, shear span-to-depth ratio, and concrete compressive
strength). However, although Warwick and Foster’s model considered the shear span-to-depth ratio and concrete strength, the
nonuniform performance of the model reveals that additional variables affected the shear behavior and must be considered.
On the other hand, it was noted that the other effectiveness factors suggested by Vecchio and Collins, Collins and Mitchell, and

13
I.S. Abbood Case Studies in Construction Materials 19 (2023) e02643

Fig. 19. Descriptive statistics of various STM in predicting shear capacity for RC deep beams without shear reinforcement [158].

Fig. 20. Descriptive statistics of various STMs in predicting shear capacity of RC deep beams with shear reinforcement [158].

Foster and Gilbert achieved very conservative results with large scatters. This is most likely because the tensile strain in concrete must
be determined according to Euler-Bernoulli’s theory (“plane sections remain plane”). Nevertheless, for RC deep beams, such a theory is
far from truth.

6. Strength of nodes

The concrete around the node is called nodal zone. The prime objective of nodal zone is to prevent the truss member from bending
due to any eccentricity in the continuous force system. The significance of these zones in design procedure is twofold. First, the stress
fields in nodal zone must be adjusted to permit forces to be transmitted safely. Second, the dimension of the nodal zone must meet
anchorage requirement-bearing plate design [159,160]. The concrete strength in nodal zones is influenced by various factors [48]
including:

• Zone confinement by support reactions, concrete strut strength, anchorage plate, and steel reinforcement.
• Strain discontinuities at nodal zone.
• The stresses created from the steel reinforcing anchorage behind the nodal zone.

The effective strength of the concrete node fce is calculated as:

fce = η f ˊc ; η ≤ 1 (3)

14
I.S. Abbood Case Studies in Construction Materials 19 (2023) e02643

where η and f ˊc are the effectiveness factor of node and concrete strength, respectively [48]. However, the strength of concrete nodes is
not significantly critical to STM [58]. The existing effectiveness factors of concrete nodes proposed by various researchers and standard
codes are summarized Table 2.

7. Tie strength and minimum reinforcements

The tie strengths identified by various code provisions are stated in Table 3. The partial safety factor of tie is typically 0.87, and it is
significantly conservative except that value of 0.70 suggested by AS 3600 [28]. Schlaich et al. [54] revealed that compressive strut has
a bowed shape, and consequently experiences transverse tensile forces. Therefore, it is crucial to provide a minimum amount of re­
inforcements to eliminate cracking of struts induced by tensile forces in order to guarantee the strut’s efficiency level indicated in
Table 1. This reinforcing significantly improves the capacity of RC deep beam in redistributing the internal forces beyond cracking as
proposed by Marti [72]. Numerical investigations [162] revealed that deep beam exhibited approximately linear elastic behavior
before cracking. To maintain the strut, adequate tension steel ties should be provided to prevent the beam from failing prematurely by
diagonal splitting failure. Foster and Gilbert [153] also reported that diagonal cracks are distributed more uniformly over the
compressive struts once enough distributing bars are provided. Furthermore, the presence of reinforcing bars decreases transverse
strains and thus improves strut efficiency. Foster and Gilbert [163] investigated the web splitting failure of an STM. It was observed

Table 2
Effectiveness factor models of concrete nodes.
Model Effectiveness factor (η) Description

Marti [72] 0.60 for general use


Collins and Mitchell [73] 0.85 CCC node
0.75 CCT node
0.60 CTT node
Schlaich et al. [54] 0.85 CCC node
0.68 CCT node
0.68 CTT node
Bergmeister et al. [159] 2.50 confined CCC node
0.76 unconfined CCC node with bearing plate
0.80 CCC node (f ˊc ≤ 27.6MPa)
0.9 − 0.25f ˊc /69 CCC node (27.6 ≤ f ˊc ≤ 69MPa)
0.65 CCC node (f ˊc ≥ 69MPa)
Jirsa et al. [161] 0.80 CCT node
0.80 CTT node
Schlaich and Schafer [53] 0.94 CCC node
0.68 CCT node
0.68 CTT node
MacGregor [154] 0.85 CCC node
0.65 CCT node
0.50 CTT node
AASHTO LRFD [20] 0.85 CCC node
0.75 CCT node
0.65 CTT node
ACI 318 [21] 0.85βn
βn = 1.00 CCC node
βn = 0.80 CCT node
βn = 0.60 CTT node
CSA A23.3 [22] 0.85 CCC node
0.75 CCT node
0.65 CTT node
CSA S806 [23] a1 ψ c CCC node
0.88a1 ψ c CCT node
a1 CTT node
NZS 3101 [25] 0.65 CCC node
0.55 CCT node
0.45 CCT node
DIN 1045–1 [26] 1.1η1 CCC node
0.75η1 CCT and CTT nodes (θ ≥ 45)
η1 = 1 normal weight concrete
η1 = 0.40 + 0.6(ρ/2200) lightweight concrete
Model Code [27] 0.85(1 − fck /250) CCC and CCT or CTT nodes (θ ≥ 55)
0.60(1 − fck /250) CCT and CTT nodes
AS 3600 [28] 0.80 − fck /200 for general use
Eurocode 2 [29] 1.00 confined CCC node
0.85 unconfined CCC node
0.70 CCT node
( )
0.70 − fck /250 > 0.5 CTT node

15
I.S. Abbood Case Studies in Construction Materials 19 (2023) e02643

Table 3
Allowable strength of ties and crack control specified by different design codes.
Model Partial safety Nominal capacity (Tie) Description Minimum crack reinforcement (crack control)
factor

AASHTO LRFD 0.87 fy Ast + Aps [fpe + fy ] - • orthogonal grid reinforcement


• spacing ≤ 12.0in.
[20]
Area Reinf.each direction
• ≥ 0.003
Gross Area Conc.
ACI 318 [21] 0.87 fy Ast + Atp [fse + Δfp ] θ ≥ 25◦ • orthogonal grid reinforcement
• spacing ≤ 12.0in.
• 0.0025/sin2 ai ≥ 0.0025
• ai ≥ 40◦
• f ˊc ≤ 6000psi
CSA A23.3 0.85 fy Ast - • orthogonal grid reinforcement
[22] • spacing ≤ 12.0in.
Area Reinf.eachdirection
• ≥ 0.002
Gross Area Conc.
CSA S806 [23] 0.65 ffu Aft only FRP bars • orthogonal grid reinforcement
• spacing ≤ 12.0in.
Area Reinf.eachdirection
• ≥ 0.002
Gross Area Conc.
CSA S6 [24] 0.85 fy Ast + Aps fpy - • orthogonal grid reinforcement
• spacing ≤ 12.0in.
Area Reinf.eachdirection
• ≥ 0.003
Gross Area Conc.
• not more 1500mm2 /m each face
NZS 3101 [25] 0.87 fy Ast + Atp [fse + Δfp ] θ ≥ 25◦ Asi
• Σ fy sin(γi ) ≥ 1.5MPa
bs si
• f ˊc ≤ 40MPa
DIN 1045–1 0.87 fyd = maximum stress of tie θ ≥ 45 ◦
Asw
• ρ = ≥ρ
[26] fp0.1k/γs = maximum stress in prestressing tie sw bw sin(α)
• ρ = 0.16(fctm /fyk )
Model Code 0.87 fytd = maximum stress of tie θ ≈ 60◦ θ ≥ 45◦ • orthogonal grid reinforcement
[27] fpyd,net = 0.9fptk/γs –σdo ≤ 600MPa = maximum stress • spacing ≤ 100mm
in prestressing tie • Ast = 0.01st bc stirrups
• Ast = 0.020st bc long. Reinforcing
• Ast = 0.015st bc long. Reinforcing (post-
tensioned regions)
AS 3600 [28] 0.70 fy Ast - -

that an increase in the minimum distributing bars corresponds to an increasing in the concrete strength. This is owing to the fact that
high-strength concrete members are often stressed to higher levels in compressive struts, exposing them to significant collapsing forces.
Under the assumption that cracked concrete holds 30% of its tensile strength, the minimum suggested distributing reinforcement
ranged from (0.2–0.4) % for concrete grade (25 ≥ f ˊc ≤ 80MPa).
As seen in Table 3, a minimum amount of crack reinforcements is recommended by all code provisions. Excluding Model Code, each

Fig. 21. Providing necessary anchorage for ties [20].

16
I.S. Abbood Case Studies in Construction Materials 19 (2023) e02643

provision is considerably specific regarding reinforcement required in terms of amount and arrangement. AASHTO, ACI 318, CSA S6,
and CSA A23.3 require, near each face, an orthogonal grid of reinforcement. AASHTO and CSA S6 require a ratio of steel reinforcement
of approximately 0.3%, and CSA A23.3 requires a reinforcement ratio of 0.2% in each direction. NZS 3101 and ACI 318 require a steel
reinforcement ratio of approximately 0.3% and 0.25%, respectively, perpendicular to the concrete strut.

8. Tie anchorage

It is necessary to ensure a safe anchorage of ties at nodes (see Fig. 21). Therefore, the minimum radius of bent and the anchorage
lengths of rebars are implemented in accordance with standard codes’ requirements. The reinforcement of tension ties should be
distributed uniformly over an effective area equal to or greater than the tie forces divided by stress limits of nodal zone. The anchorage
has to be placed inside and behind the node. It begins where the stress paths meet the rebars and are diverged. The rebar has to be
extended to the other end of node zone. Otherwise, the rebar may be extended beyond node zone if its length is less than what the code
requires. Hence, the tensile forces applied beyond the node could resist residual stresses induced within nodal zones [48].
In STM, the anchorage must be adequate and effective when the node regions completely transfer the forces. However, the question
remains in common STM regarding the anchorage requirements of whether distributing the tie reinforcing within a nodal region is
necessary [154]. Many factors could also influence the tie anchorage, such as spacing, diameter, length, and roughness of the bars, as
well as the strut and bearing plate width.

9. Strain energy minimization

In order to develop a proper truss model, it is necessary to significantly reduce strain energy. Schlaich et al. [54] claimed that it is
possible to ignore the strain energy of struts by considering ties’ energy alone in calculating the strain energy of an STM. However,
other researchers [101] have confirmed that Schlaich’s assumption is not always valid. Brown and Bayrak [101] examined the strain
energy of 596 RC deep beams (a/d ≤ 2). Two types of STMs were considered for each specimen. One model has a very simple geometry
with two inclined struts only (Fig. 22 a). The other model has a complex geometry with more struts and two vertical ties (Fig. 22 b). In
39% of cases for the first model, the strain energy in struts exceeded that in ties. In contrast, in only 5% of cases for the second model,
the strain energy in struts overcame that in ties. It has been concluded that as the complexity of an STM rises, the contribution of struts
to strain energy reduces in comparison with that of ties and it can be ignored. On the other hand, in basic models with a few struts, the
strain energy of the strut generally should not be ignored (Schlaich’s assumption is therefore not valid) [164].

10. Conclusions

STM has emerged as a promising design tool for deep concrete structural members exposed to high shear stresses, and thus it is an
alternative to commonly employed empirical models. Based on existing literature, a comprehensive review on the STM has been
presented in this paper. Accordingly, a review of the STM technique is provided with a particular focus on the basic concept of STM,
associated B- and D-regions, STM elements, STM historical development, and STM applications in RC deep beams. The review also
covers a wide range of STM aspects including strut strength, node strength, tie strength and minimum reinforcement, the effect of tie
anchorage, and strain energy minimization. Consequently, the following conclusions are drawn:

• There seems to be still no unified STM in designing RC deep beams, which is a challenge task for structural engineers in selection of
a right STM. However, researchers have recommended various guidelines and procedures to assist designers to develop their own
STMs for practical applications.
• Several authors argued that in most cases, the effectiveness factors of struts recommended by design provisions do not guarantee
sufficient levels of safety for concrete deep beams even after applying safety factors. Hence, additional experimental works are

Fig. 22. Various truss models for RC deep beams with shear reinforcement [101].

17
I.S. Abbood Case Studies in Construction Materials 19 (2023) e02643

necessary to improve the strut effectiveness factor of strut. As, the STM can be more generalized if a proper efficiency factor is
adopted.
• Limited research has investigated the RC deep beams behavior under dynamic and cyclic loading using STM. Therefore, additional
experimental works considering various geometrical limitations under various loading rates are needed to calibrate such an STM.
• Limited research has been recorded on improving the STM to investigate the serviceability behavior of RC deep beams after
cracking. Thus, such investigations are required in future.
• Lack of information has been recorded in adopting the STM for evaluating the ultimate strength of concrete deep beam reinforced
by FRP rebars. Hence, additional investigations are necessary in future.
• Future research on STM could consider the failure of shear tension in RC deep beam induced by the insufficient anchorage of
tension reinforcement.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Data availability

No data was used for the research described in the article.

References

[1] S. Fan, Y. Zhang, Y.-X. Ma, K.H. Tan, Strut-and-tie and finite element modelling of unsymmetrically-loaded deep beams, Structures 36 (2022) 805–821,
https://doi.org/10.1016/j.istruc.2021.12.037.
[2] M.V.G. Silveira, L.A.G. Bitencourt, S. Das, A performance-based optimization framework applied to a classical STM-designed deep beam, Structures 41 (2022)
488–500, https://doi.org/10.1016/j.istruc.2022.05.035.
[3] H. Karimizadeh, A. Arabzadeh, A STM-based analytical model for predicting load capacity of deep RC beams with openings, Structures 34 (2021) 1185–1200,
https://doi.org/10.1016/j.istruc.2021.08.052.
[4] K.S. Abdul-Razzaq, A.A. Dawood, Corbel strut and tie modeling – experimental verification, Structures 26 (2020) 327–339, https://doi.org/10.1016/j.
istruc.2020.04.021.
[5] P. Chetchotisak, E. Arjsri, J. Teerawong, Strut-and-tie model for shear strength prediction of RC exterior beam–column joints under seismic loading, Bull.
Earthq. Eng. 18 (4) (2020) 1525–1546, https://doi.org/10.1007/s10518-019-00756-4.
[6] S. Dey, M.M. Karthik, Modelling four-pile cap behaviour using three-dimensional compatibility strut-and-tie method, Eng. Struct. 198 (2019), 109499, https://
doi.org/10.1016/j.engstruct.2019.109499.
[7] Y.M. Yun, H.-S. Chae, J.A. Ramirez, A three-dimensional strut-and-tie model for a four-pile reinforced concrete cap, J. Adv. Concr. Technol. 17 (7) (2019)
365–380, https://doi.org/10.3151/jact.17.365.
[8] D. Tripathy, V. Singhal, Estimation of in-plane shear capacity of confined masonry walls with and without openings using strut-and-tie analysis, Eng. Struct.
188 (2019) 290–304, https://doi.org/10.1016/j.engstruct.2019.03.002.
[9] P. Chetchotisak, W. Chomchaipol, J. Teerawong, S. Shaingchin, Strut-and-tie model for predicting shear strength of squat shear walls under earthquake loads,
Eng. Struct. 256 (2022), 114042, https://doi.org/10.1016/j.engstruct.2022.114042.
[10] J. Liu, B.I. Mihaylov, A comparative study of models for shear strength of reinforced concrete deep beams, Eng. Struct. 112 (2016) 81–89, https://doi.org/
10.1016/j.engstruct.2016.01.012.
[11] M. Shahnewaz, A. Rteil, M.S. Alam, Shear strength of reinforced concrete deep beams – a review with improved model by genetic algorithm and reliability
analysis, Structures 23 (2020) 494–508, https://doi.org/10.1016/j.istruc.2019.09.006.
[12] R. Perera, J. Vique, Strut-and-tie modelling of reinforced concrete beams using genetic algorithms optimization, Constr. Build. Mater. 23 (8) (2009)
2914–2925, https://doi.org/10.1016/j.conbuildmat.2009.02.016.
[13] Y. Xia, M. Langelaar, M.A.N. Hendriks, Automated optimization-based generation and quantitative evaluation of Strut-and-tie models, Comput. Struct. 238
(2020), 106297, https://doi.org/10.1016/j.compstruc.2020.106297.
[14] J. Zhong, L. Wang, M. Zhou, Y. Li, New method for generating strut-and-tie models of three-dimensional concrete anchorage zones and box girders, J. Bridge
Eng. 22 (8) (2017), 04017047, https://doi.org/10.1061/(ASCE)BE.1943-5592.0001078.
[15] P. Nagarajan, T.M.M. Pillai, Analysis and design of simply supported deep beams using strut and tie method, Adv. Struct. Eng. 11 (5) (2008) 491–499, https://
doi.org/10.1260%2F136943308786412050.
[16] H. Chen, W.-J. Yi, H.-J. Hwang, Cracking strut-and-tie model for shear strength evaluation of reinforced concrete deep beams, Eng. Struct. 163 (2018)
396–408, https://doi.org/10.1016/j.engstruct.2018.02.077.
[17] P. Desnerck, J.M. Lees, C.T. Morley, Strut-and-tie models for deteriorated reinforced concrete half-joints, Eng. Struct. 161 (2018) 41–54, https://doi.org/
10.1016/j.engstruct.2018.01.013.
[18] J. Thomas, S. Ramadass, Improved empirical model for the strut efficiency factor and the stiffness degradation coefficient for the strength and the deflection
prediction of FRP RC deep beams, Structures 29 (2021) 2044–2066, https://doi.org/10.1016/j.istruc.2020.12.039.
[19] T.N. Tjhin, D.A. Kuchma, Integrated analysis and design tool for the strut-and-tie method, Eng. Struct. 29 (11) (2007) 3042–3052, https://doi.org/10.1016/j.
engstruct.2007.01.032.
[20] L.R.F.D. AASHTO. Bridge Design Specifications, ninth ed.,, American Association of State Highway and Transportation Officials, Washington, DC, USA, 2020,
p. 1912.
[21] ACI 318-19, Building Code Requirements For Structural Concrete And Commentary, American Concrete Institute (ACI), Farmington Hills, USA, 2019, p. 624.
[22] CSA A23.3:19. Design of Concrete Structures, seventeenth ed.,, Canadian Standards Association, Toronto, Ontario, 2019, p. 301.
[23] CSA S806-12, Design and Construction Of Building Structures With Fibre-reinforced Polymers, Canadian Standards Association, Toronto, Ontario, 2017,
p. 206.
[24] CSA S6:19. Canadian Highway Bridge Design Code, twelfth ed.,, Canadian Standards Association, Toronto, Ontario, 2019, p. 1185.
[25] NZS 3101, Concrete Structures Standard – Part 1: The Design Of Concrete Structures, Standards New Zealand, Wellington, New Zealand, 2006, p. 754.
[26] DIN 1045-1, Plain, Reinforced and Prestressed Concrete Structures – Part 1: Design and Construction, Deutsches Institut fur Normung (DIN), Berlin, Germany,
2008, p. 179.
[27] Model Code, Model Code For Concrete Structures 2010, International Federation for Structural Concrete (fib), Lausanne, Switzerland, 2012, p. 350.
[28] AS 3600:2018. Concrete Structures, fifth ed..,, Council of Standards Australia,, Sydney, NSW, Australia, 2018, p. 259.
[29] Eurocode 2, Design of Concrete Structures – Part 1-1: General Rules and Rules for Buildings, British Standards Institution, London, UK, 2015, p. 230.

18
I.S. Abbood Case Studies in Construction Materials 19 (2023) e02643

[30] K.S. Abdul-Razzaq, A. Mustafa Jalil, A.Asaad Dawood, Reinforcing struts and ties in concrete continuous deep beams, Eng. Struct. 240 (2021), 112339, https://
doi.org/10.1016/j.engstruct.2021.112339.
[31] V. Aguilar, R.W. Barnes, A. Nowak, Strength reduction factors for ACI 318 strut-and-tie method for deep beams, ACI Struct. J. 119 (2) (2022) 103–112, https://
doi.org/10.14359/51734332.
[32] A. Arabzadeh, R. Hizaji, T.Y. Yang, Experimentally studying and development of curved STM to predict the load capacity and failure mode of fixed-ended RC
deep beams, Structures 23 (2020) 289–303, https://doi.org/10.1016/j.istruc.2019.09.011.
[33] J.P. Herranz, H.S. Maria, S. Gutierrez, R. Rafael, Optimal strut-and-tie models using full homogenization optimization method, ACI Struct. J. 109 (5) (2012)
605–614, https://doi.org/10.14359/51684038.
[34] L.T. Hussein, R.M. Abbas, A semi-empirical equation based on the strut-and-tie model for the shear strength prediction of deep beams with multiple large web
openings, Eng. Technol. Appl. Sci. Res. 12 (2) (2022) 8289–8295, https://doi.org/10.48084/etasr.4743.
[35] R. Kondalraj, G.A. Rao, Efficiency of strut-and-tie model for design of reinforced concrete deep beams without web reinforcement, ACI Struct. J. 119 (3) (2022)
233–247, https://doi.org/10.14359/51734494.
[36] M.T. Ley, K.A. Riding, Widianto, S. Bae, J.E. Breen, Experimental verification of strut and tie model design method, ACI Struct. J. 104 (6) (2007) 749–755,
https://doi.org/10.14359/18957.
[37] R.T. Mabrouk, M.A. Mahmoud, M.E. Kassem, Behavior of reinforced concrete deep beams with openings under vertical loads using strut and tie model, Civ.
Eng. J. 7 (2022) 148–170, https://doi.org/10.28991/CEJ-SP2021-07-011.
[38] K.-H. Reineck, Modeling structural concrete with strut-and-tie models-summarizing discussion of the examples as per Appendix A of ACI 318-02, ACI Symp.
Publ. 208 (2002) 225–242, https://doi.org/10.14359/12423.
[39] A. Shah, E. Haq, S. Khan, Analysis and design of disturbed regions in concrete structures, Procedia Eng. 14 (2011) 3317–3324, https://doi.org/10.1016/j.
proeng.2011.07.419.
[40] A.B. Shuraim, A.K. El-Sayed, Experimental verification of strut and tie model for HSC deep beams without shear reinforcement, Eng. Struct. 117 (2016) 71–85,
https://doi.org/10.1016/j.engstruct.2016.03.002.
[41] J. Thomas, S. Ramadass, Introducing strut efficiency factor in the softened strut and tie model for the ultimate shear strength prediction of steel RC deep beams
based on experimental study, Eur. J. Environ. Civ. Eng. 26 (11) (2022) 5129–5166, https://doi.org/10.1080/19648189.2021.1885500.
[42] R.G. Tuchscherer, D.B. Birrcher, C.S. Williams, D.J. Deschenes, O. Bayrak, Evaluation of existing strut-and-tie methods and recommended improvements, ACI
Struct. J. 111 (6) (2014) 1451, https://doi.org/10.14359/51686926.
[43] T.Y. Yang, A. Amani Dashlejeh, A. Arabzadeh, R. Hizaji, New model for prediction of ultimate load of prestressed RC deep beams, Structures 23 (2020)
509–517, https://doi.org/10.1016/j.istruc.2019.12.014.
[44] W.-J. Yi, et al., Shear strength evaluation of RC D-regions based on single-panel strut-and-tie model, Eng. Struct. 265 (2022), 114500, https://doi.org/
10.1016/j.engstruct.2022.114500.
[45] J.T. Zhong, L. Wang, P. Deng, M. Zhou, A new evaluation procedure for the strut-and-tie models of the disturbed regions of reinforced concrete structures, Eng.
Struct. 148 (2017) 660–672, https://doi.org/10.1016/j.engstruct.2017.07.012.
[46] A. Ashour, K.-H. Yang, Application of plasticity theory to reinforced concrete deep beams: a review, Mag. Concr. Res. 60 (9) (2008) 657–664, https://doi.org/
10.1680/macr.2008.00038.
[47] M. Panjehpour, A.A.A. Ali, A. Mohammed Parvez, F.N. Aznieta, Y.L. Voo, An overview of strut-and-tie model and its common challenges, Int. J. Eng. Res. Afr. 8
(2012) 37–45, https://doi.org/10.4028/www.scientific.net/JERA.8.37.
[48] R.K.L. Su, A.M. Chandler, Design criteria for unified strut and tie models, Prog. Struct. Eng. Mater. 3 (3) (2001) 288–298, https://doi.org/10.1002/pse.88.
[49] Y. Xia, M. Langelaar, M.A.N. Hendriks, Optimization-based strut-and-tie model generation for reinforced concrete structures under multiple load conditions,
Eng. Struct. 266 (2022), 114501, https://doi.org/10.1016/j.engstruct.2022.114501.
[50] R.D. Cook, W.C. Young. Advanced Mechanics of Materials, second ed.,, Prentice Hall, NJ, USA, 1999, p. 504.
[51] S.E.-D.E. El-Metwally, For a consistent treatment of shear design of structural concrete B-regions, Struct. Eng. Rev. 7 (4) (1995) 323–335, https://doi.org/
10.1016%2F0952-5807(95)00008-3.
[52] S. Taufik, A. Sugianto, Analysing numerical modelling behaviour of reinforced concrete deep beam with strut-and-tie model, New Approaches Eng. Res. 12
(2021) 17–34, https://doi.org/10.9734/bpi/naer/v12/3867F.
[53] J. Schlaich, K. Schafer, Design and detailing of structural concrete using strut-and-tie models, Struct. Eng. 69 (6) (1991) 113–125.
[54] J. Schlaich, K. Schafer, M. Jennewein, Toward a consistent design of structural concrete, PCI J. 32 (3) (1987) 74–150, https://doi.org/10.15554/
pcij.05011987.74.150.
[55] S.E.-D.E. El-Metwally, W.-F. Chen. Structural Concrete: Strut-and-tie Models for Unified Design, first ed.,, CRC Press, FL, USA, 2017.
[56] A. Starčev-Ćurčin, et al., Experimental testing of reinforced concrete deep beams designed by strut-and-tie method, Appl. Sci. 10 (18) (2020) 6217, https://doi.
org/10.3390/app10186217.
[57] A. Alshegeir, J.A. Ramirez, Computer graphics in detailing strut-tie models, J. Comput. Civ. Eng. 6 (2) (1992) 220–232, https://doi.org/10.1061/(ASCE)0887-
3801(1992)6:2(220).
[58] M.D. Brown, Design for shear in reinforced concrete using strut-and-tie and sectional models. Doctoral Thesis, University of Texas at Austin, TX, USA, 2005.
[59] S.M. Lopes, R.N.F. do Carmo, Deformable strut and tie model for the calculation of the plastic rotation capacity, Comput. Struct. 84 (31) (2006) 2174–2183,
https://doi.org/10.1016/j.compstruc.2006.08.028.
[60] M.E. El-Zoughiby, Z-shaped load path: a unifying approach to developing strut-and-tie models, ACI Struct. J. 118 (3) (2021) 35–45, https://doi.org/10.14359/
51730535.
[61] D.B. Garber, Shear cracking in inverted-T straddle bents. Master Thesis, University of Texas at Austin, TX, USA, 2011.
[62] P. Chetchotisak, J. Teerawong, S. Yindeesuk, Modified interactive strut-and-tie modeling of reinforced concrete deep beams and corbels, Structures 45 (2022)
284–298, https://doi.org/10.1016/j.istruc.2022.08.116.
[63] W. Ritter, The hennebique design method (Die bauweise hennebique), Schweiz. Bauztg. 33 (7) (1899) 59–61.
[64] E. Morsch. Der betoneisenbau, seine anwendung und theorie (1902), German, Kessinger Publishing, USA, 2010, p. 120.
[65] A.N. Talbot, Tests of reinforced concrete beams: resistance to web stresses; Report No. 29, University of Illinois at Urbana Champaign, USA, 1909, 〈http://hdl.
handle.net/2142/4532〉.
[66] F.E. Richart, Investigation of web stresses in reinforced concrete beams; Report No. 166, University of Illinois at Urbana Champaign, USA, 1927, 〈http://hdl.
handle.net/2142/4391〉.
[67] E. Rausch, Design of Reinforced Concrete For Torsion And Shear, 50, Julius Springer-Verlag, Berlin, Germany, 1929.
[68] H. Kupfer, Expansion of Morsch’s truss analogy by application of the principle of minimum strain energy, CEB Bull. 40 (1964).
[69] F. Leondardt, Reducing the shear reinforcement in reinforced concrete beams and slabs, Mag. Concr. Res. 17 (53) (1965) 187–198, https://doi.org/10.1680/
macr.1965.17.53.187.
[70] P. Müller, Schlusswort und zusammenfassung, in: Plastische berechnung von stahlbetonscheiben und -balken, Vol. 83, Springer, Basel AG, Switzerland, 1978,
pp. 145–147, https://doi.org/10.1007/978-3-0348-5761-1_12.
[71] M.P. Nielsen, L.C. Hoang. Limit Analysis And Concrete Plasticity, Third ed.,, CRC Press, FL, USA, 2016, p. 816.
[72] P. Marti, Basic tools of reinforced concrete beam design, ACI J. Proc. 82 (1) (1985) 46–56, https://doi.org/10.14359/10314.
[73] M.P. Collins, D. Mitchell, Rational approach to shear design-the 1984 Canadian code provisions, ACI J. Proc. 83 (6) (1986) 925–933, https://doi.org/
10.14359/3031.
[74] D.M. Rogowsky, J.G. MacGregor, Design of reinforced concrete deep beams, Concr. Int. 8 (8) (1986) 49–58.
[75] J.A. Ramirez, J.E. Breen, Evaluation of a modified truss-model approach for beams in shear, ACI Struct. J. 88 (5) (1991) 562–572, https://doi.org/10.14359/
3145.

19
I.S. Abbood Case Studies in Construction Materials 19 (2023) e02643

[76] L. Zhou, S. Wan, Full-range nonlinear analysis of post-tensioned anchorage zones based on modified strut-and-tie model, Structures 35 (2022) 565–576,
https://doi.org/10.1016/j.istruc.2021.10.073.
[77] A. Alshegeir, J.A. Ramirez, Strut-tie approach in pretensioned beams, ACI Struct. J. 89 (3) (1992) 296–304, https://doi.org/10.14359/3240.
[78] M.P. Collins, D. Mitchell, Prestressed Concrete Structures, Prentice Hall, NJ, USA, 2001, p. 766.
[79] M.R. Khalaf, A.H.A. Al-Ahmed, Shear strength of reinforced concrete deep beams with large openings strengthened by external prestressed strands, Structures
28 (2020) 1060–1076, https://doi.org/10.1016/j.istruc.2020.09.052.
[80] Q.Q. Liang, Y.M. Xie, G. Steven, P, Generating optimal strut-and-tie models in prestressed concrete beams by performance-based optimization, ACI Struct. J. 98
(2) (2001) 226–232, https://doi.org/10.14359/10191.
[81] J.A. Ramirez, Strut-tie shear design of pretensioned concrete, ACI Struct. J. 91 (5) (1994) 572–578, https://doi.org/10.14359/4174.
[82] D.H. Sanders, J.E. Breen, Post-tensioned anchorage zones with single straight concentric anchorages, ACI Struct. J. 94 (2) (1997) 146–158, https://doi.org/
10.14359/469.
[83] K.H. Tan, A.E. Naaman, Strut-and-tie model for externally prestressed concrete beams, ACI Struct. J. 90 (6) (1993) 683–691, https://doi.org/10.14359/4492.
[84] K.H. Tan, K. Tong, C.Y. Tang, Direct strut-and-tie model for prestressed deep beams, J. Struct. Eng. 127 (9) (2001) 1076–1084, https://doi.org/10.1061/
(ASCE)0733-9445(2001)127:9(1076).
[85] G.-L. Wang, S.-P. Meng, Modified strut-and-tie model for prestressed concrete deep beams, Eng. Struct. 30 (12) (2008) 3489–3496, https://doi.org/10.1016/j.
engstruct.2008.05.020.
[86] W.B. Siao, Strut-and-tie model for shear behavior in deep beams and pile caps failing in diagonal splitting, ACI Struct. J. 90 (4) (1993) 356–363, https://doi.
org/10.14359/3978.
[87] A. Alshegeir, Analysis and design of disturbed regions with strut-tie models. Doctoral Thesis, Purdue University, USA, 1992.
[88] K.H. Tan, L.W. Weng, S. Teng, A strut-and-tie model for deep beams subjected to combined top-and-bottom loading, Struct. Eng. 75 (13) (1997) 215–223.
[89] A.-A. 445, Recent approaches to shear design of structural concrete, J. Struct. Eng. 124 (12) (1998) 1375–1417, https://doi.org/10.1061/(ASCE)0733-9445
(1998)124:12(1375).
[90] S.-J. Hwang, W.-Y. Lu, H.-J. Lee, Shear strength prediction for deep beams, ACI Struct. J. 97 (3) (2000) 367–376, https://doi.org/10.14359/9624.
[91] S.-J. Hwang, H.-J. Lee, Strength prediction for discontinuity regions by softened strut-and-tie model, J. Struct. Eng. 128 (12) (2002) 1519–1526, https://doi.
org/10.1061/(ASCE)0733-9445(2002)128:12(1519).
[92] W.-Y. Lu, I.-J. Lin, H.-W. Yu, Shear strength of reinforced concrete deep beams, ACI Struct. J. 110 (4) (2013) 671–680, https://doi.org/10.14359/51685752.
[93] K.H. Tan, C.Y. Tang, K. Tong, A direct method for deep beams with web reinforcement, Mag. Concr. Res. 55 (1) (2003) 53–63, https://doi.org/10.1680/
macr.2003.55.1.53.
[94] K.H. Tan, K. Tong, C.Y. Tang, Consistent strut-and-tie modelling of deep beams with web openings, Mag. Concr. Res. 55 (1) (2003) 65–75, https://doi.org/
10.1680/macr.2003.55.1.65.
[95] A.B. Matamoros, K.H. Wong, Design of simply supported deep beams using strut-and-tie models, Acids Struct. J. 100 (6) (2003) 704–712, https://doi.org/
10.14359/12836.
[96] G. Russo, R. Venir, M. Pauletta, Reinforced concrete deep beams- shear strength model and design formula, Acids Struct. J. 102 (3) (2005) 429–437, https://
doi.org/10.14359/14414.
[97] P. Chetchotisak, J. Teerawong, S. Yindeesuk, J. Song, New strut-and-tie-models for shear strength prediction and design of RC deep beams, Comput. Concr. 14
(1) (2014) 19–40, https://doi.org/10.12989/cac.2014.14.1.019.
[98] C.Y. Tang, K.H. Tan, Interactive mechanical model for shear strength of deep beams, J. Struct. Eng. 130 (10) (2004) 1534–1544, https://doi.org/10.1061/
(ASCE)0733-9445(2004)130:10(1534).
[99] K.H. Tan, G.H. Cheng, Size effect on shear strength of deep beams: Investigating with strut-and-tie model, J. Struct. Eng. 132 (5) (2006) 673–685, https://doi.
org/10.1061/(ASCE)0733-9445(2006)132:5(673).
[100] N. Zhang, K.-H. Tan, Size effect in RC deep beams: experimental investigation and STM verification, Eng. Struct. 29 (12) (2007) 3241–3254, https://doi.org/
10.1016/j.engstruct.2007.10.005.
[101] M.D. Brown, O. Bayrak, Design of deep beams using strut-and-tie models - part I: evaluating US provisions, ACI Struct. J. 105 (4) (2008) 395–404, https://doi.
org/10.14359/19853.
[102] M.D. Brown, O. Bayrak, Design of deep beams using strut-and-tie models - part II: design recommendations, ACI Struct. J. 105 (4) (2008) 405–413, https://doi.
org/10.14359/19854.
[103] N. Zhang, K.-H. Tan, C.-L. Leong, Single-span deep beams subjected to unsymmetrical loads, J. Struct. Eng. 135 (3) (2009) 239–252, https://doi.org/10.1061/
(ASCE)0733-9445(2009)135:3(239).
[104] K.-H. Yang, F.Ashour Ashraf, Strut-and-tie model based on crack band theory for deep beams, J. Struct. Eng. 137 (10) (2011) 1030–1038, https://doi.org/
10.1061/(ASCE)ST.1943-541X.0000351.
[105] E. Lim, S.-J. Hwang, Modeling of the strut-and-tie parameters of deep beams for shear strength prediction, Eng. Struct. 108 (2016) 104–112, https://doi.org/
10.1016/j.engstruct.2015.11.024.
[106] H. Chen, W.-J. Yi, Z.J. Ma, H.-J. Hwang, Modeling of shear mechanisms and strength of concrete deep beams reinforced with FRP bars, Compos. Struct. 234
(2020), 111715, https://doi.org/10.1016/j.compstruct.2019.111715.
[107] I.S. Abbood, Sa Odaa, K.F. Hasan, M.A. Jasim, Properties evaluation of fiber reinforced polymers and their constituent materials used in structures – a review,
Mater. Today Proc. 43 (2021) 1003–1008, https://doi.org/10.1016/j.matpr.2020.07.636.
[108] S.S. Weli, I.S. Abbood, K.F. Hasan, M.A. Jasim, Effect of steel fibers on the concrete strength grade: a review, IOP Conf. Ser.: Mater. Sci. Eng. 888 (2020),
012043, https://doi.org/10.1088/1757-899X/888/1/012043.
[109] A. Godat, O. Chaallal, Strut-and-tie method for externally bonded FRP shear-strengthened large-scale RC beams, Compos. Struct. 99 (2013) 327–338, https://
doi.org/10.1016/j.compstruct.2012.11.034.
[110] A.N. Hanoon, Modified strut-and-tie models for reinforced concrete deep beams with externally bonded CFRP systems. Doctoral Thesis, Universiti Putra
Malaysia, Malaysia, 2017.
[111] D.-J. Kim, J. Lee, Y.H. Lee, Effectiveness factor of strut-and-tie model for concrete deep beams reinforced with FRP rebars, Compos. Part B: Eng. 56 (2014)
117–125, https://doi.org/10.1016/j.compositesb.2013.08.009.
[112] S. Liu, Analysis of concrete deep beams with fibre reinforced polymer reinforcements using indeterminate strut-and-tie method. Master Thesis, University of
Waterloo, Ontario, Canada, 2022.
[113] S. Liu, M.A. Polak, Estimating shear strengths of GFRP reinforced concrete deep beams with indeterminate strut-and-tie method. Current Perspectives and New
Directions in Mechanics, Modelling and Design of Structural Systems, first ed.,, CRC Press, London, UK, 2022 pp 511-512.
[114] K. Mohamed, A.S. Farghaly, B. Benmokrane, Strut efficiency-based design for concrete deep beams reinforced with fiber-reinforced polymer bars, ACI Struct. J.
113 (4) (2016) 791–800, https://doi.org/10.14359/51688476.
[115] K. Mohamed, A.S. Farghaly, B. Benmokrane, K.W. Neale, Nonlinear finite-element analysis for the behavior prediction and strut efficiency factor of GFRP-
reinforced concrete deep beams, Eng. Struct. 137 (2017) 145–161, https://doi.org/10.1016/j.engstruct.2017.01.045.
[116] T.S. Mustafa, F.B.A. Beshara, A.S. Abd El-Maula, M.G. Fathi, Strut-and-tie model for FRP effectiveness in shear strengthening of RC deep beams, Eur. J.
Environ. Civ. Eng. (2022) 1–16, https://doi.org/10.1080/19648189.2022.2056248.
[117] M. Nehdi, Z. Omeman, H. El-Chabib, Optimal efficiency factor in strut-and-tie model for FRP-reinforced concrete short beams with (1.5 < a/d < 2.5), Mater.
Struct. 41 (10) (2008) 1713–1727, https://doi.org/10.1617/s11527-008-9359-9.
[118] M. Panjehpour, A.A.A. Ali, Y.L. Voo, F.N. Aznieta, Modification of strut effectiveness factor for reinforced concrete deep beams strengthened with CFRP
laminates, Mater. De Constr. 64 (314) (2014) 016, https://doi.org/10.3989/mc.2014.02913.

20
I.S. Abbood Case Studies in Construction Materials 19 (2023) e02643

[119] S. Park, S.Aboutaha Riyad, Strut-and-tie method for CFRP strengthened deep RC members, J. Struct. Eng. 135 (6) (2009) 632–643, https://doi.org/10.1061/
(ASCE)0733-9445(2009)135:6(632).
[120] A.N. Hanoon, M.S. Jaafar, S.R. Al Zaidee, F. Hejazi, F.N.A.A. Aziz, Effectiveness factor of the strut-and-tie model for reinforced concrete deep beams
strengthened with CFRP sheet, J. Build. Eng. 12 (2017) 8–16, https://doi.org/10.1016/j.jobe.2017.05.001.
[121] M. Deng, F. Ma, W. Ye, X. Liang, Investigation of the shear strength of HDC deep beams based on a modified direct strut-and-tie model, Constr. Build. Mater.
172 (2018) 340–348, https://doi.org/10.1016/j.conbuildmat.2018.03.274.
[122] M. Moradi, M.R. Esfahani, Application of the strut-and-tie method for steel fiber reinforced concrete deep beams, Constr. Build. Mater. 131 (2017) 423–437,
https://doi.org/10.1016/j.conbuildmat.2016.11.042.
[123] M.S.V. Sagi, C. Lakavath, S.S. Prakash, Effect of steel fibers on the shear behavior of Self-Compacting reinforced concrete deep Beams: an experimental
investigation and analytical model, Eng. Struct. 269 (2022), 114802, https://doi.org/10.1016/j.engstruct.2022.114802.
[124] S.D. Adhikary, B. Li, K. Fujikake, Strength and behavior in shear of reinforced concrete deep beams under dynamic loading conditions, Nucl. Eng. Des. 259
(2013) 14–28, https://doi.org/10.1016/j.nucengdes.2013.02.016.
[125] A.N. Hanoon, M.S. Jaafar, F. Hejazi, F.N.A.A. Aziz, Strut effectiveness factor for reinforced concrete deep beams under dynamic loading conditions, Case Stud.
Struct. Eng. 6 (2016) 84–102, https://doi.org/10.1016/j.csse.2016.08.001.
[126] M. Shahnewaz, Shear behavior of reinforced concrete deep beams under static and dynamic loads. Master Thesis, University of British Columbia Vancouver,
BC, Canada, 2013.
[127] A. Somraj, K. Fujikake, B. Li, Influence of loading rate on shear capacity of reinforced concrete beams, Int. J. Prot. Struct. 4 (4) (2013) 521–543, https://doi.
org/10.1260/2041-4196.4.4.521.
[128] S.M. Alcocer, C.M. Uribe, Monolithic and cyclic behavior of deep beams designed using strut-and-tie models, Acids Struct. J. 105 (3) (2008) 327–337, https://
doi.org/10.14359/19792.
[129] A. Arabzadeh, R. Hizaji, A simple approach to predict the shear capacity and failure mode of fix-ended reinforced concrete deep beams based on experimental
study, Int. J. Eng. 32 (4) (2019) 474–483, https://doi.org/10.5829/ije.2019.32.04a.03.
[130] B. Mihaylov, Behavior of deep reinforced concrete beams under monotonic and reversed cyclic load. Doctoral Thesis, Università degli Studi di Pavia, Pavia,
Italy, 2008.
[131] R. Perera, J. Vique, A. Arteaga, A.D. Diego, Shear capacity of reinforced concrete members strengthened in shear with FRP by using strut-and-tie models and
genetic algorithms, Compos. Part B Eng. 40 (8) (2009) 714–726, https://doi.org/10.1016/j.compositesb.2009.06.008.
[132] H. Hardjasaputra, Evolutionary structural optimization as tool in finding strut-and-tie-models for designing reinforced concrete deep beam, Procedia Eng. 125
(2015) 995–1000, https://doi.org/10.1016/j.proeng.2015.11.153.
[133] G. Yavuz, Shear strength estimation of RC deep beams using the ANN and strut-and-tie approaches, Struct. Eng. Mech. 57 (4) (2016) 657–680, https://doi.org/
10.12989/sem.2016.57.4.657.
[134] M.K. Dhahir, Shear strength of FRP reinforced deep beams without web reinforcement, Compos. Struct. 165 (2017) 223–232, https://doi.org/10.1016/j.
compstruct.2017.01.039.
[135] M.K. Dhahir, Strut and tie modeling of deep beams shear strengthened with FRP laminates, Compos. Struct. 193 (2018) 247–259, https://doi.org/10.1016/j.
compstruct.2018.03.073.
[136] A.N. Hanoon, M.S. Jaafar, F. Hejazi, F.N.A.A. Aziz, Strut-and-tie model for externally bonded CFRP-strengthened reinforced concrete deep beams based on
particle swarm optimization algorithm: CFRP debonding and rupture, Constr. Build. Mater. 147 (2017) 428–447, https://doi.org/10.1016/j.
conbuildmat.2017.04.094.
[137] J. Zaborac, J. Choi, O. Bayrak, Assessment of deep beams with inadequate web reinforcement using strut-and-tie models, Eng. Struct. 218 (2020), 110832,
https://doi.org/10.1016/j.engstruct.2020.110832.
[138] M.S. Lourenco, J.F. Almeida, Adaptive stress field models: formulation and validation, ACI Struct. J. 110 (1) (2013) 71–82, https://doi.org/10.14359/
51684331.
[139] R.M. Scott, J.B. Mander, J.M. Bracci, Compatibility strut-and-tie modeling: part I—formulation, ACI Struct. J. 109 (5) (2012) 635–644, https://doi.org/
10.14359/51684041.
[140] L. Zhou, Z. Liu, Z. He, Elastic-to-plastic strut-and-tie model for concentric anchorage zones, J. Bridge Eng. 22 (10) (2017), 04017070, https://doi.org/10.1061/
(ASCE)BE.1943-5592.0001060.
[141] L. Zhou, Z. Liu, Z. He, Elastic-to-plastic strut-and-tie model for deep beams, J. Bridge Eng. 23 (4) (2018), 04018007, https://doi.org/10.1061/(ASCE)BE.1943-
5592.0001206.
[142] M.D. Brown, C.L. Sankovich, O. Bayark, J.O. Jirsa, Behavior and efficiency of bottle-shaped struts, ACI Struct. J. 103 (3) (2006) 348–354, https://doi.org/
10.14359/15312.
[143] F.J. Vecchio, M.P. Collins, The modified compression-field theory for reinforced concrete elements subjected to shear, ACI J. Proc. 83 (2) (1986) 219–231,
https://doi.org/10.14359/10416.
[144] S.J. Foster, A.R. Malik, Evaluation of efficiency factor models used in strut-and-tie modeling of nonflexural members, J. Struct. Eng. 128 (5) (2002) 569–577,
https://doi.org/10.1061/(ASCE)0733-9445(2002)128:5(569).
[145] Y.M. Yun, J.A. Ramirez, Strength of concrete struts in three-dimensional strut-tie models, J. Struct. Eng. 142 (11) (2016), 04016117, https://doi.org/10.1061/
(ASCE)ST.1943-541X.0001584.
[146] B. Bresler, J.G. MacGregor, Review of concrete beams failing in shear, J. Struct. Div. 93 (1) (1967) 343–372, https://doi.org/10.1061/JSDEAG.0001586.
[147] J.-w Park, D. Kuchma, Strut-and-tie model analysis for strength prediction of deep beams, ACI Struct. J. 104 (6) (2007) 657–666, https://doi.org/10.14359/
18947.
[148] C.G. Quintero-Febres, G. Parra-Montesinos, J.K. Wight, Strength of struts in deep concrete members designed using strut-and-tie method, ACI Struct. J. 103 (4)
(2006) 577–586, https://doi.org/10.14359/16434.
[149] N. Zhang, K.-H. Tan, Direct strut-and-tie model for single span and continuous deep beams, Eng. Struct. 29 (11) (2007) 2987–3001, https://doi.org/10.1016/j.
engstruct.2007.02.004.
[150] K. Bergmeister, J.E. Breen, J.O. Jirsa, M.E. Kreger, Detailing in Structural Concrete; Center for Transportation Research (CTR), University of Texas at Austin,
TX, USA, 1993.
[151] F.J. Vecchio, M.P. Collins, Compression response of cracked reinforced concrete, J. Struct. Eng. 119 (12) (1993) 3590–3610, https://doi.org/10.1061/(ASCE)
0733-9445(1993)119:12(3590).
[152] W.B. Warwick, S.J. Foster, Investigation into the efficiency factor used in non-flexural reinforced concrete member design; UNICIV Report, University of New
South Wales, Australia, 1993.
[153] S.J. Foster, R.I. Gilbert, The design of nonflexural members with normal and high-strength concretes, ACI Struct. J. 93 (1) (1996) 3–10, https://doi.org/
10.14359/9671.
[154] J.G. MacGregor. Reinforced Concrete: Mechanics And Design, third ed.,, Prentice Hall, NJ, USA, 1997.
[155] W. Kaufmann, P. Marti, Structural concrete: cracked membrane model, J. Struct. Eng. 124 (12) (1998) 1467–1475, https://doi.org/10.1061/(ASCE)0733-9445
(1998)124:12(1467).
[156] L.-X.B. Zhang, T.T.C. Hsu, Behavior and analysis of 100 MPa concrete membrane elements, J. Struct. Eng. 124 (1) (1998) 24–34, https://doi.org/10.1061/
(ASCE)0733-9445(1998)124:1(24).
[157] D. Zwicky, T. Vogel, Critical inclination of compression struts in concrete beams, J. Struct. Eng. 132 (5) (2006) 686–693, https://doi.org/10.1061/(ASCE)
0733-9445(2006)132:5(686).
[158] K.S. Ismail, M. Guadagnini, K. Pilakoutas, Strut-and-tie modeling of reinforced concrete deep beams, J. Struct. Eng. 144 (2) (2018), 04017216, https://doi.org/
10.1061/(ASCE)ST.1943-541X.0001974.

21
I.S. Abbood Case Studies in Construction Materials 19 (2023) e02643

[159] K. Bergmeister, J.E. Breen, J.O. Jirsa, Dimensioning of the Nodes and Development of Reinforcement, International Association for Bridge and Structural
Engineering, Zurich, 1991.
[160] Y.M. Yun, Strength of two-dimensional nodal zones in strut–tie models, J. Struct. Eng. 132 (11) (2006) 1764–1783, https://doi.org/10.1061/(ASCE)0733-
9445(2006)132:11(1764).
[161] J.O. Jirsa, et al., Experimental studies of nodes in strut-and-tie models, IABSE Rep. Struct. Concr. 62 (1991) 525–532, https://doi.org/10.5169/seals-47681.
[162] S.J. Foster, The structural behaviour of reinforced concrete deep beams. Doctorate Thesis, University of New South Wales, Sydney NSW, Australia, 1992.
[163] S.J. Foster, R.I. Gilbert, Strut and tie modelling of non-flexural members, Aust. Civ. Eng. Trans. 39 (2/3) (1997) 87–94.
[164] F. Angotti, M. Guiglia, P. Marro, M. Orlando, Strut-and-tie models, in: F. Angotti, et al. (Eds.), Reinforced Concrete with Worked Examples, Springer
International Publishing, Cham, 2022, pp. 587–691, https://doi.org/10.1007/978-3-030-92839-1_10.

22

You might also like