Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Case Studies in Construction Materials: Jingjing Zhang, Mengzhao Wang, Kaizhao Han, Jianning Wang

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Case Studies in Construction Materials 21 (2024) e03665

Contents lists available at ScienceDirect

Case Studies in Construction Materials


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

Fatigue performance and damage of partially prestressed concrete


beam with fiber reinforcement
Jingjing Zhang a , Mengzhao Wang a , Kaizhao Han a , Jianning Wang b, *
a
School of Architectural Engineering, North China Institute of Aerospace Engineering, Langfang 065000, China
b
China National Machinery Industry Co., Ltd., Beijing 100080, China

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

Keywords: Ordinary concrete exhibits low strength and a propensity for cracking. Carbon fibers and aramid
Prestressed concrete beam fibers, as materials characterized by high strength, high toughness, and excellent fatigue resis-
Fatigue performance tance, can effectively limit the propagation of micro-cracks and enhance the strength of concrete.
Carbon fiber
To explore the influence of different fiber types on the fatigue performance of prestressed con-
Aramid fiber
crete beams, this study initiates from the perspective of beam components. Employing a com-
bination of experimental research and theoretical analysis, it conducts bending and constant-
amplitude fatigue tests on 12 prestressed concrete beams with various fiber types. Based on the
stiffness of the beam section and residual strain, the damage evolution patterns of the tested
beams were described. The results indicated that the utilization of fiber-reinforced concrete
significantly improves the fatigue life and residual bearing capacity of the beam components.
Specifically, the residual bearing capacities of beams reinforced with carbon fibers, aramid fibers,
and hybrid fibers increased by 58.6 %, 49.5 %, and 64.9 %, respectively. The analysis revealed
that residual strain is a suitable indicator for describing fatigue damage in specimens and rec-
ommends prioritizing the concrete residual strain index. These findings provided a theoretical
basis for the application of fiber-reinforced concrete in practical engineering projects and lifespan
assessment.

1. Introduction

With the economic development, the traffic volume has significantly increased, posing a severe challenge to the load capacity and
durability of highway bridges. Partially prestressed concrete beams are currently the most widely applied component form in highway
bridges [1,2]. However, due to the inherent defects of concrete materials themselves such as high brittleness, easy cracking, low tensile
strength, and poor fracture toughness, which easily lead to the spalling of beam concrete thus exposing the reinforcement and causing
corrosion, seriously affecting the service life of the bridge [3–6]. Whereas fiber, as a material with high strength, high toughness, and
good fatigue resistance [7–11], has become a research hotspot for improving the mechanical properties of concrete [12–17].
In recent years, substantial experimental study into the fatigue performance of fiber-reinforced concrete elements has been
executed. The performance of damaged reinforced concrete (RC) beams strengthened by using a thin precast UHPFRC strip adhesively
bonded to the prepared tensile surface under cyclic loads was investigated by Murthy et al. [18]. Their study only focused on the
damaged beams, not using UHPRFC in new component. Wu et al. [19] study examined the high content hybrid fiber-polymer

* Corresponding author.
E-mail address: wangjianninghebut@163.com (J. Wang).

https://doi.org/10.1016/j.cscm.2024.e03665
Received 3 July 2024; Received in revised form 7 August 2024; Accepted 20 August 2024
Available online 22 August 2024
2214-5095/© 2024 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC license
(http://creativecommons.org/licenses/by-nc/4.0/).
J. Zhang et al. Case Studies in Construction Materials 21 (2024) e03665

concrete’s (HCHFPC) flexural static and fatigue performances. They mainly studied the gaining effect of fibers with different scales on
the crack resistance of concrete. The fatigue life of fiber improved beams is not analyzed. The flexural stress-fatigue life (S-N) of ECC
was predicted by Qiu et al. [20] and compared to the outcomes of experiments. As a kind of engineering cementing composite material,
the results of ECC can not be applied to other kinds of fiber concrete. Cui et al. [21] investigated the fatigue life of polypropylene fiber
reinforced concrete (PFRC) subjected to varied degrees of repeated compressive stress using eighteen groups of prismatic specimens.
The research is mainly carried out on test blocks rather than components. The flexural fatigue behavior of an ultra-high-performance
fiber-reinforced concrete was evaluated Rios et al. [22]. A number of concrete mixtures were tested at room temperature, 100, 200,
and 300 degrees Celsius. Two kinds of reinforcement were investigated: steel fibers and steel and polypropylene fibers combined. This
paper studied the influence of temperature on fatigue properties of different fiber reinforced concrete. According to research by
Stephen et al. [23], the fatigue life of fractured concrete is only increased by greater fiber dosages at high cyclic loading magnitudes; at
smaller magnitudes, the effect of dosage is negligible. The endurance limit for fiber concretes is thought to be roughly 50 % of the
pre-crack load during the post-peak period. But the pre-cracked notched beams were tested, the results may not suitable for other FRC
beams. Chen et al. [24] used a fatigue test to investigate the manufacture of recycled tyre polymer fibre (RTPF) concrete. The flexural
strength of concrete was enhanced by 3.6–9.6 %. The study mainly compared RTPF and polypropylene fibre (PPF) reinforced concrete.
In order to explain the fatigue deformation behavior of plain and fiber-reinforced concrete in three stages, Huang et al. [25] developed
a unique model based on the three-parameter Weibull function. This study discussed the mechanical properties of fiber reinforced
concrete materials rather than specimens. By contrasting high-performance synthetic fiber pavement concrete (HPSFPC) with ordinary
pavement concrete (PPC), Long et al. [26]. examined the corrosion resistance and fatigue performance of HPSFPC. The model suited
the data quite well and effectively disclosed the mechanical characteristics of pavement concrete under the combined effects of fatigue,
corrosion, and fiber content. The influence on fatigue performance of fiber reinforced concrete beams is not studied.
In addition, some scholars have conducted research on the fatigue performance of fiber-reinforced concrete beams, specifically
concentrating on steel fibers [27–34]. Through experimental and computational research, Gao et al. [35] looked into how fiber
clustering affected the fatigue behavior of steel-fiber-reinforced concrete (SFRC) beams with reinforcements. They discovered that as
the fiber volume fraction increased from 0 to 1.0 vol%, the fatigue life of the beam increased. However, when the fiber volume fraction
reached 1.5 vol%, the fatigue performance of the beam decreased because there were a significant number of fiber clusters in the beam.
It was not mentioned whether this applies to other flexible fibers. The steel fiber reinforced concrete beams underwent both static and
fatigue testing, carried out by Adel et al. [36]. Throughout the fatigue life, it was demonstrated that the crack-bridging strength would
progressively deteriorate at various flexural fatigue stress levels. Furthermore, when the flexural fatigue stress level is modest, the
residual flexural capacity at the end of the fatigue life differs little from the initial capacity determined under static loading. Gao et al.
[37] used flexural testing to study the fatigue performance of steel-fiber-reinforced concrete (SFRC) beams. An SFRC beam with 1 %
fiber content had a twice as long fatigue life as non-fiber-reinforced concrete beams. They devised a methodology for predicting the
stress range of tensile steel bars in SFRC beams under fatigue loads. The impact of fiber content on the compressive fatigue of
steel-fiber-reinforced self-compacting concrete was assessed by Poveda et al. [38] Fibers have a positive impact on fatigue resistance.
However, an intermediate fiber content (0.6 %) yields the longest fatigue life rather than the maximum one. Dogbone specimens with
different steel fiber volume contents (0, 0.75, and 1.5 %), were tested by Isojeh et al. [39] to examine the behavior under direct tension
fatigue loading. The fatigue life of steel fiber concrete was shown to rise from 0 % to 1.5 % of the total weight under the same loading
conditions. The effectiveness of employing steel fibers to enhance the fatigue performance and load carrying capacity of rail sleepers
was examined by Parvez et al. [40] Compared to sleepers without fibers, those with 0.5 % fiber showed a greater static capacity, a
longer fatigue life, fewer deflections, and finer crack widths. All of the above are for steel fibers with high stiffness, and whether the
research findings may be extended to other flexible fibers has not been determined. Furthermore, the bonding properties of steel fiber
and concrete differ from those of flexible fiber, necessitating research into the mechanical properties of carbon fiber and aramid fiber
concrete beams.
In addition, there is relatively little research on the fatigue performance of fiber-reinforced prestressed concrete beams, especially
regarding carbon fibers and aramid fibers. By the way, a single type of fiber can only enhance one aspect performance of concrete, and
the gaining effect will not be significantly improved with the increase of volume content [41,42]. The mix design for selecting the best
hybrid fiber has been done in previous studies [43].
In this study, the static flexural strength and fatigue durability of PPCB with fiber reinforcement are studied through flexural, and
fatigue tests. For this purpose, the study has designed 12 experimental beams of fiber prestressed reinforced concrete, subjected to
bending and constant amplitude fatigue tests. The goal is to assess how fibers affect the fatigue performance of prestressed concrete
beams and analyze their damage process, providing valuable references for real-world engineering applications.

2. Material

The characteristics of the fibers, concrete and steel reinforcement materials used in the experiment are listed in Table 1, Table 2and

Table 1
Mechanical properties of fiber materials.
Fiber type Density (kg/m3) Cut length (mm) Diameter (μm) Tensile strength (MPa) Tensile modulus (GPa) Elongation (%)

Carbon fiber 1750 12 8 3530 240 1.5


Aramid fiber 1440 12 12 3150 80 3.6

2
J. Zhang et al. Case Studies in Construction Materials 21 (2024) e03665

Table 3. This experiment was conducted on the basis of previous studies, and the test results showed that the optimal fiber addition rate
was 0.6 % [44]. The dimensions and steel reinforcement layout of the test beams are illustrated in Fig. 1. The tensile stress is 0.73 times
the standard value of tensile strength, which is 1150 MPa. Plain concrete in the beam FL adopts the conventional C40 concrete mix
proportion.

3. Test procedure

3.1. Static test

Refer to GB/T 50152–2012 [45] for the flexural strength test. Electro-hydraulic servo testing machine is used for four-point loading
and the maximum load range is 200 kN, as shown in Fig. 2(a). There are four beams in the static test, L represents the ordinary concrete
beam as a contrast, CL represents the carbon fiber reinforced concrete beam, AL represents the aramid fiber reinforced concrete beam,
and CAL represents the hybrid fiber reinforced concrete beam. The loading test uses a hierarchical loading method. 3 kN is put into
each stage prior to the test beam cracking, and data collection and crack observation begin ten minutes after each stage is loaded. After
the test beam cracks, the loading for each stage is increased to 3 kN. Finally, when the load exceeds 80 % of the calculated bending
bearing capacity, the displacement control is employed to determine the accurate ultimate bearing capacity, maximum deflection, and
maximum crack width. These steps are repeated until the load equals 80 % of the estimated cracking load. Twelve 8.0-cm-long strain
gauges are affixed to the specimen at the middle and third points in order to record concrete strain data during the flexural test.

3.2. Fatigue test

To completely investigate the fatigue properties of fiber reinforced concrete, a total of eight test beams are put up for fatigue testing.
The flexural fatigue test’s maximum stress amplitude (σmax) is determined by the specimens’ ultimate flexural strength (Mu). The
maximal stress level has two grades: 0.70 and 0.80. The minimum stress level is 0.20 Mu. As a result, in order to prevent test mistakes, a
sine wave with a constant amplitude is used for cyclic loading and 3 Hz for fatigue loading. During the fatigue test, the dynamic strain
acquisition equipment records the strain data; the acquisition frequency is fixed at 10 Hz. The test apparatus is shown in Fig. 2(b). The
test loading system is shown in Fig. 3.
Prior to commencing the fatigue tests, a preliminary loading of 2 kN is applied to validate the connectivity between the data
acquisition devices and instruments and to ensure the data collection system is operational. Upon successful verification, the load is
removed. The test beams are then subjected to static force application. This application process is meticulously force-controlled and
incrementally divided into five stages, reaching up to the fatigue load’s upper threshold. Following the application of each load stage, a
duration of approximately 5 minutes is observed to facilitate the creep process in the concrete, steel reinforcement, and reinforcement
materials. During this period, measurements of strain, deflection, and the observation of crack formation are conducted, documenting
the progression and dimensions of each crack under respective loads.
Subsequently, the beams are methodically unloaded in five stages to a neutral state. After each unloading stage, a holding period of
about 5 minutes is observed, mirroring the loading procedure to record data on strain, displacement, and crack dimensions. At
specified fatigue cycles of 0.01 million, 0.05 million, 0.1 million, 0.3 million, 0.5 million, 1 million, 1.5 million; 1.8 million; and 2
million, the tests are momentarily halted, and the load is unloaded to zero. The static loading sequence is repeated for data acquisition.
Should the beam exhibit fatigue failure before reaching 2 million cycles, the procedure is halted to observe the failure characteristics
and to measure the maximum deflection. Conversely, if no fatigue failure is observed post 2 million cycles. Ordinary concrete test
beams are numbered FL, FL-1; fiber concrete test beams are divided into carbon fiber beams FCL, FCL-1, aramid fiber beams FAL, FAL-
1, and hybrid fiber beams FCAL, FCAL-1. Based on the calculation of the bending bearing capacity of the fiber concrete beams, the
numerical values of the fatigue load effect are obtained as shown in Table 4.

4. Results and analysis

4.1. Static strength and fatigue life

Three-point bending tests were conducted on four fiber-reinforced concrete beams, and the flexural capacities obtained are pre-
sented in Table 5. Under the effect of fatigue load, all beam specimens underwent fatigue failure. In the early stage of the fatigue test,
the number of cracks gradually increased, but the width development was not rapid; in the middle stage of fatigue, the cracks basically
appeared evenly, only the width kept increasing, and the emergence of new cracks led to stress redistribution; approaching the beam
failure stage, the main crack developed rapidly, and the deflection of the test beam accelerated. Ordinary beams suffered fatigue failure
due to the fracture of ordinary reinforcing bars in the tension zone, and the failure of fiber concrete beams was mainly because the

Table 2
Mechanical properties of C40 concrete.
Test specimen 1 2 3 Average

Strength (MPa) 47.3 47.6 46.8 47.2

3
J. Zhang et al. Case Studies in Construction Materials 21 (2024) e03665

Table 3
Mechanical properties of rebar materials.
Rebar Diameter (mm) Yield strength (MPa) Ultimate strength (MPa) Elongation (%)

Regular rebar 6 312 407 25


Prestressing tendon 7 1410 1603 4.2

Fig. 1. Dimensions and reinforcement of PPCB.

Fig. 2. Test apparatus: (a) Static test; (b) Fatigue test.

Fig. 3. Equal amplitude fatigue test loading system.

Table 4
Fatigue load of fiber-reinforced concrete beams.
Beam σmax (kN) σmin (kN)
FL 18.0 5.0
FL− 1 20.5 5.0
FCL 23.0 6.5
FCL− 1 26.0 6.5
FAL 20.5 6.0
FAL− 1 23.5 6.0
FCAL 21.0 6.0
FCAL− 1 24.0 6.0

Note: 1. C represents carbon fiber, A represents aramid fiber, and CA represents a


combination of both fibers. 2. σ max is the maximum fatigue load, σ min is the min-
imum fatigue load.

crack width reached the upper limit of 1.5 mm, accompanied by the crushing failure of the upper concrete.
Table 6 summarizes the data from the fatigue test results of the test beams. Fig. 4 shows the fatigue failure diagram of the beams
under the fatigue upper limit of 0.8 Mu action. It can be seen that failures happened close to the loading point, between grids 39 and 47
during the test. The fatigue life of ordinary beams dropped sharply to 51,000 times, while the lifespan of the other test beams was
around 500,000 times.

4
J. Zhang et al. Case Studies in Construction Materials 21 (2024) e03665

Table 5
Flexural test results.
Beam Cracking load Fcr (kN) Cracking moment Mcr (kN⋅m) Ultimate load Fu (kN) Ultimate moment Mu (kN⋅m)

L 12.0 7.4 38.8 23.9


CL 17.0 10.5 49.0 30.2
AL 16.1 9.9 43.5 26.8
CAL 16.5 10.2 44.9 27.7

Table 6
Fatigue test results.
Loading peak Beam type Fatigue life (×104)

0.7 Mu FL 157
FCL >200
FAL >200
FCAL >200
0.8 Mu FL− 1 5.1
FCL− 1 43.7
FAL− 1 52.4
FCAL− 1 56.6

Fig. 4. Fatigue failure diagram of test beams: (a) FL-1; (b) FCL-1; (c) FAL-1; (d) FCAL-1.

4.2. Analysis of beam deflection under fatigue loading

4.2.1. Maximum deflection


Extract the deflection values of the test beams under the maximum fatigue load and plot them as curves, as shown in Fig. 5. It can be
seen that the development of the deflection of the test beams is roughly divided into three stages. In the first stage, occurring during the
initial loading phase (before 10,000 cycles), the rate of deflection growth is relatively fast. The second stage spans from 10,000 to
100,000 load cycles, representing the normal usage state of the test beams. During this stage, the deflection changes are relatively
stable. However, the slope of the deflection curve of the ordinary beam FL-1 is larger than that of the other test beams, and the rate of
change accelerates again in the last tens of thousands of cycles. In the final stage, the stiffness of the ordinary beams drops rapidly,
eventually leading to fatigue failure at 51,000 cycles. The deflection values of beams with internal fiber are generally lower than those
of ordinary concrete beams, and the development trend is more stable, that is, the degradation of stiffness is slower, reflecting the good
fatigue resistance of fibers. Within the number of fatigue cycles, the deflections of beams FCL-1, FAL-1, and FCAL-1 increased by

5
J. Zhang et al. Case Studies in Construction Materials 21 (2024) e03665

Fig. 5. The maximum deflection curve with loading times: (a) σmax=0.7 Mu; (b) σmax=0.8 Mu.

3.9 mm, 4.6 mm, and 5.0 mm, respectively. Because aramid fiber has good fatigue and impact resistance, the energy absorbed during
the fatigue test increased, slowing down the internal damage of the test beams. Overall, the performance of hybrid fiber beams is the
best.

4.2.2. Residual deflection


From Fig. 6, it can be seen that the residual deflections of the fiber-reinforced concrete beams are less than those of ordinary beams,
and the slope of the curve development is generally less than that of ordinary beams. After 100,000 cycles of fatigue load, the residual
deflection of the test beams is close to the value of ordinary beams after 10,000 cycles, which well reflects the fatigue resistance of
fibers. To form a more intuitive comparative analysis, the deflection change coefficient is used to describe the fatigue damage
development process.
f = Δf/fmax (1)

where Δf is the residual deflection deformation of the fatigue test beam; fmax is the deflection of the test beam when the static load is
applied to the fatigue limit after each level of fatigue cycles. The change curves of the residual deformation coefficients of each test
beam under the fatigue load and the number of fatigue load cycles are plotted and fitted, with the fitting results shown in Fig. 7.

4.3. Material strain analysis under fatigue load

4.3.1. Concrete compressive strain


Five concrete strain gauges were arranged on the side of the test beam, as shown in Fig. 8. The maximum compressive strain of
concrete was used for comparative analysis in this test. Fig. 9 and Fig. 10 show the curve of concrete compressive strain versus the
number of fatigue cycles for FRC beams. It can be seen that when the load cycles to 100,000 times, the concrete compressive strains of
FCL-1, FAL-1, and FCAL-1 beams with added fibers decreased by 7.3 %, 12.4 %, and 8.2 %, respectively. This is because the addition of
fibers enhanced the interfacial bonding strength between aggregates in concrete, improving the bending fatigue performance of
concrete. Under the same load conditions, the concrete strains in the compression zone of FCL-1, FAL-1, and FCAL-1 beams are smaller
than those of FL-1 beams, and the strain of FCAL-1 beams is the smallest. This indicates that hybrid fiber-reinforced concrete over-
comes the disadvantages of single-fiber concrete, fully utilizes the advantages of both, effectively reduces or delays the generation of
cracks, and reduces the accumulation of damage.

Fig. 6. Residual deflection of fiber-reinforced concrete beams: (a) σmax=0.7 Mu; (b) σmax=0.8 Mu.

6
J. Zhang et al. Case Studies in Construction Materials 21 (2024) e03665

Fig. 7. Residual deformation coefficient: (a) FL-1; (b) FAL-1; (c) FCL-1; (d) FCAL-1.

Fig. 8. The location of strain gauges.

Fig. 9. Concrete strain variation curve with fatigue times: (a) σmax=0.7 Mu; (b) σmax=0.8 Mu.

4.3.2. Ordinary reinforcing steel strain analysis


The research shows that the fatigue of prestressed concrete beams is mainly caused by the fatigue fracture of non-prestressed
reinforcement bars in the tension area, which leads to the final failure of the test beams [46]. Therefore, the analysis of the strain
of ordinary reinforcing bars in the test beams is an important aspect of the study of bridge fatigue performance. After experiencing
different cycle numbers, the load-ordinary reinforcing steel strain curves of each test beam are shown in Fig. 11.
For the conventional beam FL-1, the strain growth rate of the reinforcing bar under the fatigue cycle was the highest. For beams
FCL-1, FAL-1, and FCAL-1, under the effect of fatigue load, the strain of the reinforcing bars grew uniformly, and the overall growth
magnitude was less than that of beam FL-1. Comparing different types of fiber-reinforced concrete (FRC) beams, based on the static

7
J. Zhang et al. Case Studies in Construction Materials 21 (2024) e03665

Fig. 10. Residual strain variation curve with fatigue times: (a) σmax=0.7 Mu; (b) σmax=0.8 Mu.

Fig. 11. Strain of ordinary reinforcing steel under fatigue Loading: (a) σmax=0.7 Mu; (b) σmax=0.8 Mu.

loading data after different fatigue cycle numbers, the strain changes of reinforcing bars in FRC beams were more stable than those in
conventional beams, tending towards linear growth. After 100,000 fatigue cycles, the strain values of ordinary reinforcing bars from
high to low were: FAL-1, FCL-1, FCAL-1, and the development pattern of residual strain was the opposite; at 100,000 loads, the
maximum strain growth magnitudes of beams FCL-1, FAL-1, and FCAL-1 were reduced by 36 %, 49 %, and 55 % compared to con-
ventional beams, indicating that the presence of fibers could better complete the transfer of load and absorption of energy, reducing the
stress borne by the reinforcing bars; and under the same load effect, after 100,000 cycles of fatigue, the strain increment of reinforcing
bars was 35 %, 27 %, and 32 % higher than that of static loading, fully reflecting the good fatigue resistance of aramid fibers; it was
found that the reinforcing bar strain curve of beam FCAL-1 was the tightest, close to linear growth, fully reflecting the positive effect of
hybrid fibers.

4.3.3. Prestressed reinforcing bar strain


The change in strain of prestressed reinforcing bars during the fatigue process can reflect the stress distribution of the test beam.
Fig. 12 shows the variation curve of prestressed reinforcing bar strain with the number of fatigue cycles. From the curve, it is observed
that during the fatigue test, the prestressed reinforcing bars did not yield, and the change in reinforcing bar strain was close to linear,
with relatively small residual strain. The curve for conventional beam FL-1 showed a diverging pattern, whereas the strain curves for
FRC beams were tighter, reflecting that under the effect of fibers, the strain of prestressed reinforcing bars was less affected by fatigue.

Fig. 12. Strain of prestressed steel bars under fatigue loading: (a) σmax=0.7 Mu, (b) σmax=0.8 Mu.

8
J. Zhang et al. Case Studies in Construction Materials 21 (2024) e03665

5. Flexural fatigue damage analysis

5.1. Fatigue damage analysis based on residual strain

According to existing damage models [47,48], the damage evolution equation is shown below for isotropic media.
∂Ψ D Y ṗ + π̇
Ḋ = − = − × (2)
∂Y S0 (1 − D)α0

where, ψ D is the dissipation potential after Legendre transformation, the damage energy release rate (Y), cumulative plastic strain rate
(ṗ), and micro-plastic strain rate (π̇) are represented by these numbers, S0 and α0 are material characteristic parameters dependent on
temperature.
When α0=0, the damage evolution rate of fatigue damage is as follows. In most circumstances, the damage rate is linearly pro-
portional to the cumulative plastic strain rate.
∂Ψ D Y
Ḋ = − = − × ṗ (3)
∂Y S0

Furthermore, the free energy density function is independent of Ḋ in the condition of one dimension, thus the damage energy
release rate can be obtained as following, and substituting into the above equation yields.

∂φ σ2
Y= − ρ = (4)
∂Ḋ 2E

σ2
Ḋ = − ε̇P (5)
2ES0
Taking the differential form of the above equation and eliminating dt (that is, the time differential), and integrating the equation
yields.
∫D ∫ εp ( 2 )
σ
dD = dεp (6)
0 εp0 − 2ES0

Further, it can be obtained as following:


( )
σ2
D= − (εP − εP0 ) + C (7)
2ES0
Considering boundary conditions, when εp < εp0 , D=0, when εp = εp0 , D=1. The result is obtained by integrating.
εP − εP0
D= (8)
εPC − εP0

where, εP stands for the material’s residual plastic strain. The cumulative residual plastic strain at the beginning of material damage is
represented by εP0 , while the cumulative residual plastic strain at fatigue failure is represented by εpc .
The damage equation between the cumulative residual plastic strain εP and the damage variable D is given above. It was established
by theoretical deduction. Obviously, by calculating the fatigue damage of concrete, reinforcing bars, and prestressed bars using the
measured residual strain data, the material fatigue damage evolution curve can be obtained, as shown in Fig. 13. From the above
figures, it can be seen that the development of material fatigue damage shows stage characteristics. Between 15 % and 40 % of the
concrete is damaged in the first and second phases of strain, and 15 % is damaged in the third stage. At the moment of fatigue failure,
the total damage was approximately 60 %. The damage increase rate is nearly constant, and there is no discernible difference between
the first and second stages of cumulative fatigue damage in steel bars and prestressed bars. Approximately 40 % of the damage occurs
during the entire fatigue phase, which is not very substantial given the increase in load cycles. It is more reasonable to use the residual
strain index of concrete to represent the fatigue damage index of beams.

5.2. Damage analysis based on component section stiffness

The stiffness of a component refers to the ability to resist deformation under load, that is, the stress required to cause unit
deformation. Stiffness degradation is the term for the phenomena wherein, under cyclic repeating load, the displacement at the peak
point rises with the number of cycles while maintaining the same peak load. Since the fatigue test uses a simply supported beam with
concentrated force loading, the moment distribution along the length of the beam is triangular, that is, the moment is maximum at mid-
span, and the stiffness is minimum. Therefore, considering the reliability of the structure and referring to the "minimum stiffness
principle", the stiffness corresponding to the mid-span position is used. The flexural stiffness B of the test beam is calculated using the
equation of the deflection curve in structural mechanics.

9
J. Zhang et al. Case Studies in Construction Materials 21 (2024) e03665

Fig. 13. Evolution law of damage of beams under fatigue loading: (a) Concrete; (b) Ordinary steel bars; (c) Prestressed steel bars.

Ml2
B=α (9)
f

where, B is the flexural stiffness of the test beam. M is the maximum moment under fatigue load. l is the calculated span of the test
beam. f is the maximum deflection under fatigue load. α is a coefficient related to the load form and support of the test beam, obtained
by the graph multiplication method as 0.0833.
Fig. 14 shows the stiffness degradation curve obtained from the fatigue test results. It can be seen that all test beams exhibit a three-

10
J. Zhang et al. Case Studies in Construction Materials 21 (2024) e03665

stage development pattern: the first stage is the rapid decay stage, accounting for approximately 10 % of the entire cycle. During this
stage, the stiffness degradation is rapid, indicating a significant impact of fatigue loading on the beam stiffness in the initial cycles. The
second stage is the stable development stage of stiffness decay, which occupies about 80 % of the entire fatigue process. The
component reaches a relatively stable state overall, and stiffness decay shows a linear development relationship. The third stage is the
rapid failure stage, representing approximately 10 % of the entire fatigue process. Stiffness decreases rapidly, ultimately leading to
destruction.
The stiffness of the test beams improved with fiber is higher than that of the prototype beams, and the stiffness degradation rate
with the number of fatigue load cycles, that is, the stiffness degradation rate, is reduced, reflecting the improved fatigue resistance of
the test beams by fibers. Among them, aramid fiber plays the most significant role in slowing down component stiffness degradation,
followed by hybrid fiber.
From the experimental phenomenon, it is known that the fatigue failure of the test beam FL is due to the fatigue fracture of the steel
bar, which may have no signs before fatigue failure, and the stiffness of the test beam may suddenly drop to 0 at the time of failure.
Therefore, it is necessary to study the stiffness degradation of the test beam at fatigue failure. The stiffness selected for the stiffness
degradation curve is the data of the test beam in normal working condition before destruction. For test beams that have not undergone
fatigue failure, the stiffness after 2 million cycles of fatigue is regarded as the ultimate stiffness.
To study the stiffness degradation pattern of the test beams under fatigue load, the stiffness of the test beams needs to be processed
to make the results more intuitive. Here, the stiffness decay coefficient K is defined as following:
Bn
K = 1− (10)
B0

where, Bn is the stiffness of the test beam under n cycles of load. B0 is the initial stiffness.
The value of the decay coefficient is specific to a particular number of fatigue cycles, K=0 means the test beam is in a non-decay
state, and the stiffness decay coefficient K is a monotonically increasing and irreversible function. Table 7 lists the related results of the
test beam stiffness analysis under 0.7 Mu fatigue load. The destruction flexural stiffness ratio of the prototype beam is 0.48; the average
destruction stiffness ratio of FRC beams is 0.71. The results show that when the decay coefficient of the prototype beam reaches 0.5, the
test beam is about to fail; for FRC beams, when the stiffness decays to 0.7 times the initial stiffness, the test beam can be considered
about to fail. According to article [49], the stiffness decay coefficient is linearly related to the fatigue life lg N, that is, from the final
stiffness value changes, the degree of component stiffness degradation can be measured. It can be seen from the data that the stiffness
decay rate of beam FAL is slow, which reflects the good fatigue resistance of aramid fiber.

6. Conclusion

To explore the impact of different fibers on the fatigue resistance of prestressed concrete beams, the article conducted amplitude
fatigue tests on four test beams, analyzed parameters such as deflection, concrete strain, and steel bar strain under fatigue load, and
discussed the cumulative fatigue damage performance based on stiffness degradation patterns, equivalent dynamic modulus, and
material residual strain. The conclusions were drawn as follows.
(1) The test beams’ fatigue life and residual bearing capacity were greatly increased by the inclusion of fibers, with aramid, carbon,
and hybrid fiber beams showing increases in residual bearing capabilities of 58.6 %, 49.5 %, and 64.9 %, respectively.
(2) In the early phases of fatigue loading, the test beam’s midspan deflection increases quickly. The concrete’s compressive zone in
the test beam gradually gets lower as the number of fatigue cycles increases. The rate of upward movement of the neutral axis in the
fiber-reinforced beam is significantly lower than that in the ordinary beam under fatigue loading.
(3) Based on residual strain data, it was analyzed that when the test beams experienced fatigue failure, the damage variable was
about 60 %. It was found that residual strain is a better indicator for describing the fatigue damage of specimens, and the residual strain
index of concrete is recommended as a priority.

Fig. 14. The stiffness degradation curve of beams: (a) σ max=0.7 Mu; (b) σmax=0.8 Mu.

11
J. Zhang et al. Case Studies in Construction Materials 21 (2024) e03665

Table 7
Results of test beam stiffness.
Beam Initial stiffness B0 (×1012 N⋅m2) Ultimate stiffness Bn (×1012 N⋅m2) Stiffness ratio Bn/B0 Stiffness decay coefficient K

FL 0.64 0.31 0.48 0.52


FCL 0.93 0.62 0.67 0.33
FAL 0.72 0.56 0.76 0.24
FCAL 0.85 0.59 0.69 0.29

CRediT authorship contribution statement

Mengzhao Wang: Investigation, Data curation. Jingjing Zhang: Writing – original draft, Validation, Methodology, Investigation.
Jianning Wang: Writing – original draft, Resources, Funding acquisition, Conceptualization. Kaizhao Han: Writing – original draft,
Visualization, Data curation.

Declaration of Competing Interest

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

Data Availability

The authors do not have permission to share data.

Acknowledgments

The authors wish to acknowledge the research funding provided by Construction of Scientific and Technological Research Project
by Hebei Department of Housing and Urban-Rural Development (2023-2156), Scientific Research Fund of Institute of Engineering
Mechanics, China Earthquake Administration (Grant No. 2023D03), SINOMACH Youth Science and Technology Fund (QNJJ-PY-2022-
02), Young Elite Scientists Sponsorship Program (BYESS2023432). All statements, results, and conclusions are those of the researchers
and do not necessarily reflect the views of these foundations.

References

[1] X.Y. Xiong, N. Hua, P.C. Yu, Static load test study on partially prestressed retard-bonded concrete beams, Build. Struct. 48 (8) (2018) 34–40.
[2] Z.F. Zhang, Study on construction technology optimization and improvement of prestressed concrete continuous beam bridge, Constr. Technol. 15 (2023)
209–211.
[3] B. Han, L. Zhang, C. Zhang, et al., Reinforcement effect and mechanism of carbon fibers to mechanical and electrically conductive properties of cement-based
materials, Constr. Build. Mater. 125 (2016) 479–486.
[4] H. Huang, Y.J. Yuan, W. Zhang, et al., Property assessment of high-performance concrete containing three types of fibers, Int. J. Concr. Struct. Mater. 15 (2021)
39.
[5] M. Zawam, K. Soudki, J.S. West, Factors affecting the time-dependent behaviour of GFRP prestressed concrete beams, J. Build. Eng. 24 (2019) 100715.
[6] S.T. Cheng, H.X. He, Y. Cheng, et al., Flexural behavior of RC beams enhanced with carbon textile and fiber reinforced concrete, J. Build. Eng. 70 (2023)
106454.
[7] D.Y. Gao, M. Zhang, H.T. Zhu, Fatigue test and stiffness calculation of steel fiber reinforced high-strength concrete beams with reinforcements, J. Build. Struct.
34 (8) (2013) 142–149.
[8] Z.B. Wang, P.Y. Wang, F.Q. Zhu, Synergy effect of hybrid steel-polyvinyl alcohol fibers in engineered cementitious composites: Fiber distribution and
mechanical performance, J. Build. Eng. 62 (2022) 105348.
[9] V.M.D. Monteiro, D.C.T. Cardoso, F.D. Silva, Mechanisms of fiber-matrix interface degradation under fatigue loading in steel FRC, J. Build. Eng. 68 (2023)
106168.
[10] D.Y. Gao, M. Zhang, J. Zhao, Calculating method for crack width of steel fiber reinforced high-strength concrete beams under fatigue loads, China Civ. Eng. J. 46
(3) (2013) 40–48.
[11] W. Wu, X. He, M. Alam, et al., Comparative research on low cycle fatigue performance of the GFRP (Steel)-RC beams, Case Stud. Constr. Mater. 19 (2023)
e02317.
[12] A. Razmi, M. Mirsayar, On the mixed mode I/II fracture properties of jute fiber-reinforced concrete, Constr. Build. Mater. 148 (2017) 512–520.
[13] I.M. Al-Damad, S.J. Foster, Behavior of postcracked steel fiber-reinforced concrete in fatigue and development of a damage prediction model, Struct. Concr. 23
(2) (2022) 1593–1610.
[14] M. Mohamadi, J.A. Mohandesi, M. Homayonifar, Fatigue behavior of polypropylene fiber reinforced concrete under constant and variable amplitude loading,
J. Compos. Mater. 47 (26) (2013) 3331–3342.
[15] J. Huang, D. Rodrigue, Stiffness behavior of sisal fiber reinforced foam concrete under flexural loading, J. Nat. Fibers 19 (15) (2022) 12251–12267.
[16] H. Rooholamini, A. Karimi, J. Safari, A statistical fatigue life prediction model applicable for fibre-reinforced roller-compacted concrete pavement, Road. Mater.
Pavement Des. 24 (1) (2023) 103–120.
[17] K. Cui, L. Xu, L. Li, et al., Mechanical performance of steel-polypropylene hybrid fiber reinforced concrete subject to uniaxial constant-amplitude cyclic
compression: Fatigue behavior and unified fatigue equation, Compos. Struct. 311 (2023) 116795.
[18] A.R. Murthy, B.L. Karihaloo, P.V. Rani, et al., Fatigue behaviour of damaged RC beams strengthened with ultra high performance fibre reinforced concrete, Int.
J. Fatigue 116 (2018) 659–668.
[19] W. Wu, X. He, Z. Yi, et al., Flexural fatigue behaviors of high-content hybrid fiber-polymer concrete, Constr. Build. Mater. 349 (2022) 128772.
[20] J. Qiu, E.H. Yang, Micromechanics-based investigation of fatigue deterioration of engineered cementitious composite (ECC), Cem. Concr. Compos. 95 (2017)
65–74.

12
J. Zhang et al. Case Studies in Construction Materials 21 (2024) e03665

[21] K. Cui, L. Xu, X. Li, et al., Fatigue life analysis of polypropylene fiber reinforced concrete under axial constant-amplitude cyclic compression, J. Clean. Prod. 319
(2021) 128610.
[22] J.D. Rios, H. Cifuentes, S. Blason, et al., Flexural fatigue behaviour of a heated ultra-high-performance fibre-reinforced concrete, Constr. Build. Mater. 276
(2021) 122209.
[23] S.J. Stephen, R. Gettu, Fatigue fracture of fibre reinforced concrete in flexure, Mater. Struct. 53 (3) (2020) 56.
[24] M. Chen, H. Zhong, M.Z. Zhang, Flexural fatigue behaviour of recycled tyre polymer fibre reinforced concrete, Cem. Concr. Compos. 105 (2020) 103441.
[25] B.T. Huang, Q.H. Li, S.L. Xu, Fatigue deformation model of plain and fiber-reinforced concrete based on Weibull function, J. Struct. Eng. 145 (1) (2019)
04018234.
[26] X.Y. Long, L.C. Cai, W.Z. Li, RSM-based assessment of pavement concrete mechanical properties under joint action of corrosion, fatigue, and fiber content,
Constr. Build. Mater. 197 (2019) 406–420.
[27] A. Medeiros, X.X. Zhang, G. Ruiz, et al., Effect of the loading frequency on the compressive fatigue behavior of plain and fiber reinforced concrete, Int. J. Fatigue
70 (2015) 342–350.
[28] S. Goel, S.P. Singh, Fatigue performance of plain and steel fibre reinforced self compacting concrete using S-N relationship, Eng. Struct. 74 (2014) 65–73.
[29] D.Y. Gao, C.S. Huang, Fatigue performance test and life calculation method of fiber reinforced asphalt concrete, J. Build. Mater. 18 (3) (2015) 505–512.
[30] S.J. Choi, J.S. Mun, K.H. Yang, et al., Compressive fatigue performance of fiber-reinforced lightweight concrete with high-volume supplementary cementitious
materials, Cem. Concr. Compos. 73 (2016) 89–97.
[31] F. Germano, G. Tiberti, G. Plizzari, Post-peak fatigue performance of steel fiber reinforced concrete under flexure, Mater. Struct. 49 (10) (2016) 4229–4245.
[32] A. Parvez, S.J. Foster, Fatigue behavior of steel-fiber-reinforced concrete beams, J. Struct. Eng. 141 (4) (2015).
[33] B.D.G. Sepulveda, P. Visintin, D.J. Oehlers, Fatigue bond-slip properties of steel reinforcing bars embedded in UHPFRC: Extraction and development of an
accumulated damage law, Case Stud. Constr. Mater. 17 (2022) e01370.
[34] J. Wang, W.Y. Ji, M.Z. An, et al., Fatigue behavior of reinforced UHPC-NC composite beams, Case Stud. Constr. Mater. 18 (2023) e02055.
[35] D.Y. Gao, Z.Q. Gu, C.J. Wei, et al., Effects of fiber clustering on fatigue behavior of steel fiber reinforced concrete beams, Constr. Build. Mater. 301 (2021)
124070.
[36] M. Adel, P. Jiradilok, K. Matsumoto, et al., Evaluation of crack-bridging strength degradation in SFRC structural beams under flexural fatigue, Compos. Struct.
244 (2020) 112267.
[37] D.Y. Gao, Z.Q. Gu, J.Y. Tang, et al., Fatigue performance and stress range modeling of SFRC beams with high-strength steel bars, Eng. Struct. 216 (2020)
110706.
[38] E. Poveda, G. Ruiz, H. Cifuentes, et al., Influence of the fiber content on the compressive low-cycle fatigue behavior of self-compacting SFRC, Int. J. Fatigue 101
(2017) 9–17.
[39] B. Isojeh, M. El-Zeghayar, F.J. Vecchio, Fatigue behavior of steel fiber concrete in direct tension, J. Mater. Civ. Eng. 29 (9) (2017) 04017130.
[40] A. Parvez, S.J. Foster, Fatigue of steel-fibre-reinforced concrete prestressed railway sleepers, Eng. Struct. 141 (2017) 241–250.
[41] S. Cattaneo, L. Biolzi, Assessment of Thermal Damage in Hybrid Fiber-Reinforced Concrete, J. Mater. Civ. Eng. 22 (9) (2010) 836–845.
[42] N. Banthia, F. Majdzadeh, J. Wu, et al., Fiber synergy in Hybrid Fiber Reinforced Concrete (HyFRC) in flexure and direct shear, Cem. Concr. Compos. 48 (2014)
91–97.
[43] J. Shi, B. Liu, S.H. Chu, et al., Recycling air-cooled blast furnace slag in fiber reinforced alkali-activated mortar, Powder Technol. 407 (2022) 117686.
[44] J. Zhang, C. Liu, J. Wang, et al., Experimental study on flexural performance of precast prestressed concrete beams (PPCB) with fiber reinforcement, Buildings
13 (8) (2023) 1982.
[45] GB/T50152-2012, Standard for Test Methods of Concrete Structure, China Building Industry Press, Beijing, 2012.
[46] P.Y. Song, J.G. Han, Failure relationship of different types of steel bars in beams under fatigue load, J. Harbin Inst. Technol. 44 (8) (2012) 96–100.
[47] G.P. Titty, Fatigue testing and performance of steel reinforcement bars, Mater. Struct. 17 (1) (1981) 43–49.
[48] J.Y. Wu, J. Li, A new energy-based elastoplastic damage model for concrete. Proceedings of the 21st international congress of theoretical and applied mechanics.
Warsaw, Poland: institute of fundamental technological research, Polish Academy of Sciences, 2004, pp. 234–244.
[49] X.Y. Luo, Y.K. Chen, P.Q. Deng, Experimental study on fatigue properties of unbounded partially prestressed concrete (UPC) beam, J. Build. Struct. 28 (3) (2007)
98–104.

13

You might also like