Home Technology Numerical study of the FRP-concrete bond behavior under thermal variations
Article Open Access

Numerical study of the FRP-concrete bond behavior under thermal variations

  • Rossana Dimitri EMAIL logo , Martina Rinaldi , Francesco Tornabene and Francesco Micelli
Published/Copyright: May 11, 2023

Abstract

In a context where daily and seasonal temperature changes or potential fire exposure can affect the mechanical response of structures strengthened with fiber-reinforced polymer (FRP) composites during their life cycle, the present work studies the bond behavior of FRP laminates glued to concrete substrates under a thermal variation. The problem is tackled computationally by means of a contact algorithm capable of handling both the normal and tangential cohesive responses, accounting for the effect of thermal variations on the interfacial strength and softening parameters, which defines the failure surface and post cracking response of the selected specimen. A parametric investigation is performed systematically to check for the effect of thermo-mechanical adhesive and geometrical properties on the debonding load of the FRP-to-concrete structural system. The computational results are successfully validated against some theoretical predictions from literature, which could serve as potential benchmarks for developing further thermo-mechanical adhesive models, even in a coupled sense, for other reinforcement-to-substrate systems, useful for design purposes in many engineering applications.

1 Introduction

The increased sensitivity toward the conservation of heritage construction and civil infrastructures has favored the study of novel methodologies and the development of innovative materials for structural repair and retrofitting [1,2]. One of the most recent innovations in strengthening existing reinforced concrete (RC) structures is related to the use of composite materials. Among such innovative materials, fiber-reinforced polymer (FRP) composites have gained an increased interest among scientists in both the reinforcement and structural repair fields [3,4,5,6]. The structural compatibility between FRP laminates and concrete substrates can be assured by chemical bonding based on the use of adhesives. In this scenario, the bond between these materials becomes a controlling factor that governs the design of strengthening interventions. It is well known from the experimental literature, that Mode II debonding is the most common failure mode in structures strengthened with externally bonded FRP plates [7,8,9], such that different setup tests have been proposed by scientists to investigate the behavior of FRP-to-substrate bonded joints under a pure Mode II loading condition [10,11,12,13,14,15,16,17,18,19,20,21,22,23,24]. Among the most common standard tests from the literature, there is the single-lap shear test and/or the double-lap shear test to investigate the bond-slip behavior of specimens under quasi-static and monotonic loading conditions, along with the debonding load, the effective bond length, and the full-range bond behavior [25,26,27,28,29,30,31,32,33,34,35,36,37,38]. Based on the existing investigations on bonded joint tests, it is noteworthy that the mechanical properties of both concrete substrates and bonding adhesives can affect the shape of a bond-slip model to yield different representative categories, namely, the elastic brittle [10,26,39,40,41,42], the bi-linear [10,14,43,44,45,46,47,48,49], the elastic–plastic–brittle [41,50,51,52,53], the trapezoidal [2,5457], the rigid-softening [10,39,41,45,46,58], and the nonlinear [12,14,29,31,34,47,5963] models. At the same time, the maximum load seems to be almost unaffected by the bond-slip model, which describes the debonding failure process of specimens in terms of load–deflection response, as found in the study of Hart-Smith [64].

The study of adhesively bonded joints started in 1938 with the pioneering works by Volkersen [65] and Goland and Reissner [66], followed by the research of Hart-Smith [64,66]. In such works, the authors described the shear response of an adhesive layer as elastic–plastic or bilinearly elastic behavior before reaching the ultimate strain, while investigating the effect of mutual interaction among fractures within an RC beam on its final debonding failure under a pure Mode II loading. Later on, a large variety of studies in the literature focused on the full-range load-slip deformation process and debonding mechanism of the FRP-to-concrete interface under a mechanical loading at room temperature. An analytical solution was proposed by Yuan et al. [41] to describe the full-range behavior of the FRP-to-substrate bonded joints by a bi-linear bond-slip model. The same study was conducted by Teng et al. [44], with the addition of a tensile force applied on the other end of the FRP sheet. Later, Chen et al. [58] employed a rigid-softening model to describe the same behavior of Yuan et al. [41]. A trapezoidal bond-slip model was applied by Fernando et al. [57] to find the full-range behavior of FRP-to-substrate bonded joints, starting from the study of a carbon-fiber-reinforced polymer (CFRP) applied on a steel structure and using a ductile nonlinear adhesive [54,56]. Besides, the exponential bond-slip model was used by Dai et al. [33] assuming an infinite bond length, while other forms of the same model were used by Biscaia et al. [42], Caggiano et al. [49], Zhou et al. [59], Cornetti and Carpinteri [60], and Yuan et al. [62]. All these works focused on the debonding behavior of FRP-to-substrate interfaces under a pure shear loading, without evaluating the possible effect of temperature on the interfacial response, as can occur for FRP-strengthened structures exposed to climate changes or fire. The experimental results from the literature indicated that thermal loading could induce two possible effects on the bond behavior of FRP-to-substrate interfaces, along with the stress transfer and load-bearing capacity in the strengthened system. The first effect relies on the thermal-induced property variations of the adhesives and adherends. As detailed in the study of Rabinovitch [67], such an effect is strictly related to the low glass transition temperature (T g) of ambient-cured bonding adhesives. For service temperatures higher than T g, the adhesive transfers from a glassy state to a rubbery one, with a consecutive degradation of its stiffness and strength, along with an overall reduction of the stress interfacial transfer and premature debonding failure of the reinforcement laminate [23,67]. The second effect, instead, is related to a thermal incompatibility between the FRP and substrate, featuring different coefficients of thermal expansion that yield a thermal stress state. More specifically, the FRP thermal expansion in the longitudinal direction can be significantly different from the thermal expansion of the concrete substrate. This difference generates a thermal stress state with a meaningful influence on the bond interfacial strength and the mechanical stress state. Among different bonded joint tests, the most commonly used for experimental investigations about the thermomechanical bond behavior of FRP-to-concrete or FRP-to-steel interfaces rely on single-lap and double-lap shear tests under different thermal conditions [6874].

Since the temperature variation affects both the interfacial stress state and material properties of the adhesive, temperature-dependent interfacial bond-slip models must be properly defined to account for any possible co-presence of interfacial thermal stresses and/or temperature-induced material property variation. In different shear test investigations from literature, these thermal stresses are ignored [71,75]. In the work by Hart-Smith [64], a double-lap joint was loaded both mechanically and thermally, to check for the solution of the interfacial problem in terms of ultimate load instead of a full-range debonding process, based on an elastic–plastic or bilinear-elastic behavior assumption, before the ultimate strain. An analytical solution was proposed by Gao et al. [76], based on a bi-linear bond-slip model, for the study of the full-range behavior of FRP-to-concrete bonded joints under Mode II thermo-mechanical loading. Their theoretical results indicated that the debonding load was highly affected by thermal stresses, making it possible to isolate them from each other. It was also found that Mode II debonding load can be influenced by a varying service temperature. More recently, Gao et al. [77] have found some closed-form analytical solutions for further bond-slip models while analyzing the influence of the bond-slip shape and bond length on both the stress transfer mechanism and debonding process of FRP-to-substrate bonded joints under thermo-mechanical loading.

Starting with the available literature on the topic, the present work examined the effect of the thermo-mechanical and geometrical properties of the adhesives on the mechanical response of chemically bonded FRP-to-concrete joints both in a local and global sense. The problem is tackled numerically based on a mixed-mode interfacial cohesive law within a generalized node-to-segment (NTS) contact algorithm, which accounts for the effect of temperature-induced variations in material properties. The generalized NTS contact algorithm is capable of handling both separation and/or penetration conditions of the adhesive layer between the substrate and reinforcement, in a unified setting. The adhesive portion is modeled here by means of a bilinear traction-separation law, accounting for the interfacial damage evolution for different thermal conditions. A large numerical investigation is performed to check for the effect of the adhesive length and temperature-dependent mechanical properties on the overall bond behavior and debonding failure processes of the jointed specimens, whose results are successfully verified against the experimental predictions from literature.

The work is organized as follows. After this introduction, the theoretical framework of the problem is briefly reviewed in Section 2, both from a thermal and mechanical perspective. The numerical implementation is described in Section 3, where the numerical predictions are compared with a set of experimental data from the literature. The concluding remarks are reported in Section 4, whose general concepts could be extended to nonlinear adhesives or more complicated thermomechanical problems.

2 Theoretical framework

2.1 Definition of the thermomechanical properties

Polymers and their composites can be used under a varying range of thermal conditions (compared to the glass transition temperature T g ), such that it becomes important to evaluate their performances under temperature variations that may produce changes in physical–mechanical properties. It is well known that the mechanical properties of polymers depend on their molecular weight and temperature. Among them, the elastic normal and shear moduli E ( t , T ) and G ( t , T ) depend on time t and temperature T , where the polymer stiffness can vary due to possible molecular rearrangements. In Figure 1, for example, we plot the modulus vs temperature curve referred to a typical polymer with a secondary relaxation and a clear subdivision in four regions. It is worth noticing the glassy zone, where the elastic modulus can present two secondary transitions (named relaxation γ and β , respectively), and a third one, called glassy transition (or transition α ). For higher temperatures, the glassy transition takes place in a transition zone with a decreasing value of the elastic modulus. Here, T g is identified experimentally as the modulus inflecting point considering the temperature. Graphically, it can be obtained as the intersection between the tangents starting from the extreme points of the drastic modulus loss range. For higher temperatures, the rubbery zone is reached. Here the modulus has lower values and keeps constant, as visible in the third zone in Figure 1, with a plateau representing the elastic zone of the rubbery state. A mixed zone takes place, which features a rapid decrease in the elastic modulus. The polymer chain movement is described with the reptation concept, as introduced by De Gennes [78] for the first time in terms of wormlike movements of the chains. In the final viscous flow, the linear polymer becomes a viscous fluid with an abrupt reduction of the stiffness property. This is not true for crosslinked polymers due to their strong chemical bonds, so their modulus keeps almost constant until degradation is reached.

Figure 1 
                  Elastic modulus of a linear polymer subjected to a temperature variation.
Figure 1

Elastic modulus of a linear polymer subjected to a temperature variation.

In order to employ polymer matrix composites in a transition zone, it is necessary to describe analytically the polymer property variation (e.g., the stiffness) caused by a temperature change. Mahieux [79,80] studied this problem and proposed a relation valid for every type of polymer (thermosetting, thermoplastic, amorphous, linear, semicrystalline, and crosslinked), for every molecular weight, and for a full range of temperature. In the rubbery zone, the temperature dependence of the stiffness property E can be defined as follows:

(1) E = ρ RT M C ,

where ρ is the density, RT is the room temperature, and M C is the molecular weight. However, this is true only for temperatures higher than the glass transition temperature T g .

Different theories allow us to define the variation of the elastic properties in the glassy state but only for small variations. Among many, the van Krevelen [81] theory suggests the following relations:

(2) G g ( T ) G g ( T r ) = T g / T r + 2 T g / T r + 2 T / T r for T < T g ,

(3) G C ( T ) G C ( T r ) = exp 2.65 T m / T r T m / T T m / T r 1 for T > T r 100 ,

(4) G SC = G g + x C 2 ( G C G g ) ,

where G is the shear modulus; T is the temperature; and the subscripts g , c, sc, m , and r refer to the glassy, crystalline, semi-crystalline, melting, and reference states, respectively. Also, in this case, note that these equations can be applied only to the glassy state (service temperature lower than T g), and do not refer to the polymer microstructure.

Among different possibilities, in this study, Mahieux’s approach [80] is adopted due to its general nature with respect to other existing theories. Such approach, indeed, describes the polymer behavior on a full-range temperature, including the transitions, and relates the polymer mechanical response to its microstructure. For a consistent definition of the local behavior for such material, during its relaxation time, spring and dashpots (dampers that withstand displacement through viscous friction) are considered in the model, such that the Young’s modulus E ( t ) varies with time t as follows:

(5) E ( t ) = i = 1 N E i e t / τ i ,

where τ i is the relaxation time. Such equation can be written in an integral form as

(6) E ( t ) = + H ( τ ) e t / τ d ( ln τ ) ,

where H is a distribution function, defined as

(7) H ( τ ) = τ E ( τ ) .

In the viscous region, the polymer reptation is usually modeled as follows:

(8) E e Q / ( RT ) ,

with Q being the activation energy of the process. The Boltzmann distribution denotes a casual process, so that the bonds are broken following an indeterminate sequence, while the Weibull distribution allows us to describe an interactive fracturing process. Note that bond breaking influences the other bonds: for example, the secondary transitions (Figure 1) result from highly localized molecular movements. Therefore, the secondary bonds must be broken to allow movement of a lateral group or a few principal chains. Due to the atomic distance variation, which induces an interaction force distribution, the secondary bonds break in different moments. Compared to the Boltzmann distribution, the Weibull distribution seems to provide a more reliable mathematical representation of the debonding process and accounts for the secondary transition β . The mechanical response is given by the movement of little chain segments (monomers). Given the high number and force of the bonds involved in this relaxation process, it is possible to associate a Weibull coefficient m 1 with transition β , so that Eq. (5) becomes

(9) H 1 ( τ ) = e ( t / τ 1 ) m 1 .

By substitution of time τ 1 with the characteristic instantaneous temperature and by introducing a conversion constant, we have

(10) H 1 ( T ) = H 1 0 e ( T / β 1 ) m 1 ,

where β 1 is a characteristic temperature (the transition temperature beta) and H 1 0 beta transition (i.e., the relaxation entity). The same approach can be applied to the other transitions for an improved number of segments involved in the relaxation stage. A novel Weibull coefficient can be associated with each stage, which refers to the number of intermolecular bonds involved. For this reason, a more general law can be applied to all transformation stages by adding all their components

(11) E = i = 1 N H i e ( T i / T ref , i ) m i .

Commercial polymers typically feature from one up to three transitions ( 1 N 3 ), whose coefficients H i (magnitude of phase transition) can be defined with different methods. Hereafter, we assume the following thermal dependence of the elastic modulus [79], as plotted in Figure 2.

  1. For materials that are not subjected to any transition before becoming rubbery/viscous

    (12) E = E 3 e ( T / T 3 ) m 3 .

  2. For materials featuring two transitions

    (13) E = ( E 2 E 3 ) e ( T / T 2 ) m 2 + E 3 e ( T / T 3 ) m 3 .

  3. For materials featuring three transitions

(14) E = ( E 1 E 2 ) e ( T / T 1 ) m 1 + ( E 2 E 3 ) e ( T / T 2 ) m 2 + E 3 e ( T / T 3 ) m 3 .

Figure 2 
                  Definition of the input parameters for the 
                        
                           
                           
                              E
                              −
                              T
                           
                           E-T
                        
                      relation.
Figure 2

Definition of the input parameters for the E T relation.

In these three relations, temperatures ( T i ) are expressed in Kelvin (K) and refer to each transition phase depicted in Figure 2. The thermal values can be defined considering the maximum points of the curve resulting from a dynamic mechanic analysis or considering the inflection points of a differential scanning calorimetry curve. At the same time, E i corresponds to the instantaneous stiffness modulus assumed by the material at the beginning of a transition phase: E 1 is the instantaneous modulus of the polymer for low temperatures, E 2 is the instantaneous modulus immediately after the beta transition, and E 3 is the instantaneous modulus at the beginning of the rubbery state (Figure 2). Moreover, the parameters m i stand for the Weibull coefficients related to the stress distribution during a debonding phenomenon. The value of m is small if there is a wide distribution of the bonding forces, and it becomes high for homogeneous materials with a limited distribution of bonding forces, or amorphous materials. Moreover, the m value depends on the impediment degree of a molecular motion (i.e., crosslinking, molecular weight, crystallinity): if the motion is limited, m assumes a low value; if the motion is strictly limited, m becomes much lower, getting a similar behavior to the Boltzmann distribution. In the viscous flow and for cross-links, the gradient slope decreases for an increased cross-linking degree. For highly cross-linked materials, the viscous flow can disappear.

2.2 Interfacial constitutive law and finite-element algorithm

For a numerical description of the mixed mode debonding process, an uncoupled cohesive law is here assumed to model the adhesive interface within the FRP-to-concrete joint both in the normal and tangential directions. An uncoupled cohesive zone model is here selected, as typically used when the interface separation is enforced to occur in a single predefined direction, such that either Mode I and II separations occur. This means that the normal and tangential stresses, p N and p T , depend on the relatively normal and tangential displacement, g N , and g T , separately.

This choice derives from the possibility of using different values of interfacial fracture energies for both Mode I and Mode II, in line with the experimental evidence. For a compressive state in the normal direction (i.e., for g N < 0 and p N < 0 ), we enforce a non-penetration condition by using a Penalty method, within a generalized contact algorithm, based on the minimization of a modified potential W ¯ = W + W C , which involves both the elastic part W and contact part, W C = ε N g N , with ε N being the Penalty parameter. We select a bilinear cohesive traction-separation law for the normal and shear directions (Figure 3) because of its capability to simply define the main three interfacial properties, namely, the fracture energies, the cohesive strengths p N max and p T max , and the elastic stiffness [82]. For such simple reasons, the bilinear model is largely used, also in design guidelines, to model the FRP-to-brittle substrate interfaces [83] in lieu of more complicated nonlinear models. Thus, the energy release rates for Modes I and II correspond to the areas under the respective curves, integrated up to the current values of g N and g T [84], such that a mixed mode fracture criterion must be properly introduced to govern a failure process. In the present work, we selected one of the simplest possible criteria to handle the fracturing process, i.e.,

(15) G I G Ic + G II G IIc = 1 ,

where G IC and G IIC stand for the fracture energies in pure Modes I and II, respectively.

Figure 3 
                  Relation at the interface between the traction and the relative displacement: (a) normal direction and (b) tangential direction.
Figure 3

Relation at the interface between the traction and the relative displacement: (a) normal direction and (b) tangential direction.

A generalized NTS contact element is adopted to handle the interfacial problem for both the normal and tangential forces within a unified formulation, as also described in the study of Dimitri et al. [24]. Based on the contact status, an automatic switching procedure is applied to select a cohesive or contact model. Each element contribution, for both cohesive and contact forces, is automatically added to the virtual global work as

(16) δ W C = F N δ g N + F T δ g T ,

where F N = p N A and F T = p T A stand for the normal and tangential interfacial forces, and A is the contact area associated with each contact element.

Figure 4 depicts the geometrical scheme of the selected FRP-to-concrete pull-out test. From a numerical standpoint, the adherend is discretized with 2D linearly elastic beam elements, while the concrete substrate is modeled with plane-stress four-node isoparametric elastic elements.

Figure 4 
                  Setup of the single-lap shear test.
Figure 4

Setup of the single-lap shear test.

The nonlinear problem is solved with a Newton–Raphson iterative scheme, where the global tangent stiffness matrix is consistently obtained from a linearization of all the terms in Eq. (16). Such linearization yields the following relation:

(17) Δ δ W C = F N g N Δ g N + F N g T Δ g T δ g N + F T g T Δ g N + F T g T Δ g T δ g T + F N Δδ g N + F T Δδ g T ,

where δ and Δ refer to the virtual variation and linearization, respectively. For more details about the geometrical quantities δ g N , δ g T , Δ g N , Δ g T , Δδ g N , and Δδ g T , the reader is referred to Ref. [85]. The partial derivatives of the normal and shear forces with respect to the normal and shear relative displacements depend on the cohesive law parameters. According to the selected traction separation law (Figure 3), it is

(18) F N g N = ε N A for g N < 0 , p N max g N max A for 0 g N < g N max , p N max g Nu g N max A for g N max g N < g Nu ,

(19) F T g T = p T max g T max A for g T < g T max , p T max g Tu g T max A for g T max g T < g Tu ,

(20) F N g T = F T g N = 0 .

3 Numerical analysis

A large numerical investigation is now performed for the pull-out test set-up illustrated in Figure 4, whose mechanical properties are here selected in line with the experimental tests performed in the study of Takeo et al. [86], for comparative purposes. A parametric study aims at analyzing the sensitivity of the mechanical response to different geometrical and mechanical properties of the FRP reinforcement and adhesive phases, accounting for different possible thermal conditions. The analysis starts considering the geometrical parameters in Table 1, where l , h , and b c stand for the length, height, and width of the concrete substrate, respectively, and t f and b f stand for the reinforcement thickness and width, respectively. The initial specimen accounts for a concrete substrate with elastic modulus E c = 32.8  GPa and reinforced with an FRP sheet with elastic modulus E f = 230  GPa, which is tested, in displacement-control mode, up to a total displacement d = 0.35 mm (Table 1). As far as the adhesive layer is concerned, the main interfacial parameters are taken as summarized in Table 2. More specifically, we assume an elastic stiffness k N = 2 , 600 MPa and k T = 1 , 100  MPa in the normal and shear direction, respectively, a Penalty parameter in compression ε N = 3 , 000  MPa, ultimate values of the normal and shear relative displacements g Nu = 0.2 mm and g Tu = 0.4 mm , and cohesive strengths p N max = 40 MPa and p T max = 120 MPa , in line with the experimental findings by Takeo et al. [86]. Hereafter, the same example is repeated systematically to study the effect of the bond length l , the elastic modulus of the reinforcing phase, E f , and the reinforcement thickness and length, t f , b f , respectively, on the overall response of the specimen in terms of load–displacement curves and peak load.

Table 1

Geometric and mechanical data

l [ mm] h [ mm] b c [ mm] t f [ mm] b f [ mm] E c [MPa] E f [ MPa] d [ mm]
150 150 150 0.167 100 32,800 2,30,000 0.35
Table 2

Adhesive data implemented in the code

k N [ MPa] k T [ MPa] p N max [ MPa] p T max [ MPa] g Nu [ mm] g Tu [ mm] ε N [ MPa]
2,600 1,100 40 120 0.2 0.4 3,000

3.1 Effect of bond length

The numerical test is repeated for an increased value of the bond length l from 50 up to 400 mm in steps of 50 mm, keeping the other parameters constant, as summarized in Table 3 and Figure 5, together with the corresponding peak load P max . It is worth observing that an increased bond length of the FRP on the substrate up to 200 mm yields larger peak loads, beyond which it remains almost constant, in line with findings by Takeo et al. [86] and Ueda et al. [87]. Similar comments can be repeated by looking at the force–displacement curves in Figure 6, which clearly show an increased initial stiffness (in the ascending branch together with an increased value of the peak load and an increased displacement capacity of the specimen, because of an improved transfer of local interfacial stresses among FRP and substrate in the bonded portion, especially in the shear direction. It is expected, indeed, that an increased bond length increases the Mode II stress components, in lieu of decreased values of Mode I stresses, and an overall decrease of the interfacial mixed-mode stress distribution within the specimen.

Table 3

Variation of the maximum load for different bond lengths

l [ mm] P max [ N] l [ mm] P max [ N]
50 5,723.2 250 13,488
100 10,106 300 13,533
150 12,466 350 13,548
200 13,287 400 13,555
Figure 5 
                  Variation of the maximum load with the bond length.
Figure 5

Variation of the maximum load with the bond length.

Figure 6 
                  Load–displacement curves for different bond lengths.
Figure 6

Load–displacement curves for different bond lengths.

3.2 Effect of the reinforcement elastic modulus

The analysis has been extended by considering three different types of composite reinforcement, namely, a glass FRP with an elastic modulus E f = 100 GPa (labeled as GFRP100), an aramid FRP with E f = 150 GPa (labeled as AFRP150), and a carbon-based reinforcement (CFRP) with a different elastic modulus E f of 190, 210, 230, 250, 390 and 450 GPa, while keeping all the other parameters constant.

In Table 4 and Figure 7, the variation of the maximum load for a varying elastic stiffness of the FRP strip is summarized. As also expected experimentally in the study of Wu et al. [88], the maximum load significantly increases for an increased elastic modulus E f , reaching the highest values for a CFRP with E f = 390 ÷ 450 GPa . In these last cases, however, the numerical values of P max are more conservative than expected experimentally by Wu et al. [88], because of the adhesive crisis estimations in lieu of their cohesive experimental results found in [88]. This can be addressed to the local behavior of the concrete which may exhibit higher properties with respect to the theoretical prediction. The overall load–displacement response of the specimen is, thus, plotted in Figure 8 for different FRPs. Based on these plots, note that as the elastic modulus E f of the FRP increases, the slope of the curves in the ascending branch increases, together with an increased peak strength P max and interfacial brittleness in the descending branch of the curves.

Table 4

Variation of the maximum load for different reinforcement elastic moduli

E f [ GPa] P max [ N] E f [ GPa] P max [ N]
100 8,863.9 230 12,466
150 10,625 250 12,799
190 11,663 390 14,362
210 12,090 450 14,775
Figure 7 
                  Variation of the maximum load with the elastic modulus of the reinforcement.
Figure 7

Variation of the maximum load with the elastic modulus of the reinforcement.

Figure 8 
                  Load–displacement curves for different fiber elastic moduli.
Figure 8

Load–displacement curves for different fiber elastic moduli.

3.3 Effect of the reinforcement thickness

We now assume six different FRP thicknesses t f , i.e., 0.167, 0.222, 0.247, 0.337, 0.55, and 1.4 mm, as commonly used in structural reinforcements. All the other parameters are kept constant. The results are plotted in Figures 9 and 10 and could be compared to the experimental predictions by Ueda et al. [87] and Tan [89]. Even in this case, an increased value of the fiber thickness corresponds to an increased value of the peak load of 7 ÷ 8.5 % , see also the results in Table 5. As also plotted in Figure 10, the global load–displacement curves increase their initial slope in the ascending branch for an increased reinforcement thickness. At the same time, the maximum strength of the specimen tends to increase, and the specimen tends to assume an elastic-brittle behavior up to the final collapse, for an increased reinforcement thickness.

Figure 9 
                  Variation of the maximum load with the fiber thickness.
Figure 9

Variation of the maximum load with the fiber thickness.

Figure 10 
                  Load–displacement curves for different fiber thicknesses.
Figure 10

Load–displacement curves for different fiber thicknesses.

Table 5

Variation of the maximum load with the fiber thickness

t f [ mm] P max [ N] t f [ mm] P max [ N]
0.167 12,466 0.337 14,858
0.222 13,553 0.55 14,944
0.247 13,926 1.4 17,066

3.4 Effect of the FRP width

The reinforcement width represents another key parameter that can significantly affect the global mechanical behavior of the adhesive joint. A larger dimension of the reinforcing FRP, indeed, can improve the interfacial distribution of stresses, and the related maximum debonding load of the specimen, whose sensitivity is studied here, by selecting the following combinations:

  1. b c = 150 mm and b f 50, 60, 80 and 100 mm;

  2. b c = 200 mm and b f 120 and 150 mm; and

  3. b c = 400 mm and b f 200 and 300 mm,

while keeping all the other parameters constant.

As summarized in Table 6 and plotted in Figure 11, the maximum load P max increases monotonically for higher values of b f , with increased structural stiffness and decreased ductility, as visible in the global load–displacement response in Figure 12. An increased interfacial stiffness, indeed, is expected to yield a more localized distribution of the cohesive stresses in the bond length, with an increased overall fragility of the specimen. Even in this case, the numerical insights are comparable with the experimental findings by Ueda et al. [87] and Tan [89].

Table 6

Variation of the maximum load with the fiber lengths

b f [ mm] P max [ N] b f [ mm] P max [ N]
50 9,457.5 120 13,180
60 10,256 150 14,018
80 11,525 200 14,897
100 12,466 300 15,843
Figure 11 
                  Variation of the maximum load with the reinforcement width.
Figure 11

Variation of the maximum load with the reinforcement width.

Figure 12 
                  Load–displacement curves for different reinforcement widths.
Figure 12

Load–displacement curves for different reinforcement widths.

3.5 Effect of temperature

We now analyze the possible effect of temperature on the mechanical bond properties of concrete members externally strengthened with FRPs while selecting an epoxy adhesive with a glass transition temperature T g ranging between 50 and 80°C . Epoxy adhesives are noteworthy for their susceptibility to mechanical degradation when exposed to high temperatures. Since adhesives are responsible for the bonding behavior among FRPs and concrete, for bond-critical conditions, the structural efficiency of the strengthening system could be highly influenced during a high-temperature exposure [90]. Based on the experimental observations from literature on shear tests, when the temperature in the FRP exceeds 200°C , the reinforcement does not contribute anymore to the load-bearing capacity, because of a softening bonding adhesion, with a general loss of the adhesive transfer between materials and a premature detachment of the FRP from the adherend. Due to the low glass transition temperature of adhesives, their elastic modulus decays drastically at temperatures slightly higher than the ambient ones. In such a context, it is important to define a theoretical correlation between temperature and the global response of the reinforced specimen. Two adhesives are considered, labeled as “Adhesive 1” and “Adhesive 2,” with elastic moduli of 4,157 MPa and 2,595 MPa, respectively. Both adhesives feature two different transition phases, moving from the glassy zone to the rubbery state, such that they lose any reserve of resistance for temperatures higher than 150 200°C . In what follows, we refer to the relation proposed by Mahieux [79] to define the variation of the elastic stiffness E, with temperature T

(21) E = ( E 2 E 3 ) e ( T / T 2 ) m 2 + E 3 e ( T / T 3 ) m 3 ,

where T i refers to temperatures at each transition, E i is the instantaneous stiffness of the adhesive at the beginning of each plateau region, and m i are the Weibull coefficients, which account for the microstructure of a polymer and the statistics of the bond breakage. More specifically, E 2 and E 3 stand for the instantaneous moduli right after the beta transition and at the beginning of the rubbery plateau, which occur at temperatures T 2 and T 3 , respectively. Based on the experimental data in previous studies [80,91,92], we set high values of m i for amorphous adhesives with homogeneous properties (see Adhesive 1) and narrow distribution of bond strengths, while setting lower values of m i for crosslinked adhesives with a reduced slope of the drop in the viscous flow region (see Adhesive 2). All the selected parameters for Eq. (21) are reported in Table 7, whereas Figures 13 and 14 show the theoretical and experimental variation of the stiffness with temperature, with a good matching among results. Based on a comparative evaluation of the plots in Figure 13a and b, it is worth observing the different behavior in the glass transition zone, with a rapid modulus drop for “Adhesive 1” in lieu of a gradual reduction with temperature for “Adhesive 2,” thus leading to a different behavior of the two adhesives in the rubbery state. Starting with the theoretical variation of the stiffness properties in Eq. (21) for the adhesive layer, we now analyze the sensitivity of the debonding response for the selected pull-out tests, both in terms of maximum load and global time histories. The systematic investigation accounts for the normal and shear elastic stiffness at different temperatures as summarized in Table 8 for both adhesives, whose properties are responsible for the slopes in the ascending branch of the interfacial model in Modes I and II. The numerical tests are performed in displacement control mode.

Table 7

Input parameters

E 2 [ MPa] E 3 [ MPa] T 2 [ °C] T 3 [ °C] m 2 m 3
Adhesive 1 4,156 300 64 120 60 40
Adhesive 2 2,595 150 128 183 20 20
Figure 13 
                  Variation of the elastic modulus of resins with temperature: (a) Adhesive 1 and (b) Adhesive 2.
Figure 13

Variation of the elastic modulus of resins with temperature: (a) Adhesive 1 and (b) Adhesive 2.

Figure 14 
                  Load–displacement response for different temperatures: (a) Adhesive 1 and (b) Adhesive 2.
Figure 14

Load–displacement response for different temperatures: (a) Adhesive 1 and (b) Adhesive 2.

Table 8

Variation of the elastic properties of the adhesives with the temperature

T [ °C] E [ MPa] G [ MPa]
Adhesive 1
25 4,154 1,731
35 4,139 1,724
45 4,039 1,683
55 3,466 1,444
65 1,467 611
75 302 126
85 293 122
95 279 116
115 165 69
125 57 24
Adhesive 2
25 2,589 1,079
40 2,578 1,074
60 2,536 1,057
90 2,281 950
110 1,786 744
125 1,174 489
140 534 222
160 129 54

Figure 14 shows the load–displacement curves for Adhesive 1 (Figure 14a), and Adhesive 2 (Figure 14b), with a noteworthy variation of the global response for an increased temperature, both in the ascending and softening branches. An increased temperature, indeed, drastically reduces the elastic stiffness and the maximum load of all the specimens, while extending the elastic stage for all the specimens, especially for temperatures higher than T g , when adhesives reach the rubbery state, due to an increased molecular motion and overall deformability. At the same time, we summarize the variation of P max with temperature in Table 9, whose results are also plotted in Figure 15 for both adhesives, with a maximum reduction of about 78% for T = 125 ° and T = 160 ° for Adhesives 1 and 2, respectively. Note that P max remains almost constant for T < T g and decreases monotonically when the glass transition takes place, as also expected from experimental evidences. Specimens based on the second adhesive feature a more gradual decay of P max compared to the first adhesive, thus reflecting in a global sense, the local decay of the interfacial properties. Moreover, for “Adhesive 2”, P max starts decreasing for a higher temperature compared to that one based on “Adhesive 1.” For affinity reasons among the global response of P max vs T plotted in Figure 15 and the local properties of adhesives in Figure 13, we propose the following relation among the debonding load P max and temperature T , which could serve as a useful theoretical tool for design purposes, namely

(22) P max ( T ) = i = 1 N α F i e ( T i / T ref , i ) β i ,

where α i and β i are parameters that depend on the transition states to which the adhesive is subjected, F i is a force parameter for a specific transition state occurring at temperature T i , and T ref , i is a reference temperature. This formulation can be considered as a starting point for future theoretical and experimental developments on the topic, in order to assess and improve the durability of glued joints, and to improve the load-bearing capacities of strengthened structural elements.

Table 9

Variation of the maximum load with temperature

T [ °C] P max [ N]
Adhesive 1
25 12,634
35 12,633
45 12,624
55 12,575
65 10,825
75 7,251.4
85 6,588.4
95 6,588.6
115 4,738.2
125 2,823.1
Adhesive 2
25 12,458
40 12,456
60 12,449
90 12,404
110 10,912
125 9,190.7
140 7,173.5
160 4,011.5
Figure 15 
                  Variation of the maximum load with temperature.
Figure 15

Variation of the maximum load with temperature.

4 Conclusions

A novel research has been carried out on the numerical modeling of the FRP-concrete bond behavior in single shear tests, subjected to a thermal variation. The problem has been approached computationally based on a contact NTS algorithm capable of handling a mixed-mode failure within a unified formulation, accounting for the effect of thermal variations on the interfacial properties of adhesives. A bilinear traction-separation law is proposed here to model the adhesive layer, which accounts for the interfacial damage evolution, under different thermal conditions. A large numerical investigation was performed to check for the effect of the adhesive geometrical properties and temperature-dependent mechanical properties on the overall bond behavior and failure modes of the selected specimens. The numerical results have been successfully verified against the experimental predictions available from the literature. Despite the simplicity of the proposed algorithm, it appears capable of capturing different aspects of the physical behavior, namely, the variation of the stiffness, displacement capacity, and maximum load for a varying bond length, FRP thickness, width, and mechanical properties of the materials. The pioneering research by Mahieux and Reifsnider [80] has been considered for the theoretical relationship between the elastic modulus of polymers and temperature within the NTS formulation in order to check for the sensitivity of the global response in terms of time history and maximum capacity for temperature-dependent adhesive properties. Based on the results, a temperature increase over the transition level, decreases the maximum load and stiffness, due to a local variation of primary and secondary bonds within polymer adhesives at a microscale, together with a varied molecular motion and decay of their elastic stiffness. A theoretical model has been, also, proposed as a useful design tool to relate the debonding load of jointed specimens and temperature, which consistently reflects, at a macroscale level, the physical and mechanical properties of adhesive resins at the microscale.

  1. Funding information: The authors state no funding involved.

  2. Author contributions: All authors have accepted responsibility for the entire content of this manuscript and approved its submission.

  3. Conflict of interest: Rossana Dimitri and Francesco Tornabene, who are the co-authors of this article, are current Editorial Board members of Curved and Layered Structures. This fact did not affect the peer-review process. The authors declare no other conflict of interest.

References

[1] Pepe M, Mazaheripour H, Barros J, Sena-Cruz J, Martinelli E. Numerical calibration of bond law for GFRP bars embedded in steel fibre-reinforced self-compacting concrete. Compos B Eng. 2013;50:403–12.10.1016/j.compositesb.2013.03.006Search in Google Scholar

[2] Ghiassi B, Marcari G, Oliveira DV, Lourenço PB. Numerical analysis of bond behavior between masonry bricks and composite materials. Eng Struct. 2012;43:210–20.10.1016/j.engstruct.2012.05.022Search in Google Scholar

[3] Teng JG, Chen JF, Smith ST, Lam L. FRP: Strengthened RC structures. West Sussex, UK: John Wiley and Sons Ltd; 2002. p. 31–46.Search in Google Scholar

[4] Barros J, Dias SJ, Lima JL. Efficacy of CFRP-based techniques for the flexural and shear strengthening of concrete beams. Cem Concr Compos. 2007;29(3):203–17.10.1016/j.cemconcomp.2006.09.001Search in Google Scholar

[5] Hollaway LC, Teng JG. Strengthening and rehabilitation of civil infrastructures using fiber-reinforced polymer (FRP) composites. Sawston, UK: Woodhead Publishing Ltd.; 2008.10.1533/9781845694890Search in Google Scholar

[6] Teng JG, Yu T, Fernando D. Strengthening of steel structures with fiber-reinforced polymer composites. J Constr Steel Res. 2012;78:131–43.10.1016/j.jcsr.2012.06.011Search in Google Scholar

[7] Teng JG, Smith ST, Yao J, Chen JF. Intermediate crack-induced debonding in RC beams and slabs. Constr Build Mater. 2003;17(6–7):447–62.10.1016/S0950-0618(03)00043-6Search in Google Scholar

[8] Teng JG, Chen JF. Mechanics of debonding in FRP-plated RC beams. Proc Inst Civ Eng: Struct Build. 2009;162(5):335–45.10.1680/stbu.2009.162.5.335Search in Google Scholar

[9] Realfonzo R, Martinelli E, Napoli A, Nunziata B. Experimental investigation of the mechanical connection between FRP laminates and concrete. Compos B Eng. 2013;45(1):341–55.10.1016/j.compositesb.2012.05.010Search in Google Scholar

[10] Täljsten B. Strengthening of concrete prisms using the plate-bonding technique. Int J Fract. 1996;82(3):253–66.10.1007/BF00013161Search in Google Scholar

[11] Chen JF, Teng JG. Anchorage strength models for FRP and steel plates bonded to concrete. J Struct Eng. 2001;127(7):784–91.10.1061/(ASCE)0733-9445(2001)127:7(784)Search in Google Scholar

[12] Dai JG, Ueda T. Local bond stress slip relations for FRP sheets-concrete interfaces. In: Tan KH, editor. Fibre Reinforced Polymer Reinforcement for Concrete Structures (In 2 Volumes): Proceedings of the Sixth International Symposium on FRP Reinforcement for Concrete Structures (FRPRCS-6); 2003 Jul 2–10; Singapore. World Scientific, 2003. p. 143–52.10.1142/9789812704863_0011Search in Google Scholar

[13] Dai J, Ueda T, Sato Y. Development of the nonlinear bond stress–slip model of fiber reinforced plastics sheet–concrete interfaces with a simple method. J Compos Constr. 2005;9(1):52–62.10.1061/(ASCE)1090-0268(2005)9:1(52)Search in Google Scholar

[14] Lu XZ, Teng JG, Ye LP, Jiang JJ. Bond–slip models for FRP sheets/plates bonded to concrete. Eng Struct. 2005;27(6):920–37.10.1016/j.engstruct.2005.01.014Search in Google Scholar

[15] Lu XZ, Ye LP, Teng JG, Jiang JJ. Meso-scale finite element model for FRP sheets/plates bonded to concrete. Eng Struct. 2005;27(4):564–75.10.1016/j.engstruct.2004.11.015Search in Google Scholar

[16] Fawzia S, Al-Mahaidi R, Zhao XL. Experimental and finite element analysis of a double strap joint between steel plates and normal modulus CFRP. Compos Struct. 2006;75(1–4):156–62.10.1016/j.compstruct.2006.04.038Search in Google Scholar

[17] Leung CK, Yang Y. Energy-based modeling approach for debonding of FRP plate from concrete substrate. J Eng Mech. 2006;132(6):583–93.10.1061/(ASCE)0733-9399(2006)132:6(583)Search in Google Scholar

[18] Seracino R, Raizal Saifulnaz MR, Oehlers DJ. Generic debonding resistance of EB and NSM plate-to-concrete joints. J Compos Constr. 2007;11(1):62–70.10.1061/(ASCE)1090-0268(2007)11:1(62)Search in Google Scholar

[19] Bocciarelli M, Colombi P, Fava G, Poggi C. Prediction of debonding strength of tensile steel/CFRP joints using fracture mechanics and stress based criteria. Eng Fract Mech. 2009;76(2):299–313.10.1016/j.engfracmech.2008.10.005Search in Google Scholar

[20] Baky HA, Ebead UA, Neale KW. Nonlinear micromechanics-based bond–slip model for FRP/concrete interfaces. Eng Struct. 2012;39:11–23.10.1016/j.engstruct.2012.01.010Search in Google Scholar

[21] Kashyap J, Willis CR, Griffith MC, Ingham JM, Masia MJ. Debonding resistance of FRP-to-clay brick masonry joints. Eng Struct. 2012;41:186–98.10.1016/j.engstruct.2012.03.032Search in Google Scholar

[22] Biolzi L, Ghittoni C, Fedele R, Rosati G. Experimental and theoretical issues in FRP-concrete bonding. Constr Build Mater. 2013;41:182–90.10.1016/j.conbuildmat.2012.11.082Search in Google Scholar

[23] Dai JG, Gao WY, Teng JG. Bond-slip model for FRP laminates externally bonded to concrete at elevated temperature. J Compos Const. 2013;17(2):217–28.10.1061/(ASCE)CC.1943-5614.0000337Search in Google Scholar

[24] Dimitri R, Trullo M, De Lorenzis L, Zavarise G. Coupled cohesive zone models for mixed-mode fracture: A comparative study. Eng Fract Mech. 2015;148:145–79.10.1016/j.engfracmech.2015.09.029Search in Google Scholar

[25] Chajes MJ, Finch WW, Thomson TA. Bond and force transfer of composite-material plates bonded to concrete. Struct J. 1996;93(2):209–17.10.14359/1491Search in Google Scholar

[26] Maeda T. A study on bond mechanism of carbon fiber sheet. Proceeding of the Third International Symposium of Non-metallic (FRP) Reinforced for Concrete Structures. Vol. 1. Sapporo, Japan: Japan Concrete Institute; 1997. p. 279–86.Search in Google Scholar

[27] Brosens K, Van Gemert D. Anchorage design for externally bonded carbon fiber reinforced polymer laminates. Spec Publ. 1999;188:635–46.Search in Google Scholar

[28] Bizindavyi L, Neale KW. Transfer lengths and bond strengths for composites bonded to concrete. J Compos Constr. 1999;3(4):153–60.10.1061/(ASCE)1090-0268(1999)3:4(153)Search in Google Scholar

[29] Focacci F, Nanni A, Bakis CE. Local bond-slip relationship for FRP reinforcement in concrete. J Compos Constr. 2000;4(1):24–31.10.1061/(ASCE)1090-0268(2000)4:1(24)Search in Google Scholar

[30] Brosens K. Anchoring of externally bonded steel plates and CFRP laminates for the strengthening of concrete elements [dissertation]. Leuven: Katholieke Universiteit Leuven; 2001.Search in Google Scholar

[31] Nakaba K, Kanakubo T, Furuta T, Yoshizawa H. Bond behavior between fiber-reinforced polymer laminates and concrete. Struct J. 2001;98(3):359–67.10.14359/10224Search in Google Scholar

[32] Yao J, Teng JG, Chen JF. Experimental study on FRP-to-concrete bonded joints. Compos B Eng. 2005;36(2):99–113.10.1016/j.compositesb.2004.06.001Search in Google Scholar

[33] Dai J, Ueda T, Sato Y. Unified analytical approaches for determining shear bond characteristics of FRP-concrete interfaces through pullout tests. J Adv Concr Tehnol. 2006;4(1):133–45.10.3151/jact.4.133Search in Google Scholar

[34] Ferracuti B, Savoia M, Mazzotti C. Interface law for FRP–concrete delamination. Compos Struct. 2007;80(4):523–31.10.1016/j.compstruct.2006.07.001Search in Google Scholar

[35] Ferrier E, Quiertant M, Benzarti K, Hamelin P. Influence of the properties of externally bonded CFRP on the shear behavior of concrete/composite adhesive joints. Compos B Eng. 2010;41(5):354–62.10.1016/j.compositesb.2010.03.007Search in Google Scholar

[36] Grande E, Imbimbo M, Sacco E. Simple model for bond behavior of masonry elements strengthened with FRP. J Compos Constr. 2011;15(3):354–63.10.1061/(ASCE)CC.1943-5614.0000170Search in Google Scholar

[37] Obaidat YT, Heyden S, Dahlblom O. Evaluation of parameters of bond action between FRP and concrete. J Compos Constr. 2013;17(5):626–35.10.1061/(ASCE)CC.1943-5614.0000378Search in Google Scholar

[38] Wu YF, Jiang C. Quantification of bond-slip relationship for externally bonded FRP-to-concrete joints. J Compos Constr. 2013;17(5):673–86.10.1061/(ASCE)CC.1943-5614.0000375Search in Google Scholar

[39] Yuan H, Wu Z, Yoshizawa H. Theoretical solutions on interfacial stress transfer of externally bonded steel/composite laminates. Doboku Gakkai Ronbunshu. 2001;2001(675):27–39.10.2208/jscej.2001.675_27Search in Google Scholar

[40] Wu Z, Yuan H, Niu H. Stress transfer and fracture propagation in different kinds of adhesive joints. J Eng Mech. 2002;128(5):562–73.10.1061/(ASCE)0733-9399(2002)128:5(562)Search in Google Scholar

[41] Yuan H, Teng JG, Seracino R, Wu ZS, Yao J. Full-range behavior of FRP-to-concrete bonded joints. Eng Struct. 2004;26(5):553–65.Search in Google Scholar

[42] Biscaia HC, Chastre C, Silva MA. Linear and nonlinear analysis of bond-slip models for interfaces between FRP composites and concrete. Compos B Eng. 2013;45(1):1554–568.10.1016/j.compositesb.2012.08.011Search in Google Scholar

[43] Brosens K, Van, Gemert D. Plate end shear design for external CFRP laminates. Aedificatio Publishers. Fract Mech Concr Struct. 1998;3:1793–804.Search in Google Scholar

[44] Teng JG, Yuan H, Chen JF. FRP-to-concrete interfaces between two adjacent cracks: theoretical model for debonding failure. Int J Solids Struct. 2006;43(18–19):5750–78.10.1016/j.ijsolstr.2005.07.023Search in Google Scholar

[45] Ali MM, Oehlers DJ, Seracino R. Vertical shear interaction model between external FRP transverse plates and internal steel stirrups. Eng Struct. 2006;28(3):381–9.10.1016/j.engstruct.2005.08.010Search in Google Scholar

[46] Ali MM, Oehlers DJ, Griffith MC, Seracino R. Interfacial stress transfer of near surface-mounted FRP-to-concrete joints. Eng Struct. 2008;30(7):1861–8.10.1016/j.engstruct.2007.12.006Search in Google Scholar

[47] Garbin E, Panizza M, Valluzzi M. Experimental assessment of bond behaviour of fibre-reinforced polymers on brick masonry. Struc Eng Int. 2010;20(4):392–9.10.2749/101686610793557582Search in Google Scholar

[48] Kashyap J, Griffith MC, Mohamed Ali MS, Oehlers DJ. Prediction of load-slip behavior of FRP retrofitted masonry. J Compos Constr. 2011;15(6):943–51.10.1061/(ASCE)CC.1943-5614.0000224Search in Google Scholar

[49] Caggiano A, Martinelli E, Faella C. A fully-analytical approach for modelling the response of FRP plates bonded to a brittle substrate. Int J Solids Struct. 2012;49(17):2291–300.10.1016/j.ijsolstr.2012.04.029Search in Google Scholar

[50] Sato Y, Kimura K, Kobatake Y. Bond behavior between CFRP sheet and concrete (Part 1). Architectural Institute of Japan. J Struct Constr Eng. 1997;62:75–82.10.3130/aijs.62.75_5Search in Google Scholar

[51] Gao B, Leung WH, Cheung CM, Kim JK, Leung CKY. Effects of adhesive properties on strengthening of concrete beams with composite strips. Proceedings of the International Conference on FRP Composites in Civil Engineering; 2001 Dec 12–15; Hong Kong, China. Elsevier, 2001. p. 423–32.Search in Google Scholar

[52] Maeda T, Komaki H, Tsubouchi K, Murakami K. Strengthening behavior of carbon fiber sheet using flexible layer. Proc Jpn Concr Inst. 2001;23(1):817–22.Search in Google Scholar

[53] Nanni A, Miller B, De Lorenzis L. Bond of fiber-reinforced polymer laminates to concrete. ACI Mater J. 2001;98(3):256–64.10.14359/10281Search in Google Scholar

[54] Fernando D. Bond Behaviour and debonding failures in CFRP-strengthened steel members [dissertation]. Hong Kong: The Hong Kong Polytechnic University; 2010.Search in Google Scholar

[55] Dehghani E, Daneshjoo F, Aghakouchak AA, Khaji N. A new bond-slip model for adhesive in CFRP–steel composite systems. Eng Struct. 2012;34:447–54.10.1016/j.engstruct.2011.08.037Search in Google Scholar

[56] Yu T, Fernando D, Teng JG, Zhao XL. Experimental study on CFRP-to-steel bonded interfaces. Compos B Eng. 2012;43(5):2279–89.10.1016/j.compositesb.2012.01.024Search in Google Scholar

[57] Fernando D, Yu T, Teng JG. Behavior of CFRP laminates bonded to a steel substrate using a ductile adhesive. J Compos Constr. 2014;18(2):04013040.10.1061/(ASCE)CC.1943-5614.0000439Search in Google Scholar

[58] Chen JF, Yuan H, Teng JG. Debonding failure along a softening FRP-to-concrete interface between two adjacent cracks in concrete members. Eng Struct. 2007;29(2):259–70.10.1016/j.engstruct.2006.04.017Search in Google Scholar

[59] Zhou YW, Wu YF, Yun Y. Analytical modeling of the bond–slip relationship at FRP-concrete interfaces for adhesively-bonded joints. Compos B Eng. 2012;41(6):423–33.10.1016/j.compositesb.2010.06.004Search in Google Scholar

[60] Cornetti P, Carpinteri A. Modelling the FRP-concrete delamination by means of an exponential softening law. Eng Struct. 2011;33(6):1988–2001.10.1016/j.engstruct.2011.02.036Search in Google Scholar

[61] Liu K, Wu YF. Analytical identification of bond–slip relationship of EB-FRP joints. Compos B Eng. 2012;43(4):1955–63.10.1016/j.compositesb.2012.01.048Search in Google Scholar

[62] Yuan H, Lu X, Hui D, Feo L. Studies on FRP-concrete interface with hardening and softening bond-slip law. Compos Struct. 2012;94(12):3781–92.10.1016/j.compstruct.2012.06.009Search in Google Scholar

[63] Biscaia HC, Chastre C, Silva MA. Nonlinear numerical analysis of the debonding failure process of FRP-to-concrete interfaces. Compos B Eng. 2013;50:210–23.10.1016/j.compositesb.2013.02.013Search in Google Scholar

[64] Hart-Smith LJ. Adhesive-bonded double-lap joints. NASA-CR-112236, NASA, California, USA; 1973.Search in Google Scholar

[65] Volkersen O. Die Nietkraftverteilung in zugbeanspruchten Nietverbindungen mit konstanten Laschenquerschnitten. Luftfahrtfor Schung. 1938;15:41–7. (in German)Search in Google Scholar

[66] Goland M, Reissner E. The stresses in cemented joints. J Appl Mech. 1944;11(1):A17–27.10.1115/1.4009336Search in Google Scholar

[67] Rabinovitch O. Impact of thermal loads on interfacial debonding in FRP strengthened beams. Int J Solids Struct. 2010;47(24):3234–44.10.1016/j.ijsolstr.2010.08.003Search in Google Scholar

[68] Wu ZS, Iwashita K, Yagashiro S, Ishikawa T, Hamaguchi Y. Temperature effect on bonding and debonding behavior between FRP sheets and concrete. J Soc. Mater Sci. 2005;54(5):474–80.10.2472/jsms.54.474Search in Google Scholar

[69] Klamer E. The influence of temperature on concrete structures strengthened with externally bonded CFRP. Research Report, Eindhoven University of Technology, Eindhoven, The Netherlands; 2006.Search in Google Scholar

[70] Klamer E. Influence of temperature on concrete beams strengthened in flexure with CFRP [dissertation]. Eindhoven: Eindhoven University of Technology; 2009.Search in Google Scholar

[71] Leone M, Matthys S, Aiello MA. Effect of elevated service temperature on bond between FRP EBR systems and concrete. Compos B Eng. 2009;40(1):85–93.10.1016/j.compositesb.2008.06.004Search in Google Scholar

[72] Al-Shawaf A, Al-Mahaidi R, Zhao XL. Effect of elevated temperature on bond behaviour of high modulus CFRP/steel double-strap joints. Aust J Struct Eng. 2009;10(1):63–74.10.1080/13287982.2009.11465033Search in Google Scholar

[73] Nguyen TC, Bai Y, Zhao XL, Al-Mahaidi R. Mechanical characterization of steel/CFRP double strap joints at elevated temperatures. Compos Struct. 2011;93(6):1604–12.10.1016/j.compstruct.2011.01.010Search in Google Scholar

[74] Kim HS, Shin YS. Flexural behavior of reinforced concrete (RC) beams retrofitted with hybrid fiber reinforced polymers (FRPs) under sustaining loads. Compos Struct. 2011;93(2):802–11.10.1016/j.compstruct.2010.07.013Search in Google Scholar

[75] Cheng L, McComb AM. Unreinforced concrete masonry walls strengthened with CFRP sheets and strips under pendulum impact. J Compos Constr. 2010;14(6):775–83.10.1061/(ASCE)CC.1943-5614.0000131Search in Google Scholar

[76] Gao WY, Teng JG, Dai JG. Effect of temperature variation on the full-range behavior of FRP-to-concrete bonded joints. J Compos Constr. 2012;16(6):671–83.10.1061/(ASCE)CC.1943-5614.0000296Search in Google Scholar

[77] Gao WY, Dai JG, Teng JG. Analysis of Mode II debonding behavior of fiber-reinforced polymer-to-substrate bonded joints subjected to combined thermal and mechanical loading. Eng Fract Mech. 2015;136:241–64.10.1016/j.engfracmech.2015.02.002Search in Google Scholar

[78] De Gennes PG. Scaling concepts in polymer physics. London, UK: Cornell University Press; 1979.Search in Google Scholar

[79] Mahieux CA. A systematic stiffness-temperature model for polymers and applications to the prediction of composite behavior [dissertation]. Blacksburg (VA): Virginia Polytechnic Institute and State University; 1999.Search in Google Scholar

[80] Mahieux CA, Reifsnider KL. Property modeling across transition temperatures in polymers: a robust stiffness–temperature model. Polymer. 2001;42(7):3281–91.10.1016/S0032-3861(00)00614-5Search in Google Scholar

[81] Van Krevelen DW. Properties of polymers. 3rd ed (completely revised). Amsterdam, Oxford, New York: Elsevier Science; 1990.Search in Google Scholar

[82] De Lorenzis L, Zavarise G. Modeling of mixed-mode debonding in the peel test applied to superficial reinforcements. Int J Solids Struct. 2008;45(20):5419–36.10.1016/j.ijsolstr.2008.05.024Search in Google Scholar

[83] Yuan H, Wu L. Full-range behavior of FRP-to-concrete bonded joints. Eng Struct. 2004;26(5):553–65.10.1016/j.engstruct.2003.11.006Search in Google Scholar

[84] Kafkalidis MS, Thouless MD. The effects of geometry and material properties on the fracture of single lap-shear joints. Int J Solids Struct. 2002;39(17):4367–83.10.1016/S0020-7683(02)00344-XSearch in Google Scholar

[85] Zavarise G. Problemi termomeccanici di contatto – aspetti fisici e computazionali [dissertation]. Padova: University of Padova; 1991. (in Italian)Search in Google Scholar

[86] Takeo K, Matsushita H, Makizumi T, Nagashima G. Bond characteristics of CFRP sheets in the CFRP bonding technique. Proc Jpn Concr Inst. 1997;19(2):1599–604.Search in Google Scholar

[87] Ueda T, Sato Y, Asano Y. Experimental study on bond strength of continuous carbon fiber sheet. Spec Publ. 1999;188:407–16.Search in Google Scholar

[88] Wu Z, Yuan H, Yoshizawa H, Kanakubo T. Experimental/analytical study on interfacial fracture energy and fracture propagation along frp-concrete interface. Spec Publ. 2001;201:133–52.Search in Google Scholar

[89] Tan Z. Experimental research for RC beam strengthened with GFRP [dissertation]. Beijing: Tsinghua University; 2002.Search in Google Scholar

[90] Firmo JP, Correia JR, Bisby LA. Fire behaviour of FRP-strengthened reinforced concrete structural elements: A state-of-the-art review. Comp B Eng. 2015;80:198–216.10.1016/j.compositesb.2015.05.045Search in Google Scholar

[91] Colangelo F, Russo P, Cimino F, Cioffi R, Farina I, Fraternali F, et al. Epoxy/glass fibres composites for civil applications: Comparison between thermal and microwave crosslinking routes. Comp B Eng. 2017;126:100–7.10.1016/j.compositesb.2017.06.003Search in Google Scholar

[92] Ghosh P, Bose NR, Mitra BC, Das S. Dynamic mechanical analysis of FRP composites based on different fiber reinforcements and epoxy resin as the matrix material. J Appl Polym Sci. 1997;64(12):2467–72.10.1002/(SICI)1097-4628(19970620)64:12<2467::AID-APP21>3.0.CO;2-XSearch in Google Scholar

Received: 2023-03-19
Accepted: 2023-03-28
Published Online: 2023-05-11

© 2023 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Research Articles
  2. Investigation of differential shrinkage stresses in a revolution shell structure due to the evolving parameters of concrete
  3. Multiphysics analysis for fluid–structure interaction of blood biological flow inside three-dimensional artery
  4. MD-based study on the deformation process of engineered Ni–Al core–shell nanowires: Toward an understanding underlying deformation mechanisms
  5. Experimental measurement and numerical predictions of thickness variation and transverse stresses in a concrete ring
  6. Studying the effect of embedded length strength of concrete and diameter of anchor on shear performance between old and new concrete
  7. Evaluation of static responses for layered composite arches
  8. Nonlocal state-space strain gradient wave propagation of magneto thermo piezoelectric functionally graded nanobeam
  9. Numerical study of the FRP-concrete bond behavior under thermal variations
  10. Parametric study of retrofitted reinforced concrete columns with steel cages and predicting load distribution and compressive stress in columns using machine learning algorithms
  11. Application of soft computing in estimating primary crack spacing of reinforced concrete structures
  12. Identification of crack location in metallic biomaterial cantilever beam subjected to moving load base on central difference approximation
  13. Numerical investigations of two vibrating cylinders in uniform flow using overset mesh
  14. Performance analysis on the structure of the bracket mounting for hybrid converter kit: Finite-element approach
  15. A new finite-element procedure for vibration analysis of FGP sandwich plates resting on Kerr foundation
  16. Strength analysis of marine biaxial warp-knitted glass fabrics as composite laminations for ship material
  17. Analysis of a thick cylindrical FGM pressure vessel with variable parameters using thermoelasticity
  18. Structural function analysis of shear walls in sustainable assembled buildings under finite element model
  19. In-plane nonlinear postbuckling and buckling analysis of Lee’s frame using absolute nodal coordinate formulation
  20. Optimization of structural parameters and numerical simulation of stress field of composite crucible based on the indirect coupling method
  21. Numerical study on crushing damage and energy absorption of multi-cell glass fibre-reinforced composite panel: Application to the crash absorber design of tsunami lifeboat
  22. Stripped and layered fabrication of minimal surface tectonics using parametric algorithms
  23. A methodological approach for detecting multiple faults in wind turbine blades based on vibration signals and machine learning
  24. Influence of the selection of different construction materials on the stress–strain state of the track
  25. A coupled hygro-elastic 3D model for steady-state analysis of functionally graded plates and shells
  26. Comparative study of shell element formulations as NLFE parameters to forecast structural crashworthiness
  27. A size-dependent 3D solution of functionally graded shallow nanoshells
  28. Special Issue: The 2nd Thematic Symposium - Integrity of Mechanical Structure and Material - Part I
  29. Correlation between lamina directions and the mechanical characteristics of laminated bamboo composite for ship structure
  30. Reliability-based assessment of ship hull girder ultimate strength
  31. Finite element method on topology optimization applied to laminate composite of fuselage structure
  32. Dynamic response of high-speed craft bottom panels subjected to slamming loadings
  33. Effect of pitting corrosion position to the strength of ship bottom plate in grounding incident
  34. Antiballistic material, testing, and procedures of curved-layered objects: A systematic review and current milestone
  35. Thin-walled cylindrical shells in engineering designs and critical infrastructures: A systematic review based on the loading response
  36. Laminar Rayleigh–Benard convection in a closed square field with meshless radial basis function method
  37. Determination of cryogenic temperature loads for finite-element model of LNG bunkering ship under LNG release accident
  38. Roundness and slenderness effects on the dynamic characteristics of spar-type floating offshore wind turbine
Downloaded on 20.12.2025 from https://www.degruyterbrill.com/document/doi/10.1515/cls-2022-0193/html
Scroll to top button