You are on page 1of 27

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/360772055

A cohesive fracture model for predicting crack spacing and crack width in
reinforced concrete structures

Article in Engineering Failure Analysis · May 2022


DOI: 10.1016/j.engfailanal.2022.106452

CITATIONS READS

36 157

5 authors, including:

Umberto De Maio Fabrizio Greco


Università della Calabria Università della Calabria
27 PUBLICATIONS 484 CITATIONS 132 PUBLICATIONS 3,126 CITATIONS

SEE PROFILE SEE PROFILE

Lorenzo Leonetti Paolo Nevone Blasi


Università della Calabria Università della Calabria
55 PUBLICATIONS 1,425 CITATIONS 38 PUBLICATIONS 702 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Andrea Pranno on 22 July 2022.

The user has requested enhancement of the downloaded file.


Preprint version of the paper: U. De Maio, F. Greco, L. Leonetti, P. Nevone Blasi, A. Pranno, A
cohesive fracture model for predicting crack spacing and crack width in reinforced concrete structures,
Engineering Failure Analysis. 139 (2022) 106452. https://doi.org/10.1016/j.engfailanal.2022.106452.

1 A cohesive fracture model for predicting crack spacing and crack width
2 in reinforced concrete structures
3 Umberto DE MAIO, Fabrizio GRECO*, Lorenzo LEONETTI, Paolo NEVONE BLASI,
4 Andrea PRANNO

5 Department of Civil Engineering, University of Calabria, Via P. Bucci Cubo39B, Rende 87036, Italy.
6 * Corresponding author: fabrizio.greco@unical.it

7 Abstract.

8 The cracking behavior analysis, in terms of crack spacing and crack width prediction, is essential to assess the
9 damage and durability of a reinforced concrete (RC) structure. However, available numerical models cannot
10 give satisfactory results in the cracking prediction due to the continuum representation of the concrete material
11 often adopted in the numerical framework. Based on the cohesive finite element method, an interelement
12 fracture model has been proposed here to investigate the cracking behavior, as well as its influence on the load-
13 carrying capacity, of RC structural elements under static loading conditions. This fracture model is employed
14 together with an embedded truss model for the rebars allowing them to be crossed by the cracks. Cracking
15 analyses are performed on RC members subjected to tension and flexural loads. Comparisons with available
16 experimental and numerical results have highlighted the effectiveness of the proposed model to investigate the
17 cracking behavior in RC structures.

18

19 Keywords. Crack spacing and crack width, reinforced concrete structures, cohesive fracture model, embedded
20 truss model.

21

22 1. Introduction

23 The cracking behavior, in terms of crack width and crack spacing, plays a crucial role in reducing both stiffness
24 and load-carrying capacity of the reinforced concrete (RC) structures subjected to static and dynamic loading
25 conditions, compromising their design requisites at the serviceability and ultimate limit states. Moreover,
26 concrete cracking strongly affects the structural integrity of the RC components favoring the development of
27 deterioration owing to chemical processes, such as alkali-aggregate reactions, leaching, and delayed ettringite
28 formation, as well as the steel reinforcement corrosion and the freeze-thaw damage process [1].
29 In order to control the cracking phenomena in concrete structures, formulas based on semi-theoretical and
30 semi-empirical models, are reported in the existing design codes, such as CEB/FIP Model Code 2010 [2],
31 Eurocode 2 [3], and American Concrete Institute (ACI) code [4], to calculate maximum and average values of
32 crack width and spacing. However, such formulas, containing coefficients determined from experiments or
33 experience, provide highly scattered predictions, reaching in some cases results with high percentage errors
34 with respect to the real cracking conditions [5,6]. These issues are mainly due to the complexity of the nonlinear
35 mechanisms which arise in the RC members, associated with the concrete microstructure as well as the
36 magnitude difference between the mechanical properties of concrete and reinforcing bars and their interaction,
37 making it difficult to establish consistent and accurate cracking analysis procedures. Indeed, the general
38 stochastic structure of aggregates within the concrete, which induces stochastic crack formation and
39 propagation, adds more complexity to the cracking behavior of concrete [7]. One of the most cracking
40 influencing factors in the RC structures is the bond behavior between steel rebars and surrounding concrete
41 because it affects the capacity of the undamaged concrete to carry tensile forces between adjacent primary
42 cracks thus providing a higher overall stiffness with respect to that obtained only by steel bars. Such
43 mechanism, commonly referred to as the tension stiffening effect, improves the load-deformation response of
44 RC structures under flexural loading conditions generally characterized by the well-known three loading stages
45 coinciding with the elastic stage, the crack formation stage, and the stabilized cracking stage occurring before
46 the rebars yielding, and mostly depend on bond characteristics of the concrete/rebars interface as reported in
47 some available experimental works [8,9]. An improvement of the concrete mechanical behavior, in terms of
48 both tension stiffening effect and cracking characteristics, was obtained by adding steel fibers and carbon
49 nanotubes within the concrete matrix, detecting a decrease in the average crack spacing and maximum crack
50 width at the increasing of the fiber content [10,11]. Moreover, some experimental results on the cracking
51 behavior of RC members reveal as the rebar type (Glass-FRP or steel bars) and the surface characteristics,
52 influence the concrete/rebar interface behavior providing beneficial effects on both crack width and spacing
53 when steel bars are used [12]. Moreover, a suitable arrangement of the reinforcing bars improves the flexural
54 stiffness of RC beams but a clear correlation between crack width and spacing is not detected when the tensile
55 rebar layout changes [6]. The use of high-strength steel rebars in concrete members allows higher axial stresses
56 to be developed than conventional bars and provides crack widths, evaluated at service load levels, within the
57 limits reported in the main design codes [13,14]. In some experimental tests, the cracking behavior of UHPC
58 members was mainly affected by the reinforcement ratio and the UHPC cover thickness, showing that the
59 diameter of rebars plays little role in the reduction of both crack spacing and maximum crack width [9,15].
60 The bonding of composite materials along the tensile faces of structural elements is a good practice to improve
61 the overall stiffness and the cracking behavior in RC and masonry structures as highlighted by some numerical
62 and experimental results [16–19]. Recent experimental works have investigated the damage processes and the
63 mechanical bond behavior of galvanized steel fabric-reinforced cementitious matrix (SFRCM) laminated
64 through non-conventional experimental setups, assessing the force-slip distribution and the crack opening
65 displacement by the DIC optical system monitoring [20,21].
66 On the other hand, a great effort has been made by the researchers to develop numerical and analytical models
67 able to accurately predict the cracking behavior with the aim of ensuring the serviceability requirements for
68 RC structures under general loading conditions. Earlier studies focused to describe the bond-slip mechanisms,
69 induced by the nonlinear cracking processes, and consequently, the tension stiffening effect in tensile RC
70 members through analytical formulations, based on fracture energy and bond stress-slip approaches [22,23].
71 However, these analytical models can be used under simplified assumptions, such as particular boundary
72 conditions and uniformly distributed cracking, in order to satisfy the requirement of the continuous
73 displacement field. In light of the analytical modeling limitations, numerical modeling has been mostly
74 employed by researchers for crack simulations. Among the main numerical modeling techniques, such as the
75 finite element method (FEM) [24], discrete element method (DEM) [25], and boundary element method
76 (BEM) [26], the FEM had received the most research interest in simulating the cracking processes in quasi-
77 brittle materials. In this contest, the damage analysis can be performed by using smeared and discrete fracture
78 approaches, according to which the damage is either continuously distributed over the computational domain
79 by a reduction of the material integrity or lumped in a discontinuity injected in the displacement field,
80 respectively [27]. However, the smeared crack models, which provide very accurate global structural responses
81 of concrete or masonry structures in the strain-softening regimes [28,29], are unable to record crack initiation
82 and propagation as well as the coalescence and branching phenomena since their essential characteristics are
83 inevitably lost in the smoothing process unless some post‐processing procedures are introduced to extract crack
84 geometries [30]. Numerical models based on the continuum mechanics are employed to capture the effects of
85 nanomaterials and the microstructure of the materials on the load-carrying capacity of the structures [31–33].
86 On the other hand, the discrete fracture models easily simulate the entire crack process by using inter- and
87 intra-element cohesive modeling techniques [34–36]. These models, also employed to simulate the interfacial
88 crack in adhesive-bonded ductile sheets [37], use various phenomenological-type traction-separation laws,
89 depending on the simulated material, to describe the mechanical behavior of traction forces acting over the
90 crack faces, but suffer from mesh dependency issues partially solved by introducing mesh-updating procedures
91 or moving mesh strategies [38,39]. To correctly evaluate the cracking behavior of RC structures, in terms of
92 crack width and crack spacing, a suitable concrete/rebars interaction modeling is needed. The simpler modeling
93 does not allow relative slip between rebars and surrounding concrete thus implying full compatibility between
94 concrete and reinforcement strains [40,41]. Such modeling provides a distributed damage in the critical tensile
95 zone rather than localized as occurs in real structures and a significant deviation of the crack width predictions
96 with respect to the experimental outcomes [42]. More sophisticated crack models employ interface elements
97 and discrete bar elements, placed between concrete bulk elements and equipped by bond stress-slip relations
98 to allow the propagation of any crack across the steel rebars avoiding undesired crack arrest effects [43,44].
99 An interesting numerical model to analyze the crack behavior which does not require the use of discrete
100 elements is proposed in [45] and takes the advantages of a crack queuing algorithm to capture the stress
101 redistribution during the crack propagation as well as the stress concentration at crack tips. Moreover, recent
102 works show as the phase field model is very efficient to simulate the damage phenomena in RC structures
103 [46,47] but it is not able to directly compute the crack width unless some post‐processing strategies are
104 incorporated [48]. Interesting modeling strategies employed general adaptive multiscale and homogenization
105 approaches to predict micro-cracking and contact evolution in both composite materials and heterogeneous
106 beams [49–53] and to reduce the computational effort of the analysis.
107 Despite the several existing models for predicting the global structural response of RC structures, numerical
108 models able to simulate the entire nonlinear cracking behavior, including crack onset and propagation as well
109 as coalescence and branching phenomena, correctly evaluating the design requisites of the structures at the
110 serviceability limit states, are still missed. To this end, in this work, an integrated numerical model, proposed
111 by some of the authors for the failure analysis of externally strengthened RC structures [18], is here improved
112 to assess its capabilities for predicting the cracking behavior, in terms of crack width and crack spacing, of
113 real-size RC structural elements under quasi-static loading conditions. The integrated model, illustrated in
114 Section 2 of this paper, adopts a cohesive zone approach and a bond-slip model, already introduced in [52,53],
115 to simulate respectively the damage phenomena in the concrete phase and the mechanical interaction between
116 steel rebars and surrounding concrete, and it is enriched by a novel user subroutine to directly compute, during
117 the simulation, the width of the propagating cracks. Cracking analyses have been performed by using the
118 proposed model, involving RC structures subjected to axial and shear/bending stress states. Original numerical
119 results, reported in Section 3, obtained by the adopted model, in terms of crack width and crack spacing, have
120 been compared with experimental results highlighting the good prediction capabilities of the model in the
121 analysis of the crack patterns (see Section 3.1). Moreover, the comparison with numerical results shows that
122 (see Section 3.2) the proposed method allows the cracking behavior in RC structures to be accurately simulated
123 while preserving the discrete nature of fracture phenomena and their interaction with the reinforcing bars, more
124 than existing smeared crack and damage models whose the fracture essential features are inevitably lost in the
125 smoothing process. Finally, a critical discussion and future perspectives of the present work are reported in
126 Section 4.
127
128 2. Numerical model for the reinforced concrete (RC) structural elements

129 In this section, the proposed numerical model to predict the cracking behavior of RC structures is briefly
130 explained. In particular, Section 2.1 shows the adopted theoretical fracture approach, able to reproduce the
131 damage in the concrete phase, and its numerical implementation within a finite element framework. Then, the
132 modeling strategy for simulating the concrete/rebar mechanical interaction is illustrated in Section 2.2.

133 2.1. Diffuse cohesive model for concrete cracking

134 The damage phenomena in the concrete phase, including crack nucleation and propagation as well as crack
135 coalescence and branching, are simulated through an inter-element cohesive fracture approach according to
136 which a set of cohesive elements, placed between all standard (bulk) finite elements of the computational mesh,
137 allow crack-induced nonlinear processes to be easily predicted.
138 The adopted theoretical fracture approach, explained in detail within [54], relies on a variational formulation
referring to a general two-dimensional fracture body, occupying the region   R bounded by the piecewise
2
139
140  , subjected to both body f and surface t forces that act on the domain  and its Neumann boundary  N
141 , respectively, whereas prescribed displacements u are imposed on the Dirichlet boundary  D . To allow the
cohesive elements to be correctly inserted, the original domain  is replaced with its planar tessellation 
h
142
143 (see Fig. 1), thus considering all mesh boundaries as cohesive interfaces, denoted as  hd , potentially subjected
144 to damage. It follows that the actual crack can be approximated by a subset of the damaged cohesive interfaces
145  ch   hd . The interfaces can be regarded as the union of the two crack faces (positive  hd + and negative  hd − )
+ −
146 where the related self-balanced cohesive tractions, i.e. tcoh and tcoh take action (see the right-hand side of Fig.
147 1).

148
149 Fig. 1. Schematic representation of the diffuse cohesive approach.

150 The nonlinear behavior of the cohesive tractions is described by a traction-separation law of the kind
151 tcoh = K u , in which K is the second‐order constitutive tensor, while u = u+ − u− is the displacement jump
152 occurring between the material bulk elements. The associated BVP is written in the following weak form: find
u U such that:
h h
153

154  C ( uh )   ( v h ) d +  h K uh  v h d =  f  v h d +  h t  v h d v h V h (1)
h \ hd d h \ hd N

155 where C and  are the fourth-order elasticity tensor and the usual linear strain operator, respectively. The
156 unknown displacement field u and the arbitrary virtual displacement field v belong respectively to the set of
157 kinematically admissible displacement U h and the set of kinematically admissible variations of the
158 approximated displacement field V h . The second‐order constitutive tensor K can be expressed as a function
159 of normal K n and tangential K s cohesive stiffness when direct cross-coupling between normal and tangential
160 deformation modes is not assumed:

161 K = Kn n  n + Ks ( I − n  n ) (2)

162 being I the second-order identity tensor, and n is the unit normal vector to  hd − . It is worth highlighting that
163 the initial stiffness components, denoted as K n0 and K s0 , play an important role of penalty parameters and must
164 be suitable set to enforce the inter-element continuity. Moreover, the undamaged interfaces, whose amount is
165 greater than the cracked elements, could seriously affect the overall stiffness of an ideally uncracked body due
166 to the artificial compliance effect introduced by the elastic branch. To reduce this compliance, avoiding at the
167 same time ill-conditioning issues of the tangent stiffness matrices in static analyses induced by the adoption of
168 very high stiffness values, in the present work, the following expressions are employed to compute the elastic
169 stiffness parameters ( K n0 and K s0 ) of the cohesive elements:

E'
170 K n0 =  , K s0 =  K s0 (3)
Lmesh

171 involving two dimensionless parameters  and  , obtained by the calibration technique proposed in [54], as
172 a function of the desired Young’s modulus reduction and the Poisson’s ratio of the material.
173 The numerical implementation of the above-explained fracture approach in a finite element framework requires
174 a preprocessing operation to build an enriched finite element mesh. To this end, the discretized computational
175 domain has been firstly split to obtain a duplication of every common node shared by each pair of bulk elements
176 and then enriched by zero-thickness interface elements inserted between the opposite faces of all neighboring
177 elements. The mechanical behavior of the embedded interface elements is described by the following cohesive
178 law written in the matrix form:

tncoh   K n0 0   n 
179  coh  = (1 − D )    (4)
ts  0 K s0   s 

180 where  n and  s are the normal and tangential components of the displacement jump u . The isotropic scalar
181 damage variable, denoted as D , ranges from 0 to 1 with an exponential-type evolution law as depicted in Fig.
182 2, and depends on an effective mixed-mode displacement jump defined as follows:

 m = (  n ) + ( s )2
2
183 (5)

184 where the symbol  is Macaulay operator which ensures the condition that the effective separation is
185 insensitive to a compressive displacement jump. The adopted traction-separation law of the cohesive forces
186 together with the damage evolution law is depicted in Fig. 2. For a deeper insight of the cohesive law, the
187 reader is addressed to [54]
188
189 Fig. 2. Adopted traction-separation relation and associated damage evolution law.

190 The mixed-mode displacements representing the crack initiation  m0 and complete separation  mf (see Fig. 2),
191 have been derived by adopting the following stress-based quadratic interaction criterion and the linear power
192 law criterion [54]:

2 2
 tn   t s  GI GII
193   +   = 1, + =1 (6)
 tnc   tsc  GIc GIIc

194 which involve both normal and tangential critical interface stresses ( t nc , t sc ) and both mode-I and mode-II
195 fracture energy ( GIc , GIIc ) of the tested material. The adopted traction-separation law, reported in Fig. 2, is not
196 able to describe in a satisfactory manner the compressive behavior of the concrete, and therefore, the crushing
197 failure mode, typically experienced in RC structures, is not predicted by the proposed model. However, the
198 interpenetration phenomenon is prevented by adding a unilateral contact constraint enforced via a penalty
199 method, which adopts the value Kn0 as regularization parameter.

200 The irreversibility condition essential to prevent any artificial self-healing of the material is enforced by means
201 of two additional weak contributions defined on the internal mesh boundaries  hd , written as follows:

 ( − max ( m ,  mmax,old ) )   ( mmax ) dΓ


max
m
d
h

202 (7)
 ( −  mmax )   ( mmax,old ) dΓ
max, old
m
 hd

203 Eqs. (7) allow recording the maximum value attained by the effective mixed-mode displacement  m (see Eq.
204 (5)) at the actual and previous loading steps of the analysis involving two auxiliary state variables, i.e.  mmax
205 and  mmax,old respectively. Such variables are set equal to zero at the beginning of the simulation. Moreover, the
206 symbol  ( ) represents the test operator for the enclosed quantity.

207 2.2. Embedded truss model for concrete/rebars interaction

208 Crack width and spacing predictions are strongly influenced by the adopted bond model to describe the
209 mechanical interaction between concrete and steel rebars. Indeed, due to such bond behavior, some parts of
210 the tensile forces are carried by the concrete tooth between two adjacent cracks providing a stress-strain relation
211 of the reinforcing bar different with respect to the bare bar one. It follows that, at the crack location, the steel
212 rebar yielding begins earlier than other points. Moreover, cracking phenomena occurring within the bonding
213 zone, induced by pull-out and splitting mechanisms depend on the adhesion and friction properties of the given
214 concrete/rebar bi-material system. Detailed modeling of such a system, including the steel reinforcement ribs
215 as well as the concrete lugs, inevitably leads to high computational efforts and modeling issues due to the need
216 of very refined computational meshes.

217

218 Fig. 3. Elastoplastic constitutive behavior and bond-slip relation adopted for the rebars (a). Schematic
219 representation of the embedded truss elements together with computational details of the crack width
220 computation (b).

221 In the present work, a simplified model, already proposed by some of the authors in [55], called embedded
222 truss model, is employed to simulate the mechanical behavior of the rebars and their interaction with the
223 concrete phase. Two-node truss elements equipped with an elastoplastic stress-strain relation, having linear
224 hardening whose tangent plastic modulus is set equal to one hundred times smaller than the elastic one (see
225 the upper portion of Fig. 3a), are employed to model the reinforcing bars. In this way, the rebars are not
226 modeled as embedded plates having the same width as the concrete section but rather as embedded 1D (truss)
227 elements having a cross-section that is different from the concrete one. Such truss elements possess an
228 equivalent section computed as the sum of the cross-section of the individual steel rebar used in the RC
229 structure, thus allowing all rebars and stirrups at a given depth to be concentrated in a 1D element. Such
230 elements are connected to the concrete mesh by means of zero-thickness concrete/rebars interface elements
231 equipped with the well-known bond-slip relation, taken from [2], valid for ribbed bars and good bond
232 conditions (see the lower portion of Fig. 3a). By embedding truss elements within the planar volumetric
233 concrete mesh, a node overlapping of the two meshes (steel and concrete) occurs (see the left portion of Fig.
234 3b). However, the adopted nonlinear interface elements, acting as nonlinear tangential springs between rebar
235 truss elements and concrete bulk elements, define the bond strength and stiffness as functions of the relative
236 tangential displacement (slip), s, between the nodes of the two meshes. Moreover, a perfect constraint in the
237 perpendicular direction to the rebars is enforced in the overlapped nodes thus allowing only tangential
238 displacements. According to the above-explained embedded truss model, the crack propagation across the
239 reinforcement layer induces a tangential displacement between truss and volumetric elements. It follows that,
240 for any truss node, two different slips can be defined for the steel/concrete element pairs. The crack width
241 computation at the level of the rebar is obtained by the sum of these two different slips, i.e. s + and s − reported
242 in the right portion of Fig. 3b.
243 The implementation of the embedded truss model results to be a fundamental modeling ingredient of the
244 proposed model for simulating the cracking behavior since avoids undesired crack arrest effects caused by the
245 presence of discrete rebar elements and allows crack width and crack spacing to be correctly calculated.

246 3. Cracking analysis of RC structural elements


247 In this section, the proposed integrated model has been employed to perform numerical simulations for
248 predicting the load-carrying capacity and the related crack patterns of real-scale RC structural elements
249 subjected to general loading conditions. In particular, in Section 3.1 the proposed model has been validated by
250 predicting both loading and cracking response of RC members under direct tensile test, providing also an
251 analysis of the mesh size influence on the numerically predicted fracture properties. In Section 3.2, cracking
252 analyses have been performed involving RC beams under flexural load with different steel reinforcement
253 arrangements. In both sections, in order to assess the prediction capabilities of the proposed model, suitable
254 comparisons with numerical and experimental results are reported.

255
256 3.1. RC structural element under tension loading conditions
257 The direct tension test experimentally performed by [9] has been simulated here by using the proposed
258 numerical fracture model. The experimental program [9] was planned to assess the influence of the concrete
259 cover thickness on the cracking behavior of the tested specimens. To this end, six axially loaded RC members
260 with a rectangular cross-section containing a 19 mm steel reinforcement bar (see Fig. 4), were experimentally
261 tested. The cross-sections of these RC elements, whose dimensions are reported in Table 1, have different
262 aspect ratios in order to vary the concrete cover thickness ratio c / Ø between 1 and 3.5 in 0.5 increments. In
263 these tensile specimens, the concrete cover thickness is the smaller length measured from the cross-section
264 boundary to the external surface of the rebars.

265
266 Fig. 4. Geometric configuration of the axially loaded RC element.

267 Table 1. Geometric parameters of the tested RC elements.


Dimensions [mm]
Specimens Cover thickness ratio c/Ø
a b L
N10 1 60 385
N15 1.5 80 290
N20 2 100 230
1500
N25 2.5 115 200
N30 3 135 170
N35 3.5 150 155
268
269 The elastic parameters of constituent materials are the following: Young’s modulus and Poisson’s ratio are
270 respectively equal to Ec = 31.48 GPa and  c = 0.2 for concrete and Es = 200 GPa and  s = 0.3 for steel of
271 the reinforcement bar. The tensile and compressive strengths of the concrete phase, taken from [9], are equal
272 to f ck = 24.8 MPa and f ct = 1.96 MPa , respectively, while the yielding stress of the rebars is equal to
273 f y = 430 MPa . The cohesive parameters required by the traction-separation law, useful to describe the
274 behavior of the cohesive tractions acting between the bulk elements, are reported in Table 2. In this pure mode-
275 I fracture test, the mode-II fracture parameters, in terms of stress ( f s ) and fracture energy ( GIIc ) are set equal
276 to the mode-I counterparts (i.e. f t and GIc ), thus allowing the mesh-induced toughening effect associated with
277 the activation of local mixed-mode conditions during the crack propagation, to be notably reduced. A plane
278 stress state is assumed. Although the thickness of the tested structural element is not very small compared to
279 the other dimensions, such a condition, currently assumed for beam problems accordingly to the Saint Venant
280 Theory, is preferred to that plane strain one because avoids a potential overestimation in both strength and
281 stiffness associated with the assumption of zero out-of-plane strain. The dimensionless parameter  , required
282 by the adopted traction-separation law to govern the damage evolution rate in quasi-brittle materials, is set
283 equal to 5 because of provides a significantly faster strength degradation after t coh [54].

284 Table 2. Cohesive parameters of the adopted traction-separation law for the tensile test.
f t [MPa] f s [MPa] GIc [N/m] GIIc [N/m] K n0 [N/m3] K s0 [N/m3] 
1.96 1.96 85 85 1.70E14 1.70E14 5

285 From a computational point of view, the concrete domain of the RC members is discretized by performing a
286 Delaunay tessellation, consisting of planar three-node bulk elements with a prescribed maximum edge length
287 of 30 mm, enriched by four-node zero-thickness interface elements placed along the boundaries of the adopted
288 mesh (see Fig. 5), while, the steel rebars are modeled by two-node truss elements. As shown in Fig. 5, cohesive
289 elements are not inserted along the steel rebar for avoiding a constraint overlapping induced by the adopted
290 traction-separation and bond-slip relations for concrete and rebar behavior, respectively. The following
291 simulations have been performed under displacement-controlled quasi-static loading conditions, adopting a
292 displacement increment of 5e−3 mm for the point P illustrated in Fig. 4.

293
294 Fig. 5. Computational discretization of the numerically tested RC elements.
295
296 Fig. 6. Numerically and experimentally predicted structural responses of the RC specimens.

297 Fig. 6 shows the numerically predicted structural responses of the six RC elements together with a comparison
298 with available experimental and numerical results taken from [9] and [43], respectively. The load versus
299 average steel strain curves of all tested specimens predicted by the proposed model result to be globally in
300 good agreement with the experimental results, showing in any curve the well-known four loading stages of an
301 RC element under tensile load, reported in Fig. 6 only for the specimen N10, i.e. the elastic stage (1), the crack
302 formation stage (2) and the stabilized cracking stage (3) before the branch associated with the rebar yielding
303 (4). Compared to the results obtained by the fiber beam-column element model adopted in [43], based on the
304 theoretical -ellipse approach, the proposed cohesive model predicts loading curves with sudden load
305 reductions associated with the crack nucleation. This behavior is clearer in the specimens with a great cover
306 thickness (N25, N30, and N35), in which the related crack patterns are characterized by transversal cracks and
307 well-defined concrete teeth (see Fig. 7) thus providing a greater tension stiffening effect than the other RC
308 specimens. Fig. 7 shows the numerically predicted crack patterns of the RC elements together with both
309 concrete normal and rebar axial stresses at yielding load. As highlighted by the experimental evidence [9] the
310 cracking predictions significantly change with the cover thickness. In particular, the crack patterns of the
311 specimens having a thin cover thickness (N10, N15, and N20) are mainly characterized by splitting cracks, i.e.
312 cracks parallel to the rebar, while in the remaining RC specimens (N25, N30, and N35), well defined transverse
313 cracks appear (see Fig. 7).
314
315 Fig. 7. Deformed configurations, crack patterns, and stress maps of the tested RC specimens.

316 Moreover, the proposed fracture model accurately captures the effect of the interaction between the reinforcing
317 bars and the surrounding cracking concrete. In particular, the stress maps of Fig. 7 show high concrete stress
318 values, close to the critical concrete tensile strength, within the concrete teeth, representing the capacity of the
319 undamaged concrete to carry tensile forces between adjacent primary cracks, while high rebar axial stress
320 values, very close to the yielding stress, in correspondence of the cracks.

321 The cracking analysis has provided the average distance between adjacent main transverse cracks, i.e. the crack
322 spacing. It is worth specifying that, here and throughout this article, the primary cracks are defined as the
323 cohesive elements having a damage variable value equal to D = 1 at yielding load. Moreover, the crack spacing
324 has been measured at the rebar level. The numerical results predicted by the model, reported in Fig. 8, have
325 the same trend as the experimental outcomes highlighting a larger crack spacing as the cover thickness
326 increases except for specimens N10 and N15 where the crack spacing is strongly influenced by the large
327 number of splitting cracks at the early loading stage (see Fig. 7). A lack of convergence of the crack spacing
328 occurs for specimens N30 and N35, since, as reported in Table 3, a different crack number with respect to the
329 available experimental range is predicted by the model. However, this result is influenced by the random
330 location of the cracks within the unstructured computational discretization predicted by the model.
331
332 Fig. 8. Numerically and experimentally predicted average crack spacing of the tested RC elements.

333 Table 3. Results of the numerical and experimental investigations


Crack number Average crack spacing [mm]
Specimens
Proposed Model Experiments Proposed Model Experiments
N10 8 7 10 162 188 136
N15 10 8 12 136 167 115
N20 10 10 9 130 136 150
N25 8 7 9 153 188 150
N30 7 6 5 185 214 250
N35 7 4 6 185 300 214
334
335 The width of the primary transverse cracks of specimen N35 predicted by the proposed model has been
336 reported in Fig. 9. In particular, in Fig. 9a the width of the seven transverse cracks are reported, numbered in
337 chronological order in the graph, and calculated during the entire simulation together with the numerical result
338 obtained by [43]. After the elastic stage, in which the concrete is intact without cracks, the curves predicted by
339 the proposed model exhibit the well-known crack formation and stabilized cracking stages, representing
340 respectively the nucleation of the seven cracks at different load levels and their stable propagation. By
341 analyzing the crack formation stage, we can note that the width of the early cracks is influenced by the
342 nucleation of the new cracks, undergoing temporary crack width reduction. Moreover, the comparison with
343 the numerical result [43], obtained by a numerical model which uses an approximated solution for the crack
344 width computation, shows a good agreement at both elastic and crack formation stages, but a lack of
345 convergence appears within the stabilized cracking stage probably due to the fact that the proposed model
346 takes into account the nonlinear processes associated with unloading stages, reducing the crack width values.
347 In Fig. 9b, the deformed configurations together with the concrete stress map at load levels corresponding to
348 the nucleation of cracks, are reported. It is worth highlighting that the new cracks occur within the concrete
349 teeth characterized by a high stress concentration and concrete normal stress that approaches the critical tensile
350 strength of the material. After the crack nucleation, a stress-free zone adjacent to the crack appears.
351
352 Fig. 9. Cracks width analysis of specimen N35: (a) comparison with a numerical result [43] and (b) deformed
353 configurations with concrete stress map.

354 These results show the good capability of the proposed integrated model to simulate the entire cracking
355 process, including crack onset and propagation, of reinforced concrete elements, providing numerical results,
356 in terms of load-carrying capacity and crack patterns, in reasonable agreement with the experimental outcomes.
357 However, the numerical predictions could be affected by the adopted computational discretization. To this end,
358 a mesh sensitivity analysis has been performed involving the N35 RC specimen. Three different unstructured
359 meshes, called Mesh 1 (the coarsest mesh), 2, and 3 (the finest mesh), have been built prescribing a maximum
360 element size of 4.5 mm, 3 mm, and 1.5 mm, respectively. The resulting meshes, reported in the graph of Fig.
361 10, consist of 362, 1088, and 4244 planar three-node bulk elements and 581, 1688, 6476 zero-thickness four-
362 node cohesive elements, respectively for Mesh1, 2, and 3. It is worth noting that Mesh 2 has been already
363 employed in the previous analyses.

364 Fig. 10a clearly shows that the global structural response, in terms of load versus average steel strain curve, is
365 almost independent of the adopted mesh size. In particular, the analyses, conducted by using the three different
366 meshes, provide the same crack load, of about 57.4 kN, representing the end of the elastic branch, and reach
367 the yielding load at the same average steel strain. However, the coarsest mesh (Mesh 1) predicts a wider crack
368 formation stage, ranging from 58 to 118 kN, than other meshes. This result is confirmed by the numerically
369 predicted crack patterns reported in Fig. 10b which highlight a different prediction of the cracking behavior,
370 in terms of crack number and related crack spacing obtained by Mesh 1 with respect to Mesh 2 and 3. Such a
371 lack of convergence of the Mesh 1 crack pattern is probably due to the fact that the adopted discretization is
372 not sufficiently refined to evaluate accurately the stress gradients within the fracture process zone. Moreover,
373 it is worth noting that the crack pattern of Mesh 1 is affected by the well-known nonuniqueness issues, widely
374 explained in [56], which influence the stability of the associated numerical computations leading to strong
375 mesh dependency in terms of no-localized fracture patterns (see Fig. 10b).
376
377 Fig. 10. Numerical results of the mesh sensitivity analysis: (a) loading curves highlighting the crack
378 formation stage and (b) deformed configurations together with stress maps at the yielding load.

379 It follows that despite the good predictions in terms of loading curves, the use of coarsest mesh provides
380 inaccurate crack patterns and should be avoided in the crack width and crack spacing analysis. Therefore, the
381 numerical results previously obtained by using the same mesh size of Mesh 2 can be judged acceptable, thus
382 confirming the reliability of the proposed model to predict the overall structural response of RC components.

383 3.2. RC structural element under flexural loading conditions

384 The adopted numerical model has been here employed to investigate the cracking behavior of reinforced
385 concrete structural elements. Specifically, a four-point bending test involving three RC beams, experimentally
386 analyzed in [6] to assess the influence of the tensile rebars arrangement on the flexural stiffness and cracking
387 resistance, has been simulated. The geometric configuration as well as the loading and boundary conditions of
388 the three tested beams, called S1-2, S1-4, and S2-3, are reported in Fig. 11. The elastic parameters of concrete
389 used in the numerical analyses, i.e. Young’s modulus and Poisson’s ratio, are Ec = 24 GPa and  c = 0.18 ,
390 respectively, while the strength parameters, in terms of compressive and tensile strength, are f c = 49.4 MPa
391 and ft = 2 MPa , respectively. The adopted tensile strength, not reported among the experimental data, is
392 chosen to match the experimental results. The mechanical properties, in terms of Young’s modulus and
393 yielding stress, of the different steel rebars, are directly taken by [6] and respectively equal to Es = 210.5 GPa
394 and f y = 632.3 MPa for Ø14 rebar and Es = 199.3 GPa , f y = 551.1 MPa for Ø22 rebar.
395
396 Fig. 11. Geometric configuration and boundary conditions of the three tested beams (all dimensions are
397 expressed in mm).

398 The computational discretization of the concrete phase consists of three-node triangular elements arranged
399 according to an unstructured isotropic (i.e. Delaunay) spatial distribution and reported in Fig. 12. A mesh
400 refinement has been performed within the critical region dominated by the shear/flexural stress state by
401 prescribing a maximum element size of 12.5 mm. In this region, zero-thickness four-node cohesive elements,
402 highlighted by blue lines in Fig. 12, are inserted between the concrete bulk elements. On other hand, the steel
403 rebars are discretized by two-node truss elements whose size is equal to the corresponding edge of the adjacent
404 concrete bulk element. It is worth noting that the effective depth of the truss elements is considered with respect
405 to the centroid of the tensile reinforcement (see Fig. 11). It follows that the meshes related to the three beams
406 differ in the numerical concrete cover thickness. Moreover, rigid steel plates length 60 mm and height 15 mm
407 are modeled under both support and load application points, as depicted in Fig. 12.

408
409 Fig. 12. Computational discretization of the tested beams together with the adopted cohesive elements.

410 The strength parameters required by the adopted mixed-mode constitutive law, in terms of normal/tangential
411 critical interface stresses ( f t and f s ) and mode-I/II fracture energies ( GIc and GIIc ), are reported in Table 4
412 together with the dimensionless parameter  and the initial normal/shear stiffness parameters ( K n0 and K s0 ).
413 For this example, the mode-II parameters, which cannot be directly identified by the experimental test, have
414 been set equal to f s = 2 ft and GIIc = 10GIc , in agreement with the interesting results obtained in [55]. The
415 following numerical simulations are conducted under plane stress state and quasi-static loading conditions,
416 adopting a displacement-control solution scheme with constant increments of the mid-span deflection equal to
417 2e-3 mm.

418 Table 4. Cohesive parameters of the adopted traction-separation law for the four-point bending test.
f t [MPa] f s [MPa] GIc [N/m] GIIc [N/m] K n0 [N/m3] K s0 [N/m3] 
2 2.83 100 1000 3.13E14 1.57E14 5

419 The numerical results, in terms of both load versus mid-span deflection curve and crack pattern, obtained by
420 the proposed model are reported in Fig. 13. The numerically predicted loading curves are in good agreement
421 with the experiments [6] and the numerical results obtained by a numerical model based on a smeared crack
422 approach [30], showing the typical trilinear behavior in which three different branches are clearly detected
423 coinciding with the elastic stage, the flexural/shear crack propagation stage, and the stage associated with the
424 rebar yielding.

425
426 Fig. 13. Global structural response of the tested beams: (a) load versus mid-span deflection curves and (b)
427 deformed configurations together with crack patterns and stress maps.
428 The crack patterns of the tested beams at the service load, assumed as 50% of the ultimate theoretical moment
429 [30] and highlighted in Fig. 13a, are represented in a very realistic manner by the proposed model (see Fig.
430 13b), showing different cracking behavior induced by the different tensile rebar arrangements. In particular,
431 in Beam S1-2 the crack propagation is hampered by the presence of the great rebar quantity (five bars with a
432 diameter of 14 mm), leading to a more diffuse cracking within the concrete cover thickness with respect to
433 Beam S1-4 and S2-3. In fact, Beam S1-2 possesses a great steel surface which induces high tangential strength
434 between rebars and surrounding concrete thus leading to small rebar/concrete relative slips according to the
435 adopted bond-slip relation. Such behavior highlights the effectiveness of the embedded truss model to capture
436 the rebar/concrete mechanical interaction.
437 Moreover, considering only the cracks within the constant bending moment region at the service load both
438 crack width and crack spacing are compared with available numerical [30] and experimental [6] results. In
439 particular, Table 5 reports the values of the maximum ( wm ) and average ( wa ) crack width predicted by the
440 numerical models together with the percentage relative deviation ( ew ) with respect to the experimental data,
441 computed as:
wnum − wexp
442 e=  100 . (8)
wexp

443 Table 5. Crack width comparison with the present and available results.
Crack Width
Maximum Average
Gribniak et al. Gribniak et al.
Rimkus et al. Rimkus et al.
Beam (Load step) Proposed Model Experiment Proposed Model Experiment
Model Model
(reference) (reference)

wm ewm wm ewm wm wa ewa wa ewa wa


[mm] [%] [mm] [%] [mm] [mm] [%] [mm] [%] [mm]
S1-2 (124 [kN]) 0.102 0.05 0.154 50.98 0.102 0.068 4.37 0.136 91.54 0.071
S1-4 (99 [kN]) 0.110 8.33 0.160 33.33 0.120 0.060 4.76 0.122 93.65 0.063
S2-3 (74 [kN]) 0.123 2.13 0.152 26.66 0.120 0.072 30.04 0.120 118.18 0.055

444 The proposed model predicts the maximum width of the three tested beams very close to the experimental ones
445 obtaining percentage relative deviations ( ewm ) below 8.33% and smaller than those predicted by [30]. Similar
446 results have been obtained for the numerically predicted average crack width except for the Beam S2-3 in
447 which a deviation ( ewa ) of about 30.04% is observed.

448 Table 6. Crack spacing comparison with the present and available results
Crack Spacing
Maximum Average
Gribniak et al. Gribniak et al.
Rimkus et al. Rimkus et al.
Beam (Load step) Proposed Model Experiment Proposed Model Experiment
Model Model
(reference) (reference)

sm esm sm esm sm sa esa sa esa sa


[mm] [%] [mm] [%] [mm] [mm] [%] [mm] [%] [mm]
S1-2 (124 [kN]) 137.50 2.96 135.00 4.73 141.70 90.00 6.44 100.50 4.47 96.20
S1-4 (99 [kN]) 125.00 16.67 103.30 31.13 150.00 86.36 17.78 150.00 42.82 105.03
S2-3 (74 [kN]) 125.00 6.79 105.00 21.70 134.10 97.73 12.33 81.80 5.98 87.00

449 Table 6 shows the comparison, in terms of both maximum ( sm ) and average ( sa ) crack spacing, between
450 numerical and experimental results. The results obtained by the proposed model are in good agreement with
451 the experiments, providing a percentage relative deviation of the maximum crack spacing ( esm ), computed
452 similarly to (8) but inserting the corresponding crack spacing values, smaller than that obtained by [30] and
453 below 16.67%. On the contrary, a better prediction of the average crack spacing is obtained by the smeared
454 crack model [30] except for the Beam S1-4. However, the deviations ( esa ) predicted by the proposed model
455 with respect to the experiments, present a maximum value of about 17.78%.

456
457 Fig. 14. Bending moment versus curvature curves of the tested RC beams.

458 Table 7. Width of the five largest cracks ( w1−5 ) and the average crack width ( wa ) measured at the load steps
459 reported in Fig. 14. All measured values are in µm.
Beam S1-2 Beam S1-4 Beam S2-3
Gribniak et al. Gribniak et al. Gribniak et al.
Proposed Model Proposed Model Proposed Model
Experiment Experiment Experiment
w1 27 50 65 60 75 40
w2 26 40 52 60 72 40
w3 25 30 47 38 54 30
Step 1 w4 25 24 43 34 53 30
w5 24 20 36 24 52 30
wa (cracks) 23 (9) 31 (10) 38 (11) 33 (11) 43 (10) 24 (9)
w1/wa 1.15 1.61 1.74 1.82 1.74 1.67
w1 38 62 84 80 94 60
w2 37 50 68 60 91 60
w3 37 50 62 54 68 50
Step 2 w4 36 40 57 50 64 40
w5 36 40 53 34 62 40
wa (cracks) 34 (10) 39 (11) 50 (11) 45 (11) 54 (12) 29 (11)
w1/wa 1.1 1.59 1.67 1.77 1.73 2.07
w1 85 102 123 120 128 100
w2 84 80 94 90 125 80
w3 83 70 92 70 98 60
w4 83 62 91 70 84 60
Step 3
w5 80 54 85 44 83 60
w max (MC) 179 208 191
wa (cracks) 68 (11) 59 (11) 73 (12) 61 (11) 75 (12) 42 (12)
w1/wa 1.26 1.73 1.69 1.96 1.71 2.41
w1 110 142 136 140 143 120
w2 104 100 105 100 142 80
w3 102 100 103 100 107 80
w4 101 100 94 70 95 80
Step 4
w5 100 90 91 60 93 80
w max (MC) 223 264 248
wa (cracks) 84 (11) 88 (11) 81 (12) 71 (12) 81 (13) 52 (13)
w1/wa 1.32 1.61 1.68 1.97 1.76 2.31

460 Fig. 14 shows the structural response of the tested beams, in terms of bending moment versus curvature curve,
461 predicted by the proposed model together with the experimentally obtained results. The numerical results are
462 in good agreement with the experimental ones, providing an ultimate bending moment for each beam,
463 coinciding with the initial point of the yielding stage in the loading curve (see Fig. 13a), very close to that
464 obtained by the experiment. Moreover, in Fig. 14 are reported the load steps adopted for analyzing in-depth
465 the predicted crack patterns. In particular, at these load steps the width of the five largest cracks ( w1−5 ) within
466 the pure bending zone and the average crack width ( wa ) have been measured. It is worth noting that the crack
467 width is the maximum crack width value computed along the associated crack path. The results of the analysis
468 together with a comparison with the experiments [6] are reported in Table 7. We can see that the numerically
469 predicted crack widths are very close to the experimental ones, as well as the average crack width and related
470 crack number, showing some result divergence in the Beam S2-3 predictions. However, as highlighted by the
471 authors of the experiments [6], the stochastic nature and complex topology of the cracks make complicate the
472 crack width experimental measurements thus providing a qualitative rather than a quantitative analysis of the
473 crack width. An important parameter to assess the different cracking behavior on varying the tensile rebar
474 arrangement is the ratio between the maximum and average crack width ( w1 / wa ) reported in Table 7. As
475 highlighted by both experimental and numerical results, such a ratio is strongly dependent on the total rebar
476 perimeter and increases with the decreasing of the employed rebar perimeter. Similar values of this ratio are
477 numerically predicted by the proposed model for Beam S1-4 and 2-3 since possessing similar total bar
478 perimeter. Moreover, the crack widths measured at the service load steps (Step 3 and 4) are compared with
479 those computed by the Model Code 2010 [2]. The mathematical model of [2] provides overestimated crack
480 widths of about two times with respect to the experiments and the proposed numerical model. However, the
481 predictions of the Model Code 2010 must be considered as the maximum crack widths that can potentially
482 occur in the reinforced concrete structures.

483
484 Fig. 15. Average (a) and maximum (b) crack spacing obtained by the proposed model and the experiments.

485 Finally, the cracking analysis has been extended to the average and maximum crack spacing computation. The
486 crack spacing predictions obtained by the proposed model for each tested beam are globally in good agreement
487 with the experimental results (see Fig. 15). In particular, in the final stages of the simulations, where the cracks
488 are almost completely developed, both average and maximum crack spacing is very close to the experimental
489 outcomes. However, Fig. 15 shows a lack of convergence for the crack spacing results in the initial stage of
490 the analysis, especially for the Beam S2-3 predictions. This issue is mainly due to the few cracks present within
491 the crack patterns in the early loading steps whose location is practically random in the computational domain
492 characterized by the constant flexural stress state.

493 4. Conclusions

494 In this work, the cracking behavior of reinforced concrete structural elements under general quasi-static loading
495 conditions, in terms of crack width and crack spacing, has been analyzed by means of a numerical fracture
496 model developed by the authors in [54,55]. The adopted model relies on a diffuse interface model and an
497 embedded truss model to capture both damage phenomena in the concrete phase and bond behavior between
498 concrete and rebars, respectively. In particular, according with an inter-element facture approach, zero-
499 thickness cohesive elements placed along all mesh boundaries allow crack onset and propagation to be easily
500 predicted, while, according with a discrete bar approach, truss elements equipped with a bond-slip relation are
501 embedded in the concrete mesh to model reinforcing bars and to capture their interaction with the surrounding
502 concrete. Numerical analyses of real-size RC members subjected to tensile and flexural loads have been
503 performed. The numerically obtained results, in terms of load-carrying capacity, crack width, and crack
504 spacing, are compared with available numerical and experimental outcomes. Moreover, the mesh dependency
505 issues have been analyzed to assess the reliability of the adopted numerical model for predicting concrete
506 cracking behavior.

507 In light of the obtained results, we can state that:

508 • the adopted model provides global structural responses, in terms of load-displacement and bending
509 moment-curvature curves, in good agreement with the experiments, predicting load levels associated with
510 the different strain stages, including both crack nucleation and stabilized cracking stages, as well as the
511 rebar yielding stage, very close to the experimental data. Moreover, the model is able to predict the well-
512 known tension stiffening effect as clearly shown by the structural responses of the axially loaded RC
513 elements;

514 • the numerically predicted crack patterns are represented in a realistic manner preserving the discrete
515 nature of cracking phenomena and resulting to be very similar to the experimental ones. Despite the
516 adoption of a 2D setting which inevitably leads to an approximation of the actual cracking behavior
517 neglecting the mode-III fracture mechanisms, the global effects of the concrete/rebars interaction have
518 been suitably captured, allowing both any crack to propagate across the reinforcing steel layer and
519 nucleation of concrete splitting cracks when small concrete covers are used. Moreover, crack coalescence
520 and branching phenomena, which typically occur in the reinforced concrete elements, are easily handled
521 without introducing complicated remeshing procedures thanks to the adoption of the diffuse interface
522 fracture model;
523 • the mesh sensitivity analysis shows numerical results almost independent of the mesh size. However, a
524 lack of convergence of the crack spacing is obtained for very coarse meshes providing no-localized
525 fracture patterns due to inaccurate stress gradients predictions within the fracture process zone induced
526 by the unrefined discretization adopted for the analysis;
527 • the crack width and crack spacing of RC structural elements predicted by the adopted model at the
528 serviceability limit state are in good agreement with that computed by the experimental test. The
529 mechanical effects of the rebar arrangement on the cracking behavior have been well captured by the
530 numerical model, highlighting that the assumed steel surface is the most cracking influencing factor, in
531 other words considering the same steel quantity, more rebars with small diameter are recommended to
532 obtain crack width reductions. The comparison with a numerical model based on a smeared crack
533 approach, in terms of maximum crack width, has highlighted the good prediction capabilities of the
534 proposed model obtaining percentage errors with respect to the experimental results smallest than those
535 obtained by the smeared crack model and less than 8.33 %. Similar results are obtained regarding the
536 crack spacing predictions computing a maximum deviation with respect to the experiments of about
537 17.78%. Such a deviation is more marked in the initial loading steps, considering both maximum and
538 average crack spacing, mainly because of the random distribution of the early cracks within the
539 computational domain. However, once the cracks are almost completely developed, spacing values
540 predicted by the adopted model are in good agreement with the experimental ones.
541 The obtained results have highlighted both effectiveness and reliability of the adopted model to predict the
542 cracking behavior over some existing fracture models since it is able to preserve the discrete nature of the
543 cracking phenomena, allowing the automatic computation of crack width and spacing without requiring a great
544 implementation effort.
545 As a future perspective of this work, the cracking analysis could be extended to the RC frame buildings
546 subjected to dynamic or cyclic loading conditions in which the damage phenomena occur in several parts of
547 the structures. However, to reduce the computational cost of the analyses the proposed fracture model should
548 be integrated within an adaptive concurrent multiscale approach similar to that developed by [50,57] for
549 composite materials.
550

551 Author Contributions:


552 Umberto De Maio: investigation, methodology, software, validation, formal analysis, writing-original draft
553 preparation, resources, visualization, and data curation. Fabrizio Greco: conceptualization, methodology,
554 investigation, supervision, writing-review and editing, funding acquisition and project administration.
555 Lorenzo Leonetti: conceptualization, investigation, resources, writing-review and editing, formal analysis.
556 Paolo Nevone Blasi: methodology, software, formal analysis, investigation, writing-original draft preparation.
557 Andrea Pranno: conceptualization, methodology, formal analysis, software, investigation, writing-review
558 and editing.

559

560 Declaration of Competing Interest


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

563

564 Acknowledgments
565 Fabrizio Greco and Paolo Nevone Blasi gratefully acknowledge financial support from the Italian Ministry of
566 Education, University and Research (MIUR) under the P.R.I.N. 2017 National Grant “Multiscale Innovative
567 Materials and Structures” (Project Code 2017J4EAYB; University of Calabria Research Unit).
568 Andrea Pranno gratefully acknowledges financial support from the Italian Ministry of Education, University
569 and Research (MIUR) under the PON 2014-2020 Azione IV.4 - Rep. N. 1062 of 10/08/2021
570

571 References

572 [1] D. Breysse, Deterioration processes in reinforced concrete: an overview, in: Non-Destructive
573 Evaluation of Reinforced Concrete Structures, Elsevier, 2010: pp. 28–56.
574 https://doi.org/10.1533/9781845699536.1.28.

575 [2] C.E.-I du B. (CEB-FIP), CEB-FIP Model Code for concrete structures 2010, (2013).

576 [3] E. C. for S (CEN), EN 1992-1-1 Eurocode 2: Design of concrete structures - part 1–1: General rules
577 and rules for buildings, (2004).

578 [4] ACI, Building code requirements for structural concrete:(ACI 318-95); and commentary (ACI
579 318r95), (1995).

580 [5] G. Kaklauskas, Flexural layered deformational model of reinforced concrete members, Magazine of
581 Concrete Research. 56 (2004) 575–584. https://doi.org/10.1680/macr.2004.56.10.575.

582 [6] V. Gribniak, A. Pérez Caldentey, G. Kaklauskas, A. Rimkus, A. Sokolov, Effect of arrangement of
583 tensile reinforcement on flexural stiffness and cracking, Engineering Structures. 124 (2016) 418–428.
584 https://doi.org/10.1016/j.engstruct.2016.06.026.

585 [7] X. Chen, J.J. Yan, H.Q. Yang, Influence of Aggregates on Cracking Sensitivity of Concrete, AMM.
586 204–208 (2012) 3299–3302. https://doi.org/10.4028/www.scientific.net/AMM.204-208.3299.

587 [8] M. Albitar, M.S. Mohamed Ali, P. Visintin, Evaluation of tension-stiffening, crack spacing and crack
588 width of geopolymer concretes, Construction and Building Materials. 160 (2018) 408–414.
589 https://doi.org/10.1016/j.conbuildmat.2017.11.085.

590 [9] G.-Y. Lee, W. Kim, Cracking and Tension Stiffening Behavior of High-Strength Concrete Tension
591 Members Subjected to Axial Load, Advances in Structural Engineering. 12 (2009) 127–137.
592 https://doi.org/10.1260/136943309788251614.

593 [10] J.R. Deluce, F.J. Vecchio, Cracking Behavior of Steel Fiber-Reinforced Concrete Members
594 Containing Conventional Reinforcement, ACI Structural Journal. 110 (2013) 481–490.

595 [11] S. Acierno, R. Barretta, R. Luciano, F. Marotti de Sciarra, P. Russo, Experimental evaluations and
596 modeling of the tensile behavior of polypropylene/single-walled carbon nanotubes fibers, Composite
597 Structures. 174 (2017) 12–18. https://doi.org/10.1016/j.compstruct.2017.04.049.

598 [12] C. Barris, L. Torres, I. Vilanova, C. Miàs, M. Llorens, Experimental study on crack width and crack
599 spacing for Glass-FRP reinforced concrete beams, Engineering Structures. 131 (2017) 231–242.
600 https://doi.org/10.1016/j.engstruct.2016.11.007.

601 [13] K.A. Harries, B.M. Shahrooz, A. Soltani, Flexural Crack Widths in Concrete Girders with High-
602 Strength Reinforcement, J. Bridge Eng. 17 (2012) 804–812. https://doi.org/10.1061/(ASCE)BE.1943-
603 5592.0000306.

604 [14] A. Soltani, K.A. Harries, B.M. Shahrooz, Crack Opening Behavior of Concrete Reinforced with High
605 Strength Reinforcing Steel, Int J Concr Struct Mater. 7 (2013) 253–264. https://doi.org/10.1007/s40069-013-
606 0054-z.
607 [15] M. Qiu, X. Shao, Y. Zhu, J. Zhan, B. Yan, Y. Wang, Experimental investigation on flexural cracking
608 behavior of ultrahigh performance concrete beams, Structural Concrete. 21 (2020) 2134–2153.
609 https://doi.org/10.1002/suco.201900339.

610 [16] U. De Maio, F. Fabbrocino, F. Greco, L. Leonetti, P. Lonetti, A study of concrete cover separation
611 failure in FRP-plated RC beams via an inter-element fracture approach, Composite Structures. 212 (2019)
612 625–636. https://doi.org/10.1016/j.compstruct.2019.01.025.

613 [17] V. Alecci, M. De Stefano, R. Luciano, L. Rovero, G. Stipo, Experimental Investigation on Bond
614 Behavior of Cement-Matrix–Based Composites for Strengthening of Masonry Structures, J. Compos. Constr.
615 20 (2016) 04015041. https://doi.org/10.1061/(ASCE)CC.1943-5614.0000598.

616 [18] U. De Maio, F. Greco, L. Leonetti, P. Nevone Blasi, A. Pranno, An investigation about debonding
617 mechanisms in FRP-strengthened RC structural elements by using a cohesive/volumetric modeling technique,
618 Theoretical and Applied Fracture Mechanics. 117 (2022) 103199.
619 https://doi.org/10.1016/j.tafmec.2021.103199.

620 [19] F. Greco, D. Gaetano, L. Leonetti, P. Lonetti, A. Pascuzzo, A. Skrame, Structural and seismic
621 vulnerability assessment of the Santa Maria Assunta Cathedral in Catanzaro (Italy): classical and advanced
622 approaches for the analysis of local and global failure mechanisms, Frattura Ed Integrità Strutturale. 60 (2022)
623 464–487.

624 [20] F.O. Falope, L. Lanzoni, A.M. Tarantino, Modified hinged beam test on steel fabric reinforced
625 cementitious matrix (SFRCM), Composites Part B: Engineering. 146 (2018) 232–243.
626 https://doi.org/10.1016/j.compositesb.2018.03.019.

627 [21] F.O. Falope, L. Lanzoni, A.M. Tarantino, Double lap shear test on steel fabric reinforced cementitious
628 matrix (SFRCM), Composite Structures. 201 (2018) 503–513.
629 https://doi.org/10.1016/j.compstruct.2018.06.001.

630 [22] A.K. Gupta, S.R. Maestrini, Tension‐Stiffness Model for Reinforced Concrete Bars, Journal of
631 Structural Engineering. 116 (1990) 769–790. https://doi.org/10.1061/(ASCE)0733-9445(1990)116:3(769).

632 [23] C. Ouyang, E. Wollrab, S.M. Kulkarni, S.P. Shah, Prediction of Cracking Response of Reinforced
633 Concrete Tensile Members, Journal of Structural Engineering. 123 (1997) 70–78.
634 https://doi.org/10.1061/(ASCE)0733-9445(1997)123:1(70).

635 [24] J. Roesler, G.H. Paulino, K. Park, C. Gaedicke, Concrete fracture prediction using bilinear softening,
636 Cement and Concrete Composites. 29 (2007) 300–312. https://doi.org/10.1016/j.cemconcomp.2006.12.002.

637 [25] B. Beckmann, K. Schicktanz, D. Reischl, M. Curbach, DEM simulation of concrete fracture and crack
638 evolution, Structural Concrete. 13 (2012) 213–220. https://doi.org/10.1002/suco.201100036.

639 [26] M.H. Aliabadi, A.L. Saleh, Fracture mechanics analysis of cracking in plain and reinforced concrete
640 using the boundary element method, Engineering Fracture Mechanics. 69 (2002) 267–280.
641 https://doi.org/10.1016/S0013-7944(01)00089-3.

642 [27] R. de Borst, J.J.C. Remmers, A. Needleman, M.-A. Abellan, Discrete vs smeared crack models for
643 concrete fracture: bridging the gap, Int. J. Numer. Anal. Meth. Geomech. 28 (2004) 583–607.
644 https://doi.org/10.1002/nag.374.
645 [28] S. Moshirabadi, M. Soltani, Implementation of smeared crack approach in rigid block and spring
646 modeling of reinforced concrete, Engineering Structures. 201 (2019) 109779.
647 https://doi.org/10.1016/j.engstruct.2019.109779.

648 [29] R. Luciano, E. Sacco, A damage model for masonry structures, European Journal of Mechanics -
649 A/Solids. 17 (1998) 285–303. https://doi.org/10.1016/S0997-7538(98)80087-9.

650 [30] A. Rimkus, V. Cervenka, V. Gribniak, J. Cervenka, Uncertainty of the smeared crack model applied
651 to RC beams, Engineering Fracture Mechanics. 233 (2020) 107088.
652 https://doi.org/10.1016/j.engfracmech.2020.107088.

653 [31] R. Barretta, F. Fabbrocino, R. Luciano, F.M. de Sciarra, G. Ruta, Buckling loads of nano-beams in
654 stress-driven nonlocal elasticity, Mechanics of Advanced Materials and Structures. 27 (2020) 869–875.
655 https://doi.org/10.1080/15376494.2018.1501523.

656 [32] A. Pranno, F. Greco, L. Leonetti, P. Lonetti, R. Luciano, U. De Maio, Band gap tuning through
657 microscopic instabilities of compressively loaded lightened nacre-like composite metamaterials, Composite
658 Structures. 282 (2022) 115032. https://doi.org/10.1016/j.compstruct.2021.115032.

659 [33] F. Greco, L. Leonetti, U. De Maio, S. Rudykh, A. Pranno, Macro- and micro-instabilities in
660 incompressible bioinspired composite materials with nacre-like microstructure, Composite Structures. 269
661 (2021) 114004. https://doi.org/10.1016/j.compstruct.2021.114004.

662 [34] U. De Maio, D. Cendón, F. Greco, L. Leonetti, P. Nevone Blasi, J. Planas, Investigation of concrete
663 cracking phenomena by using cohesive fracture-based techniques: A comparison between an embedded crack
664 model and a refined diffuse interface model, Theoretical and Applied Fracture Mechanics. 115 (2021) 103062.
665 https://doi.org/10.1016/j.tafmec.2021.103062.

666 [35] N. Moës, T. Belytschko, Extended finite element method for cohesive crack growth, Engineering
667 Fracture Mechanics. 69 (2002) 813–833. https://doi.org/10.1016/S0013-7944(01)00128-X.

668 [36] A. Skrame, A. Pascuzzo, F. Greco, L. Leonetti, D. Gaetano, Comparative finite element modelling
669 approaches for the seismic vulnerability analysis of historical masonry structures: the case study of the
670 Cathedral of Catanzaro (Italy), IJMRI. 1 (2022) 1. https://doi.org/10.1504/IJMRI.2022.10044503.

671 [37] A. Pascuzzo, A. Yudhanto, M. Alfano, G. Lubineau, On the effect of interfacial patterns on energy
672 dissipation in plastically deforming adhesive bonded ductile sheets, International Journal of Solids and
673 Structures. 198 (2020) 31–40. https://doi.org/10.1016/j.ijsolstr.2020.04.001.

674 [38] D. Ammendolea, F. Greco, P. Lonetti, R. Luciano, A. Pascuzzo, Crack propagation modeling in
675 functionally graded materials using Moving Mesh technique and interaction integral approach, Composite
676 Structures. 269 (2021) 114005. https://doi.org/10.1016/j.compstruct.2021.114005.

677 [39] F. Greco, D. Ammendolea, P. Lonetti, A. Pascuzzo, Crack propagation under thermo-mechanical
678 loadings based on moving mesh strategy, Theoretical and Applied Fracture Mechanics. 114 (2021) 103033.
679 https://doi.org/10.1016/j.tafmec.2021.103033.

680 [40] H.M. Gomes, A.M. Awruch, Some aspects on three-dimensional numerical modelling of reinforced
681 concrete structures using the finite element method, Advances in Engineering Software. 32 (2001) 257–277.
682 https://doi.org/10.1016/S0965-9978(00)00093-4.
683 [41] A. Ziari, M.R. Kianoush, Finite-Element Parametric Study of Bond and Splitting Stresses in
684 Reinforced Concrete Tie Members, J. Struct. Eng. 140 (2014) 04013106.
685 https://doi.org/10.1061/(ASCE)ST.1943-541X.0000903.

686 [42] M. Dehestani, S.S. Mousavi, Modified steel bar model incorporating bond-slip effects for embedded
687 element method, Construction and Building Materials. 81 (2015) 284–290.
688 https://doi.org/10.1016/j.conbuildmat.2015.02.027.

689 [43] L.-Y. Xu, X. Nie, M. Zhou, M.-X. Tao, Whole-process crack width prediction of reinforced concrete
690 structures considering bonding deterioration, Engineering Structures. 142 (2017) 240–254.
691 https://doi.org/10.1016/j.engstruct.2017.03.060.

692 [44] U. De Maio, N. Fantuzzi, F. Greco, L. Leonetti, A. Pranno, Failure Analysis of Ultra High-
693 Performance Fiber-Reinforced Concrete Structures Enhanced with Nanomaterials by Using a Diffuse
694 Cohesive Interface Approach, Nanomaterials. 10 (2020) 1792. https://doi.org/10.3390/nano10091792.

695 [45] F.J. Ma, A.K.H. Kwan, Crack width analysis of reinforced concrete members under flexure by finite
696 element method and crack queuing algorithm, Engineering Structures. 105 (2015) 209–219.
697 https://doi.org/10.1016/j.engstruct.2015.10.012.

698 [46] X. Hu, H. Xu, X. Xi, P. Zhang, S. Yang, Meso-scale phase field modelling of reinforced concrete
699 structures subjected to corrosion of multiple reinforcements, Construction and Building Materials. 321 (2022)
700 126376. https://doi.org/10.1016/j.conbuildmat.2022.126376.

701 [47] C. Wei, C.S. Wojnar, C. Wu, Hydro-chemo-mechanical phase field formulation for corrosion induced
702 cracking in reinforced concrete, Cement and Concrete Research. 144 (2021) 106404.
703 https://doi.org/10.1016/j.cemconres.2021.106404.

704 [48] S. Lee, M.F. Wheeler, T. Wick, Iterative coupling of flow, geomechanics and adaptive phase-field
705 fracture including level-set crack width approaches, Journal of Computational and Applied Mathematics. 314
706 (2017) 40–60. https://doi.org/10.1016/j.cam.2016.10.022.

707 [49] F. Greco, P. Lonetti, R. Luciano, P. Nevone Blasi, A. Pranno, Nonlinear effects in fracture induced
708 failure of compressively loaded fiber reinforced composites, Composite Structures. 189 (2018) 688–699.
709 https://doi.org/10.1016/j.compstruct.2018.01.014.

710 [50] F. Greco, L. Leonetti, A. Pranno, S. Rudykh, Mechanical behavior of bio-inspired nacre-like
711 composites: A hybrid multiscale modeling approach, Composite Structures. 233 (2020) 111625.
712 https://doi.org/10.1016/j.compstruct.2019.111625.

713 [51] F. Greco, L. Leonetti, C.M. Medaglia, R. Penna, A. Pranno, Nonlinear compressive failure analysis of
714 biaxially loaded fiber reinforced materials, Composites Part B: Engineering. 147 (2018) 240–251.
715 https://doi.org/10.1016/j.compositesb.2018.04.006.

716 [52] F. Gusella, F. Cluni, V. Gusella, Homogenization of dynamic behaviour of heterogeneous beams with
717 random Young’s modulus, European Journal of Mechanics - A/Solids. 73 (2019) 260–267.
718 https://doi.org/10.1016/j.euromechsol.2018.09.002.

719 [53] F. Gusella, F. Cluni, V. Gusella, Homogenization of the heterogeneous beam dynamics: The influence
720 of the random Young’s modulus mixing law, Composites Part B: Engineering. 167 (2019) 608–614.
721 https://doi.org/10.1016/j.compositesb.2019.03.025.
722 [54] U. De Maio, F. Greco, L. Leonetti, R. Luciano, P. Nevone Blasi, S. Vantadori, A refined diffuse
723 cohesive approach for the failure analysis in quasibrittle materials—part I: Theoretical formulation and
724 numerical calibration, Fatigue Fract Eng Mater Struct. 43 (2020) 221–241. https://doi.org/10.1111/ffe.13107.

725 [55] U. De Maio, F. Greco, L. Leonetti, R. Luciano, P. Nevone Blasi, S. Vantadori, A refined diffuse
726 cohesive approach for the failure analysis in quasibrittle materials—part II: Application to plain and reinforced
727 concrete structures, Fatigue Fract Eng Mater Struct. 42 (2019) 2764–2781. https://doi.org/10.1111/ffe.13115.

728 [56] A. Pascuzzo, F. Greco, L. Leonetti, P. Lonetti, A. Pranno, C. Ronchei, Investigation of mesh
729 dependency issues in the simulation of crack propagation in quasi‐brittle materials by using a diffuse interface
730 modeling approach, Fatigue Fract Eng Mat Struct. 45 (2022) 801–820. https://doi.org/10.1111/ffe.13635.

731 [57] F. Greco, L. Leonetti, P. Lonetti, R. Luciano, A. Pranno, A multiscale analysis of instability-induced
732 failure mechanisms in fiber-reinforced composite structures via alternative modeling approaches, Composite
733 Structures. 251 (2020) 112529. https://doi.org/10.1016/j.compstruct.2020.112529.

734

View publication stats

You might also like