Impact Performances of Fiber Reinforced Polymer Composit - 2023 - Composite Stru

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 67

Journal Pre-proofs

Review

Impact performances of fiber reinforced polymer composites and cables: a re-


view

Zhen Wang, Guijun Xian

PII: S0263-8223(23)00472-5
DOI: https://doi.org/10.1016/j.compstruct.2023.117128
Reference: COST 117128

To appear in: Composite Structures

Received Date: 18 July 2022


Revised Date: 28 January 2023
Accepted Date: 5 May 2023

Please cite this article as: Wang, Z., Xian, G., Impact performances of fiber reinforced polymer composites and
cables: a review, Composite Structures (2023), doi: https://doi.org/10.1016/j.compstruct.2023.117128

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version
will undergo additional copyediting, typesetting and review before it is published in its final form, but we are
providing this version to give early visibility of the article. Please note that, during the production process, errors
may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2023 Elsevier Ltd. All rights reserved.


1 Impact performances of fiber reinforced polymer
2 composites and cables: a review

4 Zhen Wang1-3, Guijun Xian 1-3, *

7 1 Key Lab of Structures Dynamic Behavior and Control (Harbin Institute of


8 Technology), Ministry of Education, Heilongjiang, Harbin 150090, China

9 2 Key Lab of Smart Prevention and Mitigation of Civil Engineering Disasters of the
10 Ministry of Industry and Information Technology, Harbin Institute of Technology,
11 Harbin 150090, China

12 3 School of Civil Engineering, Harbin Institute of Technology, Heilongjiang, Harbin


13 150090, China

14

15

16

17

18 * Correspondence: Prof. Xian, gjxian@hit.edu.cn; Tel./fax:


19 ++86(451)8628-3120;

20 Postal address: 73 Huanghe Road, Nangang District, Harbin 150090, China.


21 Abstract

22 An emerging application of fiber reinforced polymer (FRP) composites is to replace


23 steel cables for cable-supported bridges, enhancing the service lives, reducing
24 maintenance requirements, and increasing the bridge span. Vehicle collisions on the
25 cable for cable-supported bridges have been regarded as a concern since FRP
26 composites are sensitive to impact loads. The present article reviews the impact
27 performances of FRP composites and cables. Aside from the common influencing
28 factors, such as the impact energy and fiber or matrix types, the impact performances of
29 FRP cables are particularly affected by the pretension load, cable tonnage, cable forms,
30 impact position, and anchorage system. Experimental or numerical investigation of the
31 affecting factors as well as enhancement and protective methods are needed to guide the
32 anti-collision design of FRP cables in practical applications. However, high cable
33 manufacturing and testing costs, property sacrifice in material design, and lower
34 computational efficiency are possible challenges in future studies.

35

36 Keywords: Fiber reinforced polymer; Impact performance; FRP composites; FRP


37 cables

38
39 1. Introduction

40 Fiber reinforced polymer (FRP) composite, used in civil engineering, has attracted
41 growing interest in recent decades owing to its high strength-to-weight ratio, ease of
42 installation, and excellent long-term durability [1-4]. Carbon, glass, basalt, and aramid
43 fibers are commonly used in FRP composites to reinforce thermosetting and
44 thermoplastic resins [5-7]. However, one of the major concerns in their applications is
45 that these FRP composites are sensitive to impact damage [8, 9]. The visible or invisible
46 damages caused by the impact may lead to degradation in the stiffness and strength of
47 composite structures by more than 50% [8, 10]. These issues require a deep insight into
48 the impact behavior of FRP composites, utilizing experiments or numerical simulations,
49 and then a scheme to enhance their impact resistance.

50 For the FRP composites subjected to the impact, their failure mechanism is quite
51 complicated, including two or more failure modes interacting with each other, such as
52 matrix cracking, fiber breakage, and interlaminar delamination [11, 12]. At the same
53 time, the failure mechanism is dependent on the impact energies, material properties,
54 geometries, and other factors, which lead to the different mechanical responses of FRP
55 composite under impact load. According to the affecting factors, FRP composites can
56 be designed to resist the impact load [13-16]. Usually, the metal structure is insensitive
57 to the impact due to the excellent ability of energy absorption in both the elastic and
58 plastic deformation stages. Before hardening, the metal may flow for very large strains
59 at a constant rate during plastic deformation [17]. However, the failure of FRP
60 composite is brittle with irreversible intra- and inter-laminar damages under impact
61 loading, and thus it can only absorb energy through elastic deformation and damage
62 mechanisms and not via plastic deformation [17]. In order to improve the overall impact
63 resistance of FRP composites, material toughening (fiber, matrix, and interphase)
64 [18-21] is commonly used in the material design, including fiber hybridization [22-24],
65 composite fabrication using thermoplastic (TP) resins [25-27], and adding fillers [28,
66 29].

67 In cable-supported bridges, the problems of traditional steel cables, such as significant


68 self-weight, sag effect, low carrying efficiency, poor fatigue performance, and serious
69 corrosion damage, gradually emerge as bridge spans increase and the working
70 environment deteriorates [30]. An emerging application of FRP composites is to replace
71 steel cables, enhancing the service life, reducing maintenance, and increasing the bridge
72 span. However, concerns have been raised about the safety of FRP cable-supported
73 bridges since FRP cables may be subjected to low-velocity impact caused by vehicle
74 collisions (e.g., [31-36]). Several typical vehicle collisions on the cable-supported
75 bridges in the last decade are shown in Fig. 1. FRP cables are usually made of multiple
76 pultruded FRP rods or laminates. The tonnage (i.e., longitudinal load-carrying capacity)
77 of carbon fiber-reinforced polymer (CFRP) cables used in existing cable-stayed bridges

1
78 can reach more than 10 MN (e.g., The Stork Bridge [37]), but the transverse mechanical
79 performance would be much lower than the longitudinal one due to the resin-dominated
80 properties in the transverse direction. Compared to the FRP composites typically used in
81 aerospace or automotive applications, FRP cables have different molding processes,
82 section forms, loading states, and application environments [38]. Therefore, the
83 available data cannot provide a comprehensive understanding of the impact
84 performance of FRP cables. Since the FRP cable in the cable-supported bridges is still
85 an emerging technique in civil engineering applications, research on the transverse
86 impact performance is in its infancy. A review of research progress on FRP cables is
87 conducive to providing an understanding of their impact performance, identifying the
88 shortcomings in the literature, and giving direction for future research.

89 Numerical simulation is a useful tool to obtain specific information on the spatial and
90 temporal distribution of damage during the impact event and conduct the parametrical
91 study by changing boundary conditions [39]. Moreover, it was employed to study the
92 impact damage of FRP composites owing to the advantages of time-saving and
93 cost-effectiveness compared with physical impact experiments. Since the damage will
94 accumulate in the FRP composite, the progressive damage model considering damage
95 onset and growth is more acceptable than others to simulate the mechanical response
96 during the impact. The interactive strength criteria with separated expression to assess
97 fiber or matrix damage under tensile or compression loading, such as Hashin [40], Hou
98 [41, 42], and Chang-Chang [43] criteria, have been widely used in determining the
99 damage initiation of FRP composites [44]. Once the criterion is satisfied, a damage
100 evolution model reducing the initial material stiffness needs to be defined to describe
101 the process from softening to failure in FRP composites. Based on the continuum
102 damage mechanics, the failure criteria coupled with the damage evolution models would
103 be suitable for predicting the impact behavior of FRP composites [45]. Therefore, the
104 numerical modeling methods for the prediction of the low-velocity impact on FRP
105 composites were reviewed.

106 This article conducted a state-of-the-art review of the previous literature on the impact
107 performance of FRP composites and cables, including physical experiment methods,
108 impact velocity characteristics, typical mechanical responses of FRP composites,
109 improvement methods for the impact resistance of FRP composites, research progress
110 on FRP cables, and finite element modeling methods. This article mainly focused on
111 low-velocity impact events (0-11 m/s) but provided some high-velocity impact events
112 ( > 11 m/s) for comparison. In view of the extensive work conducted on FRP
113 composites, suggestions and challenges for future research on the impact performance
114 of FRP cables are provided.

115 2. Impact tests

2
116 2.1 Experimental methods

117 2.1.1 Drop-weight impact tests

118 The drop-weight impact test was extensively used to determine the impact resistance of
119 FRP composites. Fig. 2 shows a typical drop-weight impact testing instrument and the
120 fixing system [46]. The impactor would be lifted to a certain height according to the
121 impact energy and then fall to strike the target specimens, which can take the form of a
122 beam or plate. The target specimen may be simply supported or fixed. In order to
123 monitor the mechanical response of the specimen during the impact event, the impactor
124 or the support frame may be equipped with a sensor. The impact velocity history can be
125 calculated based on the impact loading history or obtained from the collection of optical
126 sensors. Compared with other low-velocity impact test methods, the drop-weight impact
127 test can measure a wider range of specimen geometries, allowing for the testing of more
128 complex components. The standard test method for the impact resistance of FRP
129 composites can be found in [47] (ASTM D7136/D7136M).

130 2.1.2 Charpy and Izod impact tests

131 Charpy impact testing was initially adopted for testing metals and then used in many
132 early impact studies on FRP composites. The energy absorption and dissipation during
133 the impact event as well as the failure mode can be tested. The specimen is usually in
134 the form of a thick beam with or without a notch and is supported freely at two ends, as
135 shown in Fig. 3 (a). The testing instrument mainly incorporates a recording scale and a
136 pendulum hammer. During the impact test, the pendulum hammer will be first raised to
137 a corresponding height according to the kinetic energy determined by the recording
138 scale and then allowed to fall following a circular trajectory to strike the midpoint of the
139 target specimen. The energy loss caused by the fracture of the specimen can be obtained
140 according to the difference between the initial and final heights of the pendulum [48,
141 49]. Besides, the strain gauge may be instrumented to the impactor, thereby allowing
142 the loading history to be collected during the impact event. Based on this information,
143 the velocity, displacement, and energy dissipation histories can be calculated according
144 to the following relationship [50]:

145 t
F (t )
v(t )  v0   dt (1)
0
m

146 t
 (t )   v(t )dt (2)
0

3
147 
E (t )   F ( )d  (3)
0

148 where v(t ) and v0 are the instantaneous and initial velocities of the impactor during

149 the impact, respectively; F (t ) ,  (t ) , and E (t ) are the impact force, displacement,
150 and energy histories, respectively; F ( ) is the impact force-displacement history; t
151 and m are the time and mass of the impactor, respectively.

152 The standard test method for the Charpy impact resistance of FRP composites can be
153 found in ASTM D6110 [51] and EN ISO 179 [52]. This test method can be performed
154 in a fully automated way with a small technical effort, and the specimen can be
155 inexpensively machined. However, the specimen is a short and thick beam, which may
156 not be representative of typical engineering applications. Besides, the specimen would
157 be tested in a destructive way, which may cause the failure mode to be inconsistent with
158 the low-velocity impact event. In the contact force history, high-frequency oscillations
159 often exist due to the natural response of the hammer. A high-frequency measuring
160 system greater than 10 kHz to ensure the accuracy of the collected data was suggested
161 in [50]. Moreover, the additional filtering of the contact force history can be used to
162 alleviate the frequency response. The set-up and process of Izod impact tests are
163 identical to those of Charpy impact tests, but the specimen is clamped as a vertical
164 cantilever beam and the pendulum hammer hits the unsupported end of the beam, as
165 shown in Fig. 3 (b). The standard test method for the Izod impact resistance of FRP
166 composites can be found in [53] (ASTM D256). According to the standard test methods,
167 the Charpy or Izod impact test can be employed to investigate the behavior of specified
168 types of specimens under the impact conditions defined and for estimating the
169 brittleness and toughness of specimens. It can also be employed to obtain comparative
170 data from similar types of material [52].

171 2.1.3 Hopkinson-bar tests

172 Hopkinson-bar tests are used to evaluate the basic material properties under high-strain
173 rate loadings ranging from 50 to 104 s-1 [54]. Gama et al. [55] reviewed the development
174 of the Hopkinson-bar test method and outlined the experimental procedure, including
175 bar calibration, specimen design, data analysis, etc. A modern split-Hopkinson pressure
176 bar (SHPB) technique is shown in Fig. 4 (a), which mainly consists of a striker bar, an
177 incident/input bar, and a transmitter/output bar. A pair of strain gauges is bonded to the
178 surface of the incident and transmitter bars, and the specimen is sandwiched between
179 these two bars. A gas gun can be employed to propel the striker bar, which hits the
180 incident/input bar, producing the stress wave that propagates through the incident bar to
181 the specimen interface. Strain gauges will record the incident and transmitted stress

4
182 waves, which can be analyzed and used to determine a dynamic stress/strain curve [56].
183 There are no uniform rules for specimen types, and specimens are often made from
184 exploratory experiments. The manufacture of the specimen should conform to the
185 general assumptions made for the SHPB technique as follows [55]: (1) the specimen
186 deforms uniformly (i.e., no friction and no inertia effects); (2) the specimen is in stress
187 equilibrium; and (3) the specimen is under a uniaxial stress condition.

188 2.1.4 Gas gun impact tests

189 The gas gun impact system was widely adopted for high-velocity impacts at ballistic
190 strain rates, as shown in Fig. 4 (b). The projectile is supported by a sabot at one end of
191 the system and propelled and accelerated by high-pressure gas (e.g., nitrogen or helium)
192 in the gun barrel so as to directly strike the specimen fixed at the other end. The
193 separator was used to separate the projectile from the sabot. The specimen can take any
194 form (mostly in plate form) according to experimental or engineering requirements. The
195 initial impact velocity can be adjusted by changing the air pressure. In order to measure
196 the projectile impact velocity, the velocimeter system can be mounted between the gun
197 muzzle and the specimen. The laser beams are obstructed successively when the
198 projectile crosses this system, triggering the transformation of the voltage signal.
199 According to the time interval between two trigger signals and the distance of the
200 velocimeters, the impact velocity can be obtained [57]. Broadly, the test is not a
201 complete destructive test but often leads to the specimen having large-scale damage
202 and/or perforation. The gas gun impact test is a very useful method to conduct the
203 high-velocity impact test, but it usually has the disadvantage that little test information
204 can be obtained during impact [58]. As a result, the instrumented gas gun has been
205 developed, which allows load-displacement signals to be acquired from the mechanical
206 response of the specimen via data collection [59].

207 2.2 Impact velocity

208 The impact event can have various ranges of velocity and can be divided into four
209 categories: low-, high-, ballistic-, and hyper-velocity. The low-velocity impact
210 (0-11m/s) may occur from dropped items or vehicle collisions. High-velocity impacts
211 (>11 m/s) may occur for aircraft that are struck by debris on the runway during takeoff
212 or landing. Ballistic (>500m/s) and hyper-velocity impacts (>2000m/s) may occur in
213 military and spacecraft applications, respectively [60]. A more concise classification of
214 impact events may be low-velocity impact by a large mass and high-velocity impact by
215 a small mass. For low-velocity impact events, the contact duration between the impactor
216 and target is long enough that the entire target will respond to the impact, and hence
217 more energy will be absorbed. Contrarily, for high-velocity impact events, the duration
218 between the impactor and target is much shorter, and the target does not have enough
219 time to generate quasi-static failure mechanisms [61]. The impact will induce a more

5
220 localized type of target response, resulting in energy dissipation in a relatively small
221 region.

222 The impact load triggers the elastic waves propagating from the impact point to the
223 adjacent region. Material damping and energy diffusion lead to the decaying response of
224 the elastic waves caused by the impact load [62]. Hence, the duration of the impact
225 corresponds to the different responses of the elastic waves. Three types of elastic waves
226 associated with impact time proposed by Olsson [62] are summarized as

227 (1) Dilatational waves: very short impact time; easily visible damage; ballistic impact
228 (Fig. 5 (a));

229 (2) Flexural & shear governed waves: short impact times; invisible damage; runway
230 debris impact (Fig. 5 (b));

231 (3) Quasi-static waves: longer impact time than the time that waves reach the target
232 boundaries; invisible damage; drop-weight impact (Fig. 5 (c)).

233 The target size and boundary conditions have an influence on the quasi-static waves but
234 not on the dilatational waves or the flexural & shear governed waves. The responses of
235 quasi-static waves and flexural & shear governed waves are also defined as
236 boundary-controlled and wave-controlled impacts, respectively [63]. For the former, the
237 whole target is deformed under the impact load, and the contact force and deformation
238 are in phase, as shown in Fig. 6 (a). For the latter, the reaction of the target is localized
239 around the impact point, and the contact force and deformation are out-of-phase, as
240 shown in Fig. 6 (b).

241 3. Impact response of FRP composites

242 3.1 Failure mechanism

243 The damage mechanism of the FRP composite is very complicated during the impact
244 event. The typical failure mechanisms are summarized as follows [64, 65]: (1) matrix
245 cracking and fiber/matrix interface debonding caused by the transverse shear stress and
246 local compression stress in the top layers; (2) matrix cracking, fiber breakage, and fiber
247 pull-out at the bottom surface owing to the bending stress; (3) fiber micro-buckling
248 caused by compression loading; (4) interlaminar delamination caused by the
249 interlaminar normal and shear stresses; (5) penetration. The failure mechanism is related
250 to the impact energy, which results in different mechanical responses for FRP
251 composites. Fig. 7 depicts the failure mechanism of FRP composites under low,
252 intermediate, and high impact energies. Under low impact energy, matrix cracking is the
253 dominant failure mode, leaving a small permanent dent in the FRP composite. Under
254 intermediate and high impact energies, fiber breakage and fiber pull-out are increased,

6
255 leading to the perforation and even penetration of the FRP composite [66]. The failure
256 modes result in different contact force-time, contact force-displacement, and
257 energy-time responses in FRP composites.

258 3.2 Contact force-time responses

259 Fig. 8 (a) shows three typical contact force-time responses with different impact
260 energies under the low-velocity impact loading. The oscillations in the curves were
261 caused by transient stress waves in the transverse direction of the FRP composites [67].
262 In the initial stage, the contact force increases with time, but the slope is decreased with
263 an oscillation at the characteristic force Fini due to the Hertzian failure. This force can

264 be considered a pointer to the ability of FRP composites to resist interlaminar


265 delamination initiation [68]. The perturbation of the contact force after Fini implies the

266 growth of the in-plane damage, and then the contact force increases to the maximum
267 impact load Fmax . For low impact energy, the contact force gradually decreases and

268 reaches zero. For intermediate and high impact energies, the contact force decreases
269 with further propagation of the in-plane damage, and the impactor penetrates into the
270 target. The energy stored in the impactor in the intermediate energy case is exhausted
271 before the front of the impactor reaches the back layer of the FRP composite. The front
272 of the impactor passes through the composite materials in the high energy case. The
273 plateau stage (the post-perforation frictional section [69]) in the force-time responses is
274 caused by friction between the impactor and target.

275 3.3 Contact force-displacement responses

276 Fig. 8 (a) presents the typical contact force-displacement responses for three types of
277 impact energy events. For low impact energy, the contact force increases with the
278 displacement until it reaches the maximum loading value, and then the contact force and
279 displacement decrease simultaneously due to the impactor rebound. This type of curve
280 implies that the target does not experience too serious damage, and the dominant failure
281 mode is matrix cracking. Differently, the contact force-displacement responses for the
282 other two energy events show that the contact force decreases with the increased
283 displacement after the ascending stages. This means that the impact induces serious
284 damage and failure in the matrix and interphase (i.e., fiber to matrix and ply to ply), and
285 fiber tensile breakage and fiber compressive bucking take place in the target specimen
286 and intensify as the impact energy increases. It also means that most of the impact
287 energy is dissipated in the form of fiber breakage, and the deformation caused by the
288 impact is unrecoverable. These two curves also represent the perforation and penetration
289 failure modes, respectively. The shaded area below the load-displacement curves

7
290 denotes the energy absorbed by the target (i.e., absorbed energy Ea ), while the

291 post-perforation section was excluded for the purpose of identifying the true energy
292 absorption caused by damage formation in the target [69].

293 3.4 Energy- and velocity-time responses

294 Fig. 8 (b) depicts the typical energy- and velocity-time responses under the low,
295 intermediate, and high energy impacts. The kinetic energy stored in the impactor was
296 transferred to the target during the impact event. For low impact energy, the energy
297 initially increases with time and then decreases at the maximum energy (i.e., total
298 energy) and finally reaches a stable value. The final value is defined as the energy
299 absorbed by the target specimen, which is dissipated in the form of fiber/matrix damage,
300 delamination, and friction between the impactor and the target or among neighboring
301 plies within the FRP composite. The value from the maximum energy to the final
302 energy is regarded as the elastic energy Eel , which leads to the elastic deformation of

303 the target. For the intermediate and high energy impacts, the initial ascending stage is
304 the same as the low energy event, while the energy continues to increase due to the
305 perforation and penetration failures. Therefore, an estimated point in the contact
306 force-displacement response is used to define the end of the energy absorbed by the
307 specimen, as shown in Fig. 8.

308 The velocity of the impactor possesses the highest value at the beginning of the impact
309 event, as shown in Fig. 8 (b). For low impact energy, the velocity of the impactor
310 decreases to zero when its kinetic energy is completely transferred to the target.
311 Subsequently, the velocity increases since the elastic energy stored by the target
312 specimen is transferred back to the impactor, which makes it rebound. The value of the
313 velocity at zero corresponds to the maximum energy value. For the other two energy
314 impacts, the velocity-time responses have no ascending stages, implying no rebounding
315 due to the perforation and penetration failure mechanisms.

316 4. Improvement of impact resistance of FRP composites

317 FRP composites mainly include fiber, matrix, and their interphase. In view of failure
318 mechanisms, various methods have been conducted for the purpose of improving the
319 impact resistance of these three components to delay or avoid matrix cracking, fiber
320 breakage, and interlaminar delamination. Among the enhancement methods, efficient
321 fabrication, lower cost, and less consumption of current mechanical properties are
322 always considered in the application. At present, fiber hybridization, matrix toughening,
323 and modification (e.g., using thermoplastic resins and adding fillers) are the most
324 accepted methods to improve the impact resistance of FRP composites.

8
325 4.1 Fiber hybridization

326 The fiber system is the primary load-bearing constituent, providing the composite
327 material with the majority of its strength and stiffness [10]. Carbon, glass, aramid, and
328 basalt fibers are primarily used in FRP composites. Carbon fibers are the most widely
329 used in many structural applications owing to the designable elastic moduli ranging
330 from 207 GPa to 1035 GPa [70] and the highest tensile strength, but they are also the
331 most brittle with an elongation of 0.5% to 2.4%. Glass fibers are also frequently applied
332 to reinforce the polymer matrix due to their low cost, high tensile strength, and excellent
333 insulating properties [71]. The disadvantages are a relatively low tensile modulus and
334 fatigue resistance. Aramid fibers are highly crystalline aromatic polyamide fibers,
335 possessing the lowest density and the highest strength-to-weight ratio among other
336 fibers. Aramid fibers have high tensile strength and excellent impact resistance but a
337 large creep deformation. Aramid fibers are always used in anti-ballistic composites,
338 such as body armor [72], but they are not widely used in structural engineering fields
339 these days. Basalt fibers are made from finely powdered basalt, which is burned at
340 around 1500–1700 °C and then extruded from a glassy molten liquid into the form of
341 thin threads [73]. The mechanical properties of basalt fiber are similar to those of glass
342 fiber, but it has superior thermal characteristics. Basalt fiber is considered an alternative
343 to glass fiber, but its mechanical properties are mainly dependent on the quality of the
344 natural basalt. In FRP composites, the fundamental mechanical properties provide a
345 useful evaluation of the energy absorption and dissipation, and the area under the
346 stress-strain relationship provides an important implication for the impact resistance of
347 composites. A larger area means a higher energy-absorbing ability under an impact
348 load.

349 To enhance the impact resistance of FRP composites, a hybrid fiber system is one of the
350 most effective methods. The commonly used hybrid patterns for FRP composite (Fig. 9)
351 are summarized as follows [74]:

352 (1) In the layer-by-layer hybridization, the two types of fiber fabric are stacked onto
353 each other (Fig. 9 (a)), which is the most simple and economical method to fabricate a
354 hybrid composite.

355 (2) In the yarn-by-yarn hybridization, the two types of fiber yarn are mixed within the
356 layers, such as different yarns co-woven into a fabric (Fig. 9 (b)) or intralayer
357 hybridization with parallel bundles.

358 (3) In the fiber-by-fiber hybridization, the two types of fiber are mixed or co-mingled
359 directly, as shown in Fig. 9 (c).

360 Hybridization in FRP composites often combines the load-carrying capacity of high

9
361 modulus fibers with the toughness of low modulus fibers for the purposes of: (1)
362 improving the in-plane and/or out-of-plane toughness of the brittle FRP composite (e.g.,
363 CFRP); (2) enhancing the in-plane and/or out-of-plane load-carrying capacity and
364 stiffness of the ductile FRP composite (e.g., natural fiber composite); (3) decreasing the
365 cost of FRP composites. However, hybridization cannot always ensure a desirable
366 increase in impact resistance. The fiber layer configuration [75] should be carefully
367 considered since it can change the flexural properties and the damage mechanisms [74].
368 For instance, Chen et al. [76] employed the combination of carbon, glass, and aramid
369 fibers to fabricate FRP composites with the sandwich-like layup sequence and found
370 that the carbon layers placed in the center layer present superior impact resistance
371 compared with those in the top and bottom layers. Papa et al. [77] studied the impact
372 performance of the carbon/glass hybrid composites with symmetric and asymmetric
373 layup sequences. They found that the asymmetrical layup sequence with the glass fiber
374 situated on the impact side revealed the best impact performance compared to the
375 symmetric layup and the asymmetric layup with the glass fiber placed on the bottom. In
376 the previous review paper, Swolfs et al. [74] summarized the influence of two types of
377 layup, including layer positioning and dispersion, on the impact resistance of hybrid
378 fiber composites and concluded that increased dispersion improves impact resistance
379 and post-impact properties, and positioning the fibers with the highest energy
380 absorption on the outside enables the composite to absorb more energy.

381 Natural fibers have the advantage of low density, acceptable specific strength
382 properties, renewability, and low cost, while lower durability, high moisture absorption,
383 and lower impact strength are the prominent shortcomings [78]. Natural fibers are
384 usually acquired from animals, plants, or rocks, such as bamboo, hemp, sisal, flax, and
385 basalt [78]. Natural-fiber-reinforced polymer composites are still limited to
386 non-structural or sub-structural applications, such as vehicle, airplane, and building
387 interiors [79]. A hybridized fiber system makes natural fibers have great potential to be
388 applied in engineering structures owing to a balance between reduced costs and optimal
389 mechanical properties compared to single fiber systems [80, 81]. Incorporating
390 synthetic fibers into natural fibers can enhance the in-plane and out-of-plane properties
391 of natural fiber composites. Petrucci et al. [82] studied the impact resistance of hybrid
392 composites using flax, hemp, and glass fibers and found a larger deformation in the
393 absence of glass fiber. Bandaru et al. [83] studied the impact response of 3D
394 angle-interlock Kevlar/basalt fiber composites and found that the composites using
395 basalt yarns as warp reinforcement and Kevlar yarns as weft and binders result in an
396 improvement of 7.67-46.49% in energy absorption compared to the non-hybrid
397 composites. Natural fibers are hydrophilic because of their porous structure, which
398 allows for water absorption [78], but they are vulnerable to water molecules due to the
399 hydroxyl groups on the surface [84, 85]. The stacking sequence, which places the
400 synthetic fibers as the outer layers and the natural fibers as the core, not only provides a

10
401 barrier to the water molecules but also enhances the impact resistance. Some recent
402 papers [81, 84, 86] found that placing the carbon fiber on the surface side with a
403 symmetric layup can provide minimum water absorption and maximum impact
404 resistance compared with other forms of fiber configuration.

405 4.2 Composite fabrication using TP resins

406 The matrix in the FRP composite serves to protect, align, and stabilize the fibers,
407 ensuring stress transfer between fibers. The matrix system can be divided into two
408 categories: thermoplastic (TP) and thermoset (TS) resins. TS resins are the traditional
409 matrix system and are widely used in FRP composites for various engineering
410 applications. Compared with TP resins, the most prominent advantage of the TS resin
411 lies in its low viscosity, which makes for easier wetting out between the fiber and
412 matrix [87]. The impregnation quality is very important to achieve fiber-matrix
413 interaction and good mechanical properties for FRP composites [88]. Furthermore, TS
414 resins exhibit significantly less creep and stress relaxation. However, their impact
415 resistance is limited by their low fracture strain. Composites with the TS matrix system
416 are more susceptible to interlaminar delamination failure due to their brittle nature [78].
417 The development and fabrication of the TS matrix with high toughness may be one of
418 the most effective solutions for improving impact resistance.

419 TP resins have better impact resistance and damage tolerance properties owing to their
420 higher fracture toughness compared with TS resins. The impact resistance and damage
421 tolerance for TP and TS composites can be found in recent review papers [60, 67]. Shah
422 et al. [89] studied both the TP and TS matrix-based 3D glass fabric composites under
423 single and multiple low-velocity impact events. The TP composites show much better
424 performance under a single impact event, including less damaged area, higher damage
425 transition energy, and higher load-bearing capacity. For multiple impact events, the TP
426 composites survive more impacts and indicate superior impact resistance compared with
427 TS composites. The authors [90] also investigated the post-impact properties of TP and
428 TS composites and found that TP composites are less susceptible to impact-induced
429 damage than TS composites. Earlier, Vieille et al. [27] found that the post-impact
430 residual strength of PEEK-based laminates is about 10% higher than epoxy-based
431 laminates at low impact energy, while the residual strength is virtually the same for the
432 two types of laminates at high impact energy due to a significant permanent indentation
433 for TP laminates. The same authors [91] investigated the impact performance of TP and
434 TS composites subjected to different impact energies and found that the delamination in
435 TP laminates is lower than that in TS ones since the fiber bridging refrains the plies
436 from opening in mode I failure and reduces the growth of interlaminar and intralaminar
437 cracks in modes II–III failure. Recently, Kazemi et al. [92] studied the low-velocity
438 impact on the composite laminate using a newly developed TP resin, Elium®, that can
439 be cured at room temperature. Compared to TS-based laminates, TP-based laminates

11
440 have a higher impact load and improved structural integrity by 240%. The newly
441 developed TP resins can save production time and avoid thermal stress concentration at
442 a high temperature during fabrication.

443 The fracture toughness of TP composites is ten times higher than that of TS composites
444 [93, 94], which enables TP ones to present better impact performance. However, most
445 TP composites need to be fabricated at a high process temperature with expensive
446 equipment. Moreover, incorporation of continuous fibers into TP resins is difficult due
447 to the high melt viscosity of TP resins, which leads to poor interphase and integrity.

448 4.3 Interface system reinforcement

449 The bond performance between fiber and matrix is important to the mechanical
450 behavior of FRP composites. Excellent interphase can decrease the stress concentration
451 in the loading transfer process from matrix to fiber and improve the interfacial
452 performance of composites [95], while weak interphase leads to the FRP composite
453 being more susceptible to fiber pull-out damage or interlaminar delamination damage
454 under impact loads. The stiffness of the structure will be reduced depending on the
455 delamination area. This will intensify the fiber and matrix damage and then prompt
456 delamination propagation under high transverse shear loading and further degradation in
457 the stiffness and strength of FRP composites. Several methods have been verified to be
458 effective in improving the interlaminar properties of FRP composites, such as
459 3D-weaving, stitching, and Z-pinning [96]. However, these methods may significantly
460 reduce the in-plane properties of FRP composites. In order to avoid the consumption of
461 in-plane properties, matrix toughening by adding micro- or nano-sized fillers into the
462 matrix (Fig. 10 (a)), interleaving film, fibers, or particles between neighboring plies
463 (Fig. 10 (b)), and grafting CNTs directly on the fibers (Fig. 10 (c)) can be used to
464 enhance the interface properties of FRP composites.

465 Nanoparticles applied to the matrix can be classified into zero- (e.g., rubber, carbon
466 black, and fullerene), one- (e.g., single-walled carbon nanotubes (CNTs), multi-walled
467 CNTs, and cup-stacked CNTs), two- (e.g., nano-plates, nano-sheets, and nano-clays),
468 and three- (e.g., nano-carbon aerogels) dimensional structures [97]. Most of the
469 published studies present a positive effect of the matrix adding nanoparticles on the
470 interlaminar Mode I fracture toughness, interlaminar Mode II fracture toughness,
471 delamination area, and post-impact strength. For instance, Yokozeki et al. [98] studied
472 CFRP laminates with modified epoxy matrix using cup-stacked CNTs and found that
473 the fracture toughness in mode I and mode II was increased by 97% and 30%,
474 respectively. Ashrafi et al. [99] modified the matrix using single-walled CNTs and
475 found a 5% decrease in the impact damage area, a 13% increase in mode I fracture
476 toughness, and a 28% growth in mode II fracture toughness. Landowski et al. [100]
477 studied the impact damage of carbon fiber-epoxy composites with the addition of 1~8

12
478 wt% SiO2 nanoparticles in the matrix and found a maximum reduction of 28% in the
479 damage size. A combination of two or more fillers with different length scales can also
480 improve the impact resistance of the composite materials. Ravindran et al. [101] studied
481 the impact resistance of carbon fiber-epoxy composites enhanced using carbon
482 nanofibers (CNFs) or carbon short fibers (CSFs) and a combination of both. The authors
483 found that the addition of CNFs or CSFs in the matrix reduces the delamination area
484 under the impact load, and a combination of these two fillers results in the highest
485 reduction owing to the enhancement of both interlaminar mode I and mode II fracture
486 toughness properties. Besides, the impact resistance and interlaminar fracture toughness
487 were improved without consuming the mechanical properties such as flexural and
488 interlaminar shear strengths.

489 Differently, some works show an insignificant or even adverse effect on the impact
490 characteristics of the FRP composites by adding nanoparticles into the matrix [98, 102,
491 103], which may be attributed to the type and content of the nanoparticles and
492 manufacturing methodologies [29] or the impact energies and impact velocity [102]. On
493 the other hand, adding fillers may lead to an increase in the initial viscosity of the resin
494 matrix [104, 105], which limits flow for a given cure profile and makes it more
495 challenging to achieve a void-free composite due to incomplete fiber wet-out [106].

496 The interlaminar properties of the FRP composite can be enhanced through interleaving
497 films, sheets, or particles between neighboring layers, as shown in Fig. 10 (b). Hence, it
498 is reasonable to expect an increase in the impact properties using this method due to the
499 dominant failure mode of delamination under impact load. Akangah et al. [107] studied
500 the impact resistance of CFRP laminate interleaved with electrospun Nylon-66
501 nanofabric and found that the threshold impact force was increased by about 60% and
502 the impact damage growth rate was decreased from 0.115 to 0.105 mm2/N. Xin et al.
503 [96] assessed the impact and post-impact properties of CFRP composites inserted with
504 continuous sheets of randomly oriented multi-walled CNTs. The delamination area was
505 reduced by 11%-39% after impact events with 5J, 10J, and 20J impact energies, and the
506 residual flexural strength was increased by 24% after 20 J impact. The improvement in
507 the impact properties is due to nanotube pull-out and crack bridging on the microscopic
508 scale, which increase fracture toughness and provide stronger adhesion at the
509 fiber/matrix interface.

510 Grafting CNTs directly on the fibers is also an effective method to improve the
511 interfacial properties between the fiber and matrix, as shown in Fig. 10 (c). The
512 improvement of the interface properties lies in increased chemical bonding, mechanical
513 interlocking, and/or local stiffening of the polymer chains at the fiber/matrix interface,
514 all of which enable the interface to transfer stress more effectively [95]. Wu et al. [108]
515 studied the interfacial properties of carbon fiber reinforced methylphenylsilicone resin
516 composites that were modified using multi-walled CNTs grafted onto the carbon fibers

13
517 and employed saline as the bridging agent. The results show an improvement of 53.1%
518 and 33.17% in the interlaminar shear strength and the impact toughness, respectively.
519 Zhao et al. [109] employed a layer-by-layer grafting method to deposit CNTs on the
520 fiber surface and assessed the interfacial properties of the modified CFRP composites.
521 The authors found that the interfacial shear strength was increased by 117.7% and the
522 impact strength was enhanced by 49.8% without reducing the in-plane properties. The
523 mechanism of improvement in impact strength by the grafting method was also
524 disclosed by the authors, as shown in Fig. 11. Because of the weak adhesion in the
525 interface for the neat composite, the crack tip caused by the impact force can directly
526 contact the fiber surface, whereas the CNFs on the fiber surface can alleviate the
527 interfacial stress concentration and cause the crack tip to turn to the interphase.

528 5. Impact performance of FRP cables

529 The FRP composite applied to civil engineering structures has aroused big interest in
530 the last two decades [30]. Nowadays, FRP cables used in cable-supported bridges have
531 different forms, such as parallel wire cable, twisted wire cable, plate cable, and sheet
532 cable [30, 37], as shown in Fig. 12. FRP cables are connected to the bridge through
533 anchorage systems. The bonding material or mechanical interlock may serve as the
534 connection between the FRP composites and the anchorage (Fig. 13).

535 5.1 Factors affecting the impact resistance of FRP cables

536 The threat to FRP cables from vehicle collisions has attracted attention in recent
537 research [110-112]. Fang et al. [110, 112] conducted a drop-weight impact test to
538 investigate the transverse impact behavior of the scaled CFRP cable made of CFRP
539 twisted wires, as shown in Fig. 14 (a-b). Prior to the impact event, a pretension load was
540 applied to the CFRP cable to simulate the loading state of the cable in cable-supported
541 bridges. A 30% pretension level makes the specimen more susceptible to damage due to
542 the combination of the tension, bending, and shear loads. The effect of impact energy on
543 the impact response of CFRP cable was investigated by the above-mentioned group.
544 The damage patterns of the scaled cables are indentation, partial fracture, and complete
545 fracture as the impact energy increases. Under a higher energy impact, CFRP cables
546 present a brittle failure after the impact load reaches the maximum value, as shown in
547 Fig. 14 (c). Fang et al. [113] also studied the low-velocity impact behavior of CFRP
548 wire under different pretension loads. The experimental results show that the stiffness of
549 CFRP wire increases with the increased pretension load, but the maximum impact load
550 seems to be irrelevant to the pretension load level, as shown in Fig. 14 (d). Recently,
551 Wang et al. [38] reviewed the low-velocity impact on the CFRP cables and concluded
552 that the existing studies cannot provide a full understanding of the impact behavior of
553 CFRP cables and that more research should focus on large-tonnage CFRP cables and a
554 strict anti-collision design should be carried out for the CFRP cable.

14
555 Aside from the CFRP cables, some studies [114, 115] have focused on the low-velocity
556 impact behaviors of BFRP tendons. Zhu et al. [114] considered the effect of the impact
557 energy and the pretension load on the impact performance of BFRP tendons and the
558 residual tensile strength after impact. They found that the failure mode changed from
559 resin failure to partial fiber fracture as the pretension load increased, and the relations
560 between the pretension load and the mechanical responses, such as the residual
561 deformation, energy consumption, and impact contact time, depended on the impact
562 energy. Furthermore, as impact energy increases, the residual tensile strength of BFRP
563 tendons decreases. Li et al. [115] studied the impact behavior of BFRP tendon after
564 exposure to the saline-alkaline environment and found the impact resistance decreased
565 as exposure time increased, due to resin hydrolysis, basalt fiber corrosion, and interface
566 deterioration.

567 For FRP cables, the studies mainly considered the effects of pretension load and impact
568 energy. Other important factors, however, such as cable tonnages, cable forms, and
569 impact positions, were not investigated in the previous research. The understanding of
570 the impact behaviors of FRP cables is not enough for the design of the cable to resist
571 impact loads.

572 5.2 Reliability of the anchorage system

573 The anchorage system of FRP cables withstands not only the tensile force but also the
574 combined shear force and bending moment [30]. Under the impact load, the sudden
575 increase in tensile load may cause the debonding of FRP cable from the anchorage
576 system, and the shear and compressive stresses may cause a notch effect [116], leading
577 to failure in the anchorage zone. Recently, Fang et al. [117] considered the effect of an
578 impact load on the bond behavior of ultra-high performance concrete (UHPC)
579 bond-type anchorage systems for CFRP tendons through static and impact tests. They
580 found that slip failure from the anchorage system occurs for the specimen with a
581 200-mm-long bond length under impact loads but not for the same specimen under
582 static loads. They attributed the failure to weakened mechanical interlocking under
583 impact loads. However, the findings may not be suitable for the most commonly used
584 anchorage systems with epoxy adhesives. In the future, more work needs to be carried
585 out on the reliability between the FRP cable and the anchorage system.

586 5.3 Material design for enhancing impact resistance of FRP cables

587 To the best knowledge of the authors, no work has been done on how to enhance the
588 impact resistance of FRP cables [110, 111, 113, 115]. Based on the research conducted
589 on FRP composites, material design through fiber hybridization, matrix toughening, and
590 fiber/matrix interface reinforcement are expected to be applicable to enhancing the
591 impact resistance of FRP cables. For the hybrid fiber system, property sacrifices, such

15
592 as tensile strength and fatigue performance, should be controlled in the material design
593 to achieve the optimal properties of FRP cables. Matrix toughening may be used with
594 the pultrusion process for the production of FRP rods/plates, but adding fillers may lead
595 to an increase in the viscosity of the resin, which is unacceptable for the pultrusion
596 fabrication. It is necessary to ensure that the fillers do not increase the viscosity of the
597 resin above the critical viscosity for pultrusion [105]. Using high-toughness resin
598 matrices, such as polyurethane, may be a helpful method for enhancing the impact
599 resistance of pultruded rods or plates [118-120]. In addition, using TP resin may not be
600 an ideal method at present due to its high cost, difficult wetting out of continuous fibers,
601 and possible creep.

602 5.4 Protective design for FRP cables

603 The requirement for the impact resistance of FRP cables may not be achieved only
604 through material design. Adding an outside layer of protection, which has the added
605 benefit of allowing the energy absorption capacity of FRP cables to be tailored to
606 individual needs, is regarded as one possible strategy for dramatically enhancing their
607 impact performance. Fang et al. [121] investigated the Charpy impact resistance of
608 CFRP tendons with protective layers of polyvinyl chloride (PVC) and polyethylene
609 (PE). They found that the energy absorption capacity was improved by 21%-433%
610 using the protective layer, and the PE layers showed the highest protection effectiveness
611 in terms of the increased absorbed energy and additional structural weight. An empirical
612 expression was proposed to calculate the relationship between the thickness of PE
613 layers and the energy absorption capacity.

614 High-density polyethylene (HDPE) layers are always used in the fabrication of steel
615 cables [122], mainly acting as a protective layer (e.g., anti-corrosion and anti-scratch).
616 HDPE has exceptional impact strength, making it one of the best impact-resistant
617 thermoplastics available [123]. This method can be easily applied to the FRP cable
618 during fabrication. However, the protective effectiveness of the HDPE layer on the FRP
619 cables subjected to the impact load has not been investigated. The HDPE layer may be
620 unable to withstand the high-energy impact caused by vehicle collisions. Additional
621 protective systems need to be considered to protect impact-prone positions and weak
622 cable positions (e.g., the transition area between the anchorage and cable). The
623 protective methods, such as material selection and configuration design, need to be
624 investigated systematically to develop an effective energy absorption system for FRP
625 cables.

626

627 6. Numerical modeling of the impact response of FRP composites

16
628 Numerical modeling is a useful approach to predicting the damage and failure of FRP
629 composites under the impact load, which can save the cost and time of physical testing
630 [124] and reproduce the impact event as well as describe the internal damage
631 mechanisms in detail through an appropriate numerical model [125, 126]. Therefore, the
632 finite element (FE) approach is preferred by many researchers to predict the damage to
633 FRP composites caused by the impact event. In the FE simulation, it is necessary to
634 build a model following the practical conditions (e.g., loading conditions, geometry,
635 geometric dimensions, boundary conditions, etc.) as closely as possible, although some
636 assumptions may be inevitable in some cases in order to simplify the model and
637 improve the computing efficiency [127]. The mesh size and type should match the
638 model complexity to improve computational efficiency and accuracy.

639 The damage mechanism of FRP composites subjected to impact loads is extremely
640 complex, with several failure modes interacting with one another. These behaviors can
641 be predicted using constitutive models provided by several analytical studies [128-131].
642 The continuum damage mechanics (CDM), coupled with the damage initiation criterion
643 and the damage evolution model, have become the most popular approaches to
644 simulating the damage onset and growth in FRP composites [132]. In damage initiation
645 criteria, polynomial expressions were employed to define the interaction between the
646 stress or strain caused by the load and the peak stress or strain. In the damage evolution
647 model, stiffness degradation based on the initial elastic modulus was used to model the
648 softening and failure of FRP composites. Therefore, reliable testing data (i.e., original
649 material properties) is required to obtain an accurate simulation. The material
650 parameters of the FRP composites used in the FE model are collected in Table A.2. The
651 damage in FRP laminated composites was generally classified by two main failure
652 modes: (1) intralaminar damage, including fiber and matrix damages; and (2)
653 interlaminar damage or delamination between neighboring layers.

654 6.1 Intralaminar damage prediction

655 Fiber tension, fiber compression, matrix tension, matrix compression, and shear failure
656 were regarded as the main intralaminar failure modes for the FRP composites under the
657 different stress states. Therefore, it is necessary to obtain an accurate model that can
658 predict these typical failure modes to simulate the mechanical response of FRP
659 composites in all possible stress configurations.

660 6.1.1 Fiber/matrix tension/compression damage prediction

661 Previously, several criteria based on static testing were proposed to predict the
662 intralaminar damage of the FRP composite. For instance, Tsai-Wu [133] proposed the
663 simplest method by using the maximum stress with seven parameters interacting with
664 each other to predict damage initiation. Hill [134, 135] proposed the Tsai-Hill criterion,

17
665 which is derived from a yield criterion for orthotropic ductile material and applied to
666 orthotropic material by extending the classical von Mises yield condition. However,
667 these failure models cannot provide a valid prediction since they are derived from the
668 metal theories, in which composites were regarded as homogenous materials without
669 considering the heterogeneity and interactions of their constituents [136]. To improve
670 prediction accuracy, the damage criterion considering separated failure modes in the
671 FRP composite was proposed in the previous studies, as shown in Fig. 15. Hashin [40]
672 considered fiber and matrix failure in unidirectional fiber composites as a separate
673 damage mode and proposed a failure criterion in the form of quadratic stress
674 polynomials. The Chang-Chang [43] criterion was proposed using a laminated
675 composite with a circular hole at the center that contains the stress concentration. Based
676 on a “modified Mohr hypothesis’’, Puck & Schürmann [137] proposed a fracture plane
677 in the damage initiation criterion that the matrix damage is prone to initiate with an
678 inclined angle. Similarly, the LaRC04 criterion [138] is a fracture-plane based failure
679 criterion, but its fracture conditions are derived from a fracture mechanical point of
680 view, and it extends the fracture behavior in the fiber direction under compression and
681 considers the so-called in-situ effect [139]. In the Second World-Wide Failure Exercise
682 (WWFE-II) [140], the performance of 12 leading failure criteria for predicting the
683 mechanical behavior of FRP composites when subjected to 3D states of stress was
684 evaluated through challenging tests [141]. Fig. 16 presents the ranking of the theories
685 for the combined quantitative and qualitative assessment, where theories Pinho [142],
686 Carrere [143], and Cuntze [144] show better performance by comparison between the
687 prediction and the experiment.

688 Once the failure criterion is satisfied, the damaged region of the composite material
689 softens, resulting in a reduction in its stress-carrying capacity due to local stiffness
690 degradation. Some researchers considered parameters for the stiffness degradation to
691 simulate the process of damage accumulation (e.g., [44, 145-148]), which obviously
692 lacks physical meaning and is inadequate for a wide range of situations. Based on the
693 CDM and fracture mechanism, the damage evolution model is more physical and
694 acceptable to describe the stiffness degradation of the FRP composite after the damage
695 initiation. In this model, the initial stiffness was degraded by means of a damage
696 variable ranging from zero (undamaged material) to unity (material failure). The
697 exponential and linear models using the equivalent strain/displacement methods are two
698 of the most prevailing models to describe the stiffness degradation of FRP composites,
699 as shown in Fig. 17 [136, 149]. The commonly used damage initiation criterion and
700 damage evolution model based on the CDM framework to predict fiber/matrix
701 tensile/compressive failure of FRP composites are given in Appendix A.

702 Nowadays, part of the criteria has been employed to simulate the initiation of damage in
703 FRP composites under the impact load. Previous studies [150, 151] evaluated the

18
704 accuracy of various failure criteria in predicting the impact response of FRP composites.
705 The authors found that the Puck & Schürmann criterion has the minimum deviation but
706 consumes more time on computation, while the Hashin and Chang-Chang criteria are
707 more suitable to be used due to the acceptable predicted results and the simple
708 expression. Similarly, Shao et al. [152] compared the Hashin and Chang-Chang criteria
709 and found a better predicted result for lower impact energy by the use of the
710 Chang-Chang criterion but a closer prediction for higher impact energy when the
711 Hashin criterion was adopted. Li et al. [153] compared five criteria, including max
712 stress, Tsai–Wu, Hashin, Puck, and Hou criteria. The max stress criterion leads to the
713 lowest total damage area because it ignores the shear stress components, and the Tsai–
714 Wu criterion presents the greatest irreversible displacement since the damage in the
715 tensile and compressive directions were not considered separately. Besides, the
716 equivalent displacement and strain methods have a different influence on the numerical
717 results. The former method shows a larger damage area, while the latter indicates a
718 more rapid reduction in stiffness. Lopes et al. [154] developed an FE model using
719 LaRC04 failure criteria [138] coupled with an exponential damage evolution model to
720 predict the low-velocity impact damage on dispersed stacking sequence laminates. The
721 authors found that the FE model can provide a reliable prediction for impact energies up
722 to 30 J, but the accuracy appears to decrease as the impact energy and number of
723 interfaces increase. Zhou et al. [132, 155] proposed a modified method to calculate the
724 equivalent stress in the 3D Hashin criterion, which considers the through-thickness
725 normal stress component. The authors found that the accuracy of the numerical
726 simulation can be improved during lower energy impacts, but the positive effect was not
727 apparent for higher energy impacts.

728 The numerical simulation is not only used in predicting the impact response of 2D FRP
729 composites but also that of 3D composites. Bandaru et al. [83] employed the
730 Chang-Chang failure criterion to predict the impact response of 3D woven composites.
731 The mechanical response and failure modes in terms of fiber breakage, matrix cracking,
732 and yarn pullout are well predicted using this model. Miao et al. [156] built a 3D FE
733 model in which ductile and shear damage criteria [157] were used to simulate the
734 damage onset and growth of resin matrix and the Hashin criterion to predict the damage
735 of impregnated yarn. In this model, resin matrix and reinforcement, including warp
736 yarns, weft yarns, and Z-binders, were considered in the 3D woven composite. The
737 characteristic damage modes were well captured by this model. Muñoz et al. [158]
738 employed the simplified version of the LaRC04 failure criterion [138, 159] to simulate
739 the impact damage of each ply for the 3D woven composite. The authors found that the
740 impact response was reasonably predicted by the FE model, and the modeling strategies
741 of unidirectional laminates (e.g., homogenization of ply properties and the use of CDM)
742 can also be implemented into 3D woven composites. Recently, Shah et al. [160]
743 proposed a 3D multi-scale FE model to predict the impact response of

19
744 thermoplastic-based 3D woven composites. In this FE model based on the CDM
745 approach, the macro-level stress-strain response is obtained using the solid element. The
746 failure in individual impregnated yarns and matrix regions was calculated according to
747 the deformations of each element by the use of a meso-scale analytical unit-cell model
748 of the 3D orthogonal woven composite.

749 For the simulation of low-velocity impact, the basic material properties used in the FE
750 model can be approximately obtained from the quasi-static test due to the relatively
751 lower strain rate in the low-velocity impact events [161]. However, this strategy is not
752 true as the impact velocity increases due to the strain-rate properties of FRP composites.
753 The material properties at different strain rates can be tested using the HSPB system.
754 Cui et al. [162] employed the Puck and LaRC failure criteria to model the impact
755 response of IM7/8552 composites, with the impact velocity ranging from 21 m/s up to
756 157 m/s. The material properties tested at a high strain rate were implemented into the
757 Puck criterion, while a rate-dependent model was introduced to the LaRC criterion. The
758 comparison between the experiment and simulation shows a deviation of macroscopic
759 damage to the laminate at elevated projectile velocities, due to a lack of dynamic
760 material data and a mismatch between the complex fiber architecture and its numerical
761 representation.

762 6.1.2 Shear damage prediction

763 FRP composites may show nonlinear and irreversible shear behavior due to matrix
764 plasticity, matrix cracking, and fiber/matrix interface debonding. Several methods have
765 been proposed to describe the nonlinear shear behavior of FRP composites, such as
766 continuum theories of plasticity [163], continuum damage mechanics [164], or a
767 combination of both theories [165]. For in-plane shear loading, deformations of FRP
768 composites are dominated by the matrix performance, which may behave inelastic or
769 irreversible due to the matrix cracking and plasticity, which can lead to permanent
770 deformation in the ply. To describe matrix cracking and fiber/matrix interface
771 debonding, Ladeveze and Dantec [166] proposed a shear damage model that considers
772 both plasticity and damage theories. This model was expressed using material stiffness
773 loss and the emergence of permanent strains. Further, Donadon et al. [130] improved
774 this shear damage model, which required fewer material parameters and defined the
775 final failure in shear when the crack saturation parameter reaches a unit value, as shown
776 in Fig. 18. Detailed information about the shear damage model proposed by Donadon et
777 al. [130] can be found in Appendix A. Based on the previous concept, other developed
778 shear damage models can be found in [167-171].

779 For the FRP composites under the impact loading, the FE simulation can employ the
780 non-linear shear damage model coupled with intralaminar damage criteria to simulate
781 matrix damage in shear and fiber and matrix damage in tension and compression.

20
782 Faggiani et al. [172] employed the nonlinear shear damage model proposed by Donadon
783 et al. [130] to represent the shear response of a stiffened composite panel under impact
784 load and used the quadratic strain criterion and Puck criterion to predict fiber/matrix
785 tensile/compressive damage. This FE model presents a good prediction of the
786 mechanical response and the extent of damage as compared with the experiment.
787 Similarly, Shi et al. [45, 173] proposed a 3D FE model with Hashin and Puck criteria to
788 predict fiber and matrix damage in tension and compression and a nonlinear shear
789 damage model with a semi-empirical expression [174, 175] to simulate shear behavior.
790 The peak force, energy, and displacement were well reproduced, while the impact
791 response during the impactor rebound was less accurately predicted. The difference
792 between experimental and numerical results was reduced with the increased impact
793 energy. Liu et al. [176] proposed a 3D FE model to model the impact response of hybrid
794 unidirectional/woven CFRP laminates, which incorporated a unidirectional constitutive
795 model, a woven constitutive model, and a nonlinear shear damage model to simulate the
796 damage within a ply. For the multiple impact events, Rezasefat et al. [177] also
797 employed the shear damage model proposed by Falzon et al. [171] to simulate the shear
798 damage of CFRP composite plates. Other attempts using the nonlinear shear damage
799 model to represent the shear behavior under the impact load can be found in [170, 178].

800 6.1.3 Elastic-plastic model

801 In FRP composites, the physical nonlinearity can also be considered a plastic behavior.
802 Sun and Chen [179] have proposed a 3D elastic-plastic model to represent this behavior.
803 In this model, effective stress and plastic strain increment were used to define the
804 nonlinear stress-strain relationship for the FRP composite, and a one-parameter flow
805 rule was developed to describe the elastic-plastic behavior. It should be noted that, since
806 FRP composites behave linearly up to failure in the fiber direction, the plasticity in the
807 fiber direction is not considered in the elastic-plastic model. In the FE approach, this is a
808 simple and accurate model to incorporate into CDM [180]. This elastic-plastic model,
809 coupled with the intralaminar damage criteria, has been employed to simulate the
810 nonlinear behavior of FRP composites under the impact load. The elastic-plastic model
811 may perform the same as the elastic model under lower energy impacts, while the
812 elastic model cannot provide a valid prediction of the sudden load drop, plastic
813 deformation, and fiber damage under higher energy impacts as compared with the
814 elastic-plastic model [181]. In the case of graphite/epoxy composites, a previous study
815 [182] also shows a better prediction for the damage-induced plastic model than the
816 damage model that does not consider the dependence of plasticity. Recently, Liu et al.
817 [183] improved the elastic-plastic model by considering the plastic damage in the
818 tensile, compressive, and shear directions and employed the Northwestern University
819 (NU) failure criterion [184-186] to predict the initiation of intralaminar damage. The
820 elastic-plastic model provides a closer prediction of the localization and the degree of

21
821 the permanent indentation in contrast to the same model, which only considered the
822 elastic behavior.

823 6.2 Interlaminar damage prediction

824 The interlaminar damage denotes the delamination failure between neighboring layers.
825 This failure mode can be modeled using the cohesive zone method, which can describe
826 the crack propagation in mode I, mode II, and mixed-mode failure. For the modeling of
827 FRP composites under the impact load, a bilinear cohesive mixed-mode damage model
828 [187] was commonly employed in the FE simulation. The cohesive element will be
829 inserted between neighboring layers to define the weak location (interface) where the
830 crack tends to occur. The quadratic stress criterion [188] was generally adopted for the
831 prediction of the damage onset between layers under impact load as follows:

832 tn
2
ts2 tt2
  1 (4)
N2 S2 T 2

833 where tn , ts , and tt mean the normal and shear tractions; and N , S , and T are

834 the normal and shear strengths. The B-K criterion [189] can be used to predict the
835 damage propagation between layers as follows:

836 
G 
G  G  (G  G )  S 
C C
n
C
s
C
n (5)
 GT 

837 where G C , GnC , and GsC are the total, normal, and shear fracture toughness,

838 respectively; GS means the dissipated energy in the out-of-plane direction; GT

839 denotes the total dissipated energy in all directions;  is an empirical parameter.

840 7. Challenges

841 The current study on the impact performance of the FRP cable is in its early stages.
842 Further research using experimental or numerical methods is required to understand the
843 impact performance of FRP cables, but the following challenges may arise:

844 (1) High manufacturing and testing costs

845 The cost of cable fabrication is directly related to the tonnage of the cable. The raw
846 materials and processing of large-tonnage FRP cables would be more expensive than

22
847 the smaller ones. On the other hand, the cost of the impact equipment and fixing system
848 also corresponds to the tonnage of the cable in testing. The scaled specimen is necessary
849 for future experimental studies, but it should conform to the configuration of the FRP
850 cables in practical applications.

851 (2) Property sacrifice in material design

852 Material toughening through fiber hybridization and matrix modification is an effective
853 way to enhance the impact resistance of FRP composites. However, the drawback is that
854 it may consume the original properties of a single fiber or neat matrix. For instance,
855 fiber hybridization may reduce tensile strength, tensile stiffness, fatigue performance,
856 and durability. The properties of hybrid fiber composites need to be assessed according
857 to the requirements of FRP cables. Matrix toughening may increase resin viscosity,
858 making fiber-resin impregnation more difficult. It is necessary to select appropriate
859 modifiers or fabrication methods to control the resin below the critical viscosity for
860 pultrusion.

861 (3) Lower computational efficiency

862 The numerical simulation is an effective way to reduce the cost of a full-scaled test for
863 the FRP cable with a large tonnage (e.g., more than 10 MN) and can conduct a
864 parametric study by changing the boundary and loading conditions. However, the
865 meso-scale finite element model with a large number of elements is time-consuming to
866 compute. An accurate numerical model with high computational efficiency is a
867 challenge for the simulation of the large-tonnage cable under the impact load. The
868 simplification of the material model and/or the geometry model is necessary for the
869 computation.

870 (4) Other challenges

871 In the structural application, analytical models or semi-empirical expressions would be


872 easier to use for engineers and designers and would be necessary for designing the FRP
873 cables. Several other challenges, including 1) anisotropic behavior; 2) a variety of cable
874 forms; and 3) the absence of a uniform test method, may contribute to the difficulty in
875 the anti-collision design of FRP cables (e.g., proposing an analytical model, data
876 comparison, and regressive analysis).

877

878 8. Conclusions

879 The sensitivity of FRP composite to impact loads has raised concerns about the safety
880 of FRP cables subjected to vehicle collisions. In practical applications, the anti-collision

23
881 design is necessary for FRP cables. A comprehensive review of the impact
882 performances of FRP composites and cables is carried out. Suggestions and challenges
883 for future studies of FRP cables are discussed in the present article.

884 The impact energy and fiber or matrix types are identified as the most important factors
885 affecting the impact performance of FRP composites. Aside from these factors, the
886 impact behavior of FRP cables is particularly affected by the pretension load, cable
887 tonnage, cable forms, impact position, and anchorage system. The pretension load leads
888 to an increase in the transverse stiffness of the cable and makes the cable more
889 susceptible to damage due to the combination of the tension, bending, and shear loads.
890 FRP cables have different tonnages, forms, and anchorage systems. The anchorage zone
891 is regarded as the weakest position to resist the impact load since the transverse load can
892 cause high shear and compressive stresses in this zone.

893 Although FRP cables are sensitive to impact, their impact resistance can be designed
894 through material toughening and protective methods. The fiber and matrix with high
895 toughness can absorb more impact energy. Fiber hybridization, matrix toughening, and
896 fiber/matrix interface reinforcement can be considered to improve the impact
897 performance of FRP cables. Adding an outside layer of protection is regarded as one
898 possible strategy for dramatically enhancing the impact performance of FRP cables. A
899 high-density polyethylene (HDPE) protective layer can be easily applied to the FRP
900 cables during fabrication and shows a good energy absorption capacity. Additional
901 protective systems are also needed to protect the impact-prone region and the weak
902 position of FRP cables.

903 Numerous studies have been conducted on the simulation of the impact behavior of FRP
904 composites. The separated material models coupled with the damage evolution model
905 can accurately predict the failure of FRP composites under impact loads. This numerical
906 method is suitable for predicting the impact performances of FRP cables and has the
907 advantage of reducing the cost of the experiment. The selection of appropriate material
908 models, boundary conditions, element types, and interactions is essential to accurately
909 predicting the impact response of FRP cables and their anchorage systems.

910 At present, there is insufficient understanding of the impact behavior of FRP cables to
911 implement an effective anti-collision design. The following prominent problems need to
912 be studied for FRP cables in future work:

913 (1) The research on the affecting factors needs to be further investigated, including: 1)
914 impact energy; 2) pretension load; 3) cable tonnage; 4) cable form; 5) impact position;
915 and 6) anchorage system. More experiments are also needed to investigate the influence
916 of impact damage or impact energy on the residual properties of FRP cables.

24
917 (2) The material design and protective method for improving the transverse impact
918 performance of FRP cables need to be investigated.

919 (3) Numerical models with high computational efficiency, analytical models, or
920 semi-empirical expressions need to be proposed to predict the impact responses of FRP
921 cables.

922 (4) To guide the anti-collision design of FRP cables in the future, a reliable standard or
923 method based on numerous experimental and theoretical results is needed.

924

925

926 Acknowledgements

927 This work was financially supported by the National Key Research and Development
928 Program of China (2021YFB3704403), and the NSFC with Grant No. 51878223.

929

930 Appendix A. Intralaminar Damage Prediction of FRP Composites

931 A.1 Fiber/matrix damage prediction

932 A.1.1 Damage initiation criteria

933 (1) Hashin criterion [40]

934 Fiber tension failure (  11  0 )

2 2 2
     
935 Fft   11    12    13   1 (A-1)
 X T   S12   S13 

936 Fiber compression failure (  11  0 )

2
  
937 Ffc   11   1 (A-2)
 XC 

938 Matrix tension failure (  22   33  0 )

25
2 2 2
    33  1   12    13 
Fmt   22   2  23   22 33       1
2
939 (A-3)
 YT  S 23  S12   S13 

940 Matrix compression failure (  22   33  0 )

2
    33   22   33  Y  2  1 2
  12    13 
2

Fmc   22      1  2  23   22 33       1
C 2
941
 2 S 23  YC  2 S 23   S 23  S12   S13 
942 (A-4)

943 where  ij ( i, j  1, 2, 3 ) is the effective stress tensor (Fig. A.1 (a)); X T and X C are

944 the fiber tension and compression strengths, respectively; YT and YC are the matrix

945 tensile and compressive strengths, respectively; and S12 , S13 , and S 23 are the shear

946 strengths of the 1-2, 1-3, and 2-3 planes, respectively.

947 (2) Hou criterion [41, 42]

948 Fiber tension failure (  11  0 )

2 2 2
     
949 Fft   11    12    13   1 (A-5)
 X T   S12   S13 

950 Fiber compression failure (  11  0 )

2 2 2
        
951 Ffc   11    12    13   1 (A-6)
 X c   S12   S13 

952 Matrix tension failure (  22  0 )

2 2 2
     
953 Fmt   22    12    23   1 (A-7)
 YT   S12   S 23 

26
954 Matrix compression failure (  22  0 )

2 2
1    Y 2   
955 Fmc   22   C 2 22  22   12   1 (A-8)
4  S12  4 S12YC YC  S12 

956 (3) Chang-Chang criterion [43]

957 Matrix tension failure (  22  0 )

2 2
   
958 Fmt   22    12   1 (A-9)
 YT   S12 

959 Matrix compression failure (  22  0 )

2 2
    Y  
960 Fmc   22    12   C 222  22  1 (A-10)
 S12   S12  4 S12 YC

961 where fiber damage initiation can be calculated according to Eqs. (A-1-2).

962 (4) Puck & Schürmann criterion [137]

963 Puck & Schürmann proposed that the fracture was caused by the stress acting on a
964 fracture plane with inclination angle  , as shown in Fig. A.1 (b). The inter-fiber
965 damage is triggered by the normal and shear stresses on the fracture plane. The effective
966 stress tensor and the inclination angle can be used to determine these stresses. For the
967 unidirectional composite under uniaxial transverse compressive load,  is
968 approximately 53 ° [130]. The fiber tensile/compressive damage initiation can be
969 calculated using the Hashin criterion, and the inter-fiber failure criterion was defined as

970 Inter-fiber tension (  n    0 )

2 2 2
  ( )    nt ( )    nl ( ) 
971 Fmt ( )   n   A    (A-11)
 YT   S    S12 

972 Inter-fiber compression (  n    0 )

27
2 2
  ( )    nl ( ) 
973 Fmc ( )   A nt    (A-12)
 S   nt n ( )   S12  nl n ( ) 

974 with

 n     22 cos 2    33 sin 2   2 23 cos  sin 



 nt    ( 22   33 ) cos  sin   2 23 (2 cos   1)
2


975  nl     13 sin    12 cos  (A-13)

 nt nt A YC  1  sin  
 nt  tan(2  90 ),S A  S ,S 23  2  cos   ,   2  90
o o

 23 12  

976 where SA is the transverse shear strength in the fracture plane; nt and nl are the

977 friction coefficients according to the Mohr–Coulomb failure theory.

978 A.1.2 Damage evolution model

979 The damage variable for each failure mode I can be calculated as follows:

980 a) Equivalent displacement method:

 If,eq  I ,eq   I0,eq 


981 dI  (d I   0,1 , I =ft , fc, mt , mc) (A-14)
 I ,eq  If,eq   I0,eq 

982 with  I0,eq   I ,eq FI ,  If,eq  2GI  I0,eq , and  I0,eq   I ,eq FI ; where  I0,eq is the

983 damage initiation equivalent displacement;  If,eq means the final equivalent

984 displacement;  I0,eq is the equivalent stress at the instant of damage initiation; GI is

985 the fracture toughness for each failure mode; FI is the value of the failure criterion for

986 each failure mode;  I ,eq and  I ,eq can be calculated according to Table A.1, which

987 are the equivalent displacement and equivalent stress for a failure mode, respectively.

988 b) Equivalent strain method:

28
 If,eq   I ,eq   I0,eq 
989 dI  (d I   0,1 , I =ft , fc, mt , mc) (A-15)
 I ,eq   If,eq   I0,eq 

990 with  I0,eq  X I E0, J ,  If,eq  2GI  X I lc  ; where  I0,eq and  If,eq mean the initial and

991 final equivalent strains, respectively; XI is the fiber or matrix strength; E0,J

992 represents the initial modulus of the composite lamina in the corresponding direction;
993 lc is the characteristic length of the element; The equivalent strain  I ,eq can be

994 calculated according to Table A.1.

995 For the Puck failure criterion, the inter-fiber failure of the damage initiation equivalent
996 strain is recorded by the calculation once the damage initiation criterion is satisfied,
997 while the final equivalent strain can be calculated as follows:

2G2kc
998  f
 0 (A-16)
I , eq
 I ,eq lc

999 where G2kc means the fracture toughness corresponding to transverse tension and

1000 compression;  I0,eq is the damage initiation equivalent stress.

1001 A.1.3 Continuum damage model

1002 The degraded compliance matrix can be obtained by adding the damage variable for
1003 each failure mode as follows:

29
 1   
d E  21  31 
E22 E33
 f 11 
  12 1  
  32 
 E11 d m E22 E33 
  13  1 
   23 
1004 S d   E11 E22 E33  (A-17)
 1 
 
 d f d mG12 
 1 
 d f d mG23 
 1 
 d f d mG31 

1005 where d f and d m are the fiber and matrix damage variables, respectively. The

1006 degraded stiffness matrix can be written as:

( 12   23 31) d f E11


 d f E11 (1  d m 23 32 ) d f d m E11 ( 31 +d m 21 32) 
 d m E22(1- d f  13 31) d m E22( 32  d f  12 31) 
 
1 (1- d f d m 12 21)
E33 
1007 Cd   
 sym d f d m G12 
 d f d m G23 
 
 d f d m G31 

1008 (A-18)

1009 with

d f  (1  d ft )(1  d fc )

1010 d m  (1  S mt d mt )(1  S mc d mc ) (A-19)
   1  d d    d    d    2d d   
 f m 12 21 m 23 32 f 13 31 f m 21 32 13

1011 where d ft , d fc , d mt , and d mc are damage variables for the fiber tension, fiber

1012 compression, matrix tension, and matrix compression failure modes, respectively; S mt

1013 and S mc are the empirical parameters to control the shear stiffness loss caused by the

1014 matrix tension and compression failure modes.

1015 A.2 Non-linear shear damage model [130]

30
1016 A strain-based criterion was used to detect the damage initiation for the shear failure
1017 modes.

2
 
1018 F ( ij )   ij0   1
s
(A-20)
ij  
 ij 

1019 where  ij0 means the elastic strain limit for matrix cracking.

1020 The non-linear shear damage model was shown in Fig. 18, where a polynomial cubic
1021 stress-strain relationship was used to express the non-linear shear behavior:

1022  ij ( ij )  c1 ij3  c2 ij2  c3 ij (A-21)

1023 where c1 , c2 , and c3 are the parameters obtained from fitting the polynomial

1024 expression with experimental stress-strain curves. The total strain of the material can be
1025 separated into:

1026  ij   ije   ijd (A-22)

1027 with

1028  ijd   ijed   ijin (A-23)

1029  ijin   ij   ije   ijed (A-24)

1030 where  ijed and  ijin are the elastic-damage and inelastic-damage parts of the damage

1031 strain, respectively. The ordinary elastic strain  ije and the elastic damage strain  ijed

1032 are reversible during the unloading process. Therefore, the total reversible elastic strain
1033  ijE can be expressed as

1034  ijE   ije   ijed (A-25)

1035 with

31
 ij
1036  ije  (A-26)
Gij0

 ij dij
1037  ijed  (A-27)
Gij0 (1  dij )

1038 where Gij0 is the initial shear modulus of the material, and dij is the damage variable

1039 and can be calculated according to:

1  a ij if  ij0   ij   ijmax

1040 dij ( ij )    ijf   ijmax    ijf   ijmax 
     f   max
  max
 1     max
1   if  ij   ij   ij
max f

 1 a 1 a
 ijf   ijmax   ij  ij
ij ij
   ij ij   

1041 (A-28)

1042 where a is a material parameter that can be obtained from the cyclic
1043 loading-unloading tension shear testing using ± 45 ° specimen by measuring the
1044 gradual shear stiffness reduction;  ijmax and  ijf are the maximum strain and failure

1045 strain, respectively;  ijf can be calculated by the shear fracture energy and the shear

1046 strength.

32
1047 Reference

1048 [1] Xian G, Guo R, Li C, Wang Y. Mechanical performance evolution and life
1049 prediction of prestressed cfrp plate exposed to hygrothermal and freeze-thaw
1050 environments. Composite Structures. 2022;293:115719.

1051 [2] Li C, Guo R, Xian G, Li H. Effects of elevated temperature, hydraulic pressure and
1052 fatigue loading on the property evolution of a carbon/glass fiber hybrid rod. Polymer
1053 Testing. 2020;90:106761.

1054 [3] Miller TC, Chajes MJ, Mertz DR, Hastings JN. Strengthening of a steel bridge
1055 girder using CFRP plates. Journal of Bridge Engineering. 2001;6:514-22.

1056 [4] Guo R, Li C, Xian G. Water absorption and long-term thermal and mechanical
1057 properties of carbon/glass hybrid rod for bridge cable. Engineering Structures.
1058 2023;274:115176.

1059 [5] Teng JG, Yu T, Fernando D. Strengthening of steel structures with fiber-reinforced
1060 polymer composites. Journal of Constructional Steel Research. 2012;78:131-43.

1061 [6] Zhou P, Li C, Bai Y, Dong S, Xian G, Vedernikov A, et al. Durability study on the
1062 interlaminar shear behavior of glass-fibre reinforced polypropylene (GFRPP) bars for
1063 marine applications. Construction and Building Materials. 2022;349:128694.

1064 [7] Dong S, Zhou P, Guo R, Li C, Xian G. Durability study of glass fiber reinforced
1065 polypropylene sheet under simulated seawater sea sand concrete environment. Journal
1066 of Materials Research and Technology. 2022;20:1079-92.

1067 [8] Davies GAO, Hitchings D, Zhou G. Impact damage and residual strengths of woven
1068 fabric glass/polyester laminates. Composites Part A: Applied Science and
1069 Manufacturing. 1996;27:1147-56.

1070 [9] Maio L, Monaco E, Ricci F, Lecce L. Simulation of low velocity impact on
1071 composite laminates with progressive failure analysis. Composite Structures.
1072 2013;103:75-85.

1073 [10] Richardson M, Wisheart M. Review of low-velocity impact properties of


1074 composite materials. Composites Part A: Applied Science and Manufacturing.
1075 1996;27:1123-31.

1076 [11] Tehrani M, Boroujeni A, Hartman T, Haugh T, Case S, Al-Haik M. Mechanical


1077 characterization and impact damage assessment of a woven carbon fiber reinforced
1078 carbon nanotube–epoxy composite. Composites Science and Technology. 2013;75:42-8.

33
1079 [12] Khan SU, Kim J-K. Impact and delamination failure of multiscale carbon
1080 nanotube-fiber reinforced polymer composites: a review. International Journal of
1081 Aeronautical and Space Sciences. 2011;12:115-33.

1082 [13] Abdewi EF, Sulaiman S, Hamouda A, Mahdi E. Quasi-static axial and lateral
1083 crushing of radial corrugated composite tubes. Thin-Walled Structures. 2008;46:320-32.

1084 [14] Elgalai AM, Mahdi E, Hamouda A, Sahari B. Crushing response of composite
1085 corrugated tubes to quasi-static axial loading. Composite Structures. 2004;66:665-71.

1086 [15] Mahdi E, Hamouda A, Mokhtar A, Majid D. Many aspects to improve damage
1087 tolerance of collapsible composite energy absorber devices. Composite Structures.
1088 2005;67:175-87.

1089 [16] Mahdi E, Sebaey T. Crushing behavior of hybrid hexagonal/octagonal cellular


1090 composite system: Aramid/carbon hybrid composite. Materials & Design.
1091 2014;63:6-13.

1092 [17] Agrawal S, Singh KK, Sarkar P. Impact damage on fibre-reinforced polymer
1093 matrix composite–a review. Journal of composite materials. 2014;48:317-32.

1094 [18] Koricho EG, Khomenko A, Haq M, Drzal LT, Belingardi G, Martorana B. Effect
1095 of hybrid (micro- and nano-) fillers on impact response of GFRP composite. Composite
1096 Structures. 2015;134:789-98.

1097 [19] Subadra SP, Griskevicius P, Yousef S. Low velocity impact and pseudo-ductile
1098 behaviour of carbon/glass/epoxy and carbon/glass/PMMA hybrid composite laminates
1099 for aircraft application at service temperature. Polymer Testing. 2020;89:106711.

1100 [20] Gokuldass R, Ramesh R. Mechanical and low velocity impact behaviour of
1101 intra-ply glass/kevlar fibre reinforced nano-silica and micro-rubber modified epoxy
1102 resin hybrid composite. Materials Research Express. 2019;6:055302.

1103 [21] Kinvi-Dossou G, Matadi Boumbimba R, Bonfoh N, Garzon-Hernandez S,


1104 Garcia-Gonzalez D, Gerard P, et al. Innovative acrylic thermoplastic composites versus
1105 conventional composites: Improving the impact performances. Composite Structures.
1106 2019;217:1-13.

1107 [22] Hosur M, Adbullah M, Jeelani S. Studies on the low-velocity impact response of
1108 woven hybrid composites. Composite Structures. 2005;67:253-62.

1109 [23] De Rosa IM, Santulli C, Sarasini F, Valente M. Post-impact damage


1110 characterization of hybrid configurations of jute/glass polyester laminates using acoustic
1111 emission and IR thermography. Composites Science and Technology. 2009;69:1142-50.

34
1112 [24] Sarasini F, Tirillò J, Valente M, Valente T, Cioffi S, Iannace S, et al. Effect of
1113 basalt fiber hybridization on the impact behavior under low impact velocity of
1114 glass/basalt woven fabric/epoxy resin composites. Composites Part A: Applied Science
1115 and Manufacturing. 2013;47:109-23.

1116 [25] Kazemi ME, Shanmugam L, Dadashi A, Shakouri M, Lu D, Du Z, et al.


1117 Investigating the roles of fiber, resin, and stacking sequence on the low-velocity impact
1118 response of novel hybrid thermoplastic composites. Composites Part B: Engineering.
1119 2021;207:108554.

1120 [26] Sonnenfeld C, Mendil-Jakani H, Agogué R, Nunez P, Beauchêne P.


1121 Thermoplastic/thermoset multilayer composites: A way to improve the impact damage
1122 tolerance of thermosetting resin matrix composites. Composite Structures.
1123 2017;171:298-305.

1124 [27] Vieille B, Casado VM, Bouvet C. Influence of matrix toughness and ductility on
1125 the compression-after-impact behavior of woven-ply thermoplastic- and
1126 thermosetting-composites: A comparative study. Composite Structures.
1127 2014;110:207-18.

1128 [28] Nash N, Young T, McGrail P, Stanley W. Inclusion of a thermoplastic phase to


1129 improve impact and post-impact performances of carbon fibre reinforced thermosetting
1130 composites—A review. Materials & Design. 2015;85:582-97.

1131 [29] Saghafi H, Fotouhi M, Minak G. Improvement of the impact properties of


1132 composite laminates by means of nano-modification of the matrix—A review. Applied
1133 Sciences. 2018;8:2406.

1134 [30] Yang Y, Fahmy MFM, Guan S, Pan Z, Zhan Y, Zhao T. Properties and
1135 applications of FRP cable on long-span cable-supported bridges: A review. Composites
1136 Part B: Engineering. 2020;190:107934.

1137 [31] Gallery: UPDATE: Lions Gate Bridge accident sends 2 to hospital.
1138 https://www.nsnews.com/local-news/update-lions-gate-bridge-accident-sends-2-to-hosp
1139 ital-2914006. 11/4/2012.

1140 [32] Semi-truck blown over on Mackinac Bridge; is that the beginning of our severe
1141 weather?
1142 https://www.mlive.com/weather/2013/07/semi-truck_blown_over_on_macki.html#incar
1143 t_river_default. 19/7/2013.

1144 [33] Traffic Control of Fuhe Bridge Section of Chengdu Ring Expressway Damaged by
1145 Freight Cars. http://scnews.newssc.org/system/20170705/000795347.html. 5/7/2017.

35
1146 [34] Review of the bus crash in Wanzhou, Chongqing.
1147 https://v.qq.com/x/page/f0778uw36jz.html. 2/11/2018.

1148 [35] Multiple lanes closed on Ravenel Bridge due to two car collision.
1149 https://www.counton2.com/news/local-news/multiple-lanes-closed-on-ravenel-bridge-d
1150 ue-to-two-car-collision/. 1/2/2020.

1151 [36] Multi-vehicle crash on Zakim Bridge leaves some with non-life-threatening
1152 injuries.
1153 https://www.bostonglobe.com/2020/12/30/metro/multi-vehicle-crash-zakim-bridge-leav
1154 es-some-with-non-life-threatening-injuries/. 30/12/2020.

1155 [37] Liu Y, Zwingmann B, Schlaich M. Carbon fiber reinforced polymer for cable
1156 structures—A review. Polymers. 2015;7:2078-99.

1157 [38] Wang A, Liu X, Yue Q. Low-velocity impact resistance of carbon fiber reinforced
1158 polymer composite and its cables: A review. Acta Materiae Compositae Sinica.
1159 2022;39:5049-61.

1160 [39] Zhang C, Duodu EA, Gu J. Finite element modeling of damage development in
1161 cross-ply composite laminates subjected to low velocity impact. Composite Structures.
1162 2017;173:219-27.

1163 [40] Hashin Z. Failure criteria for unidirectional fiber composites. 1980.

1164 [41] Hou JP, Petrinic N, Ruiz C. A delamination criterion for laminated composites
1165 under low-velocity impact. Composites Science and Technology. 2001;61:2069-74.

1166 [42] Hou JP, Petrinic N, Ruiz C, Hallett SR. Prediction of impact damage in composite
1167 plates. Composites Science and Technology. 2000;60:273-81.

1168 [43] Chang F-K, Chang K-Y. A progressive damage model for laminated composites
1169 containing stress concentrations. Journal of composite materials. 1987;21:834-55.

1170 [44] Zhang J, Zhang X. An efficient approach for predicting low-velocity impact force
1171 and damage in composite laminates. Composite Structures. 2015;130:85-94.

1172 [45] Shi Y, Swait T, Soutis C. Modelling damage evolution in composite laminates
1173 subjected to low velocity impact. Composite Structures. 2012;94:2902-13.

1174 [46] Wang A. Effect of flax surface modification and carbon fiber hybrid on the
1175 properties of flax reinforced polymer composites. Harbin: Harbin Institute of
1176 Technology, 2021.

36
1177 [47] ASTM. Standard, Standard test method for measuring the damage resistance of a
1178 fiber-reinforced polymer matrix composite to a drop-weight impact event. ASTM
1179 International West Conshohocken (PA); 2012.

1180 [48] Bozkurt ÖY, Erkliğ A, Bulut M. Hybridization effects on charpy impact behavior
1181 of basalt/aramid fiber reinforced hybrid composite laminates. Polymer Composites.
1182 2018;39:467-75.

1183 [49] Caminero M, Rodríguez G, Muñoz V. Effect of stacking sequence on Charpy


1184 impact and flexural damage behavior of composite laminates. Composite Structures.
1185 2016;136:345-57.

1186 [50] Kalthoff J. Characterization of the dynamic failure behaviour of a


1187 glass-fiber/vinyl-ester at different temperatures by means of instrumented Charpy
1188 impact testing. Composites Part B: Engineering. 2004;35:657-63.

1189 [51] ASTM D6110-18: standard test method for determining the charpy impact
1190 resistance of notched specimens of plastics. ASTM International West Conshohocken;
1191 2018.

1192 [52] PN-EN ISO 179-1: Plastics—Determination of Charpy Impact Properties—Part 1:


1193 Non-Instrumented Impact Test. Polish Committee for Standardization Warszawa,
1194 Poland; 2010.

1195 [53] ASTM D256: Standard test methods for determining the Izod pendulum impact
1196 resistance of plastics. ASTM West Conshohocken; 2010.

1197 [54] Nemat-Nasser S. Introduction to high strain rate testing. ASM handbook.
1198 2000;8:427-8.

1199 [55] Gama BA, Lopatnikov SL, Gillespie Jr JW. Hopkinson bar experimental
1200 technique: a critical review. Appl Mech Rev. 2004;57:223-50.

1201 [56] Gómez-del Río T, Barbero E, Zaera R, Navarro C. Dynamic tensile behaviour at
1202 low temperature of CFRP using a split Hopkinson pressure bar. Composites Science and
1203 Technology. 2005;65:61-71.

1204 [57] Liu L, Zhao Z, Chen W, Shuang C, Luo G. An experimental investigation on high
1205 velocity impact behavior of hygrothermal aged CFRP composites. Composite
1206 Structures. 2018;204:645-57.

1207 [58] Cantwell WJ, Morton J. The impact resistance of composite materials — a review.
1208 Composites. 1991;22:347-62.

37
1209 [59] Caprino G, Lopresto V, Santoro D. Ballistic impact behaviour of stitched
1210 graphite/epoxy laminates. Composites Science and Technology. 2007;67:325-35.

1211 [60] Andrew JJ, Srinivasan SM, Arockiarajan A, Dhakal HN. Parameters influencing
1212 the impact response of fiber-reinforced polymer matrix composite materials: A critical
1213 review. Composite Structures. 2019;224.

1214 [61] Grytten F. Low-velocity penetration of aluminium plates. 2008.

1215 [62] Olsson R. Mass criterion for wave controlled impact response of composite plates.
1216 Composites Part A: Applied Science and Manufacturing. 2000;31:879-87.

1217 [63] Olsson R. Closed form prediction of peak load and delamination onset under small
1218 mass impact. Composite Structures. 2003;59:341-9.

1219 [64] Greenhalgh E, Hiley M. The assessment of novel materials and processes for the
1220 impact tolerant design of stiffened composite aerospace structures. Composites Part A:
1221 Applied Science and Manufacturing. 2003;34:151-61.

1222 [65] Andrew JJ, Srinivasan SM, Arockiarajan A. Influence of patch lay-up
1223 configuration and hybridization on low velocity impact and post-impact tensile response
1224 of repaired glass fiber reinforced plastic composites. Journal of composite materials.
1225 2019;53:3-17.

1226 [66] Zhou J, Liao B, Shi Y, Ning L, Zuo Y, Jia L. Experimental investigation of the
1227 double impact position effect on the mechanical behavior of low-velocity impact in
1228 CFRP laminates. Composites Part B: Engineering. 2020;193:108020.

1229 [67] Shah S, Karuppanan S, Megat-Yusoff P, Sajid Z. Impact resistance and damage
1230 tolerance of fiber reinforced composites: A review. Composite Structures.
1231 2019;217:100-21.

1232 [68] Zhou J, Liao B, Shi Y, Zuo Y, Tuo H, Jia L. Low-velocity impact behavior and
1233 residual tensile strength of CFRP laminates. Composites Part B: Engineering.
1234 2019;161:300-13.

1235 [69] Atas C, Sayman O. An overall view on impact response of woven fabric composite
1236 plates. Composite Structures. 2008;82:336-45.

1237 [70] Mallick PK. Fiber-reinforced composites: materials, manufacturing, and design:
1238 CRC press, 2007.

1239 [71] kumre A, Rana RS, Purohit R. A Review on mechanical property of sisal glass
1240 fiber reinforced polymer composites. Materials Today: Proceedings. 2017;4:3466-76.

38
1241 [72] Gore PM, Kandasubramanian B. Functionalized aramid fibers and composites for
1242 protective applications: a review. Industrial & Engineering Chemistry Research.
1243 2018;57:16537-63.

1244 [73] Dhand V, Mittal G, Rhee KY, Park S-J, Hui D. A short review on basalt fiber
1245 reinforced polymer composites. Composites Part B: Engineering. 2015;73:166-80.

1246 [74] Swolfs Y, Gorbatikh L, Verpoest I. Fibre hybridisation in polymer composites: A


1247 review. Composites Part A: Applied Science and Manufacturing. 2014;67:181-200.

1248 [75] Nisini E, Santulli C, Liverani A. Mechanical and impact characterization of hybrid
1249 composite laminates with carbon, basalt and flax fibres. Composites Part B:
1250 Engineering. 2017;127:92-9.

1251 [76] Chen D, Luo Q, Meng M, Li Q, Sun G. Low velocity impact behavior of interlayer
1252 hybrid composite laminates with carbon/glass/basalt fibres. Composites Part B:
1253 Engineering. 2019;176:107191.

1254 [77] Papa I, Boccarusso L, Langella A, Lopresto V. Carbon/glass hybrid composite


1255 laminates in vinylester resin: Bending and low velocity impact tests. Composite
1256 Structures. 2020;232:111571.

1257 [78] Tang Y, Ye L, Zhang Z, Friedrich K. Interlaminar fracture toughness and CAI
1258 strength of fibre-reinforced composites with nanoparticles–A review. Composites
1259 Science and Technology. 2013;86:26-37.

1260 [79] Wang A, Xia D, Xian G, Li H. Effect of nanoclay grafting onto flax fibers on the
1261 interfacial shear strength and mechanical properties of flax/epoxy composites. Polymer
1262 Composites. 2019;40:3482-92.

1263 [80] Vignesh V, Balaji A, Rabi BRM, Rajini N, Ayrilmis N, Karthikeyan M, et al.
1264 Cellulosic fiber based hybrid composites: A comparative investigation into their
1265 structurally influencing mechanical properties. Construction and Building Materials.
1266 2021;271:121587.

1267 [81] Wang A, Wang X, Xian G. The influence of stacking sequence on the low-velocity
1268 impact response and damping behavior of carbon and flax fabric reinforced hybrid
1269 composites. Polymer Testing. 2021;104:107384.

1270 [82] Petrucci R, Santulli C, Puglia D, Nisini E, Sarasini F, Tirillò J, et al. Impact and
1271 post-impact damage characterisation of hybrid composite laminates based on basalt
1272 fibres in combination with flax, hemp and glass fibres manufactured by vacuum
1273 infusion. Composites Part B: Engineering. 2015;69:507-15.

39
1274 [83] Bandaru AK, Patel S, Sachan Y, Alagirusamy R, Bhatnagar N, Ahmad S. Low
1275 velocity impact response of 3D angle-interlock Kevlar/basalt reinforced polypropylene
1276 composites. Materials & Design. 2016;105:323-32.

1277 [84] Wang A, Wang X, Xian G. Mechanical, low-velocity impact, and hydrothermal
1278 aging properties of flax/carbon hybrid composite plates. Polymer Testing.
1279 2020;90:106759.

1280 [85] El Hachem Z, Célino A, Challita G, Moya M-J, Fréour S. Hygroscopic multi-scale
1281 behavior of polypropylene matrix reinforced with flax fibers. Industrial Crops and
1282 Products. 2019;140:111634.

1283 [86] Sujon MAS, Habib MA, Abedin MZ. Experimental investigation of the mechanical
1284 and water absorption properties on fiber stacking sequence and orientation of
1285 jute/carbon epoxy hybrid composites. Journal of Materials Research and Technology.
1286 2020;9:10970-81.

1287 [87] Vaidya U, Chawla K. Processing of fibre reinforced thermoplastic composites.


1288 International Materials Reviews. 2008;53:185-218.

1289 [88] Monticeli FM, Neves RM, Ornaghi Jr HL, Almeida Jr JHS. A systematic review on
1290 high ‐ performance fiber ‐ reinforced 3D printed thermoset composites. Polymer
1291 Composites. 2021;42:3702-15.

1292 [89] Shah SZH, Megat-Yusoff PSM, Karuppanan S, Choudhry RS, Ahmad F, Sajid Z,
1293 et al. Performance comparison of resin-infused thermoplastic and thermoset 3D fabric
1294 composites under impact loading. International Journal of Mechanical Sciences.
1295 2021;189:105984.

1296 [90] Shah SZH, Megat-Yusoff PSM, Karuppanan S, Choudhry RS, Ud Din I, Othman
1297 AR, et al. Compression and buckling after impact response of resin-infused
1298 thermoplastic and thermoset 3D woven composites. Composites Part B: Engineering.
1299 2021;207:108592.

1300 [91] Vieille B, Casado VM, Bouvet C. About the impact behavior of woven-ply carbon
1301 fiber-reinforced thermoplastic-and thermosetting-composites: a comparative study.
1302 Composite Structures. 2013;101:9-21.

1303 [92] Kazemi ME, Shanmugam L, Li Z, Ma R, Yang L, Yang J. Low-velocity impact


1304 behaviors of a fully thermoplastic composite laminate fabricated with an innovative
1305 acrylic resin. Composite Structures. 2020;250:112604.

1306 [93] Beland S. High performance thermoplastic resins and their composites: William

40
1307 Andrew, 1990.

1308 [94] Friedrich K, Gogeva T, Fakirov S. Thermoplastic impregnated fiber bundles:


1309 Manufacturing of laminates and fracture mechanics characterization. Composites
1310 Science and Technology. 1988;33:97-120.

1311 [95] Tzounis L, Kirsten M, Simon F, Mäder E, Stamm M. The interphase


1312 microstructure and electrical properties of glass fibers covalently and non-covalently
1313 bonded with multiwall carbon nanotubes. Carbon. 2014;73:310-24.

1314 [96] Xin W, Sarasini F, Tirillò J, Bavasso I, Sbardella F, Lampani L, et al. Impact and
1315 post-impact properties of multiscale carbon fiber composites interleaved with carbon
1316 nanotube sheets. Composites Part B: Engineering. 2020;183:107711.

1317 [97] Sharma VK, McDonald TJ, Kim H, Garg VK. Magnetic graphene–carbon
1318 nanotube iron nanocomposites as adsorbents and antibacterial agents for water
1319 purification. Advances in Colloid and Interface Science. 2015;225:229-40.

1320 [98] Yokozeki T, Iwahori Y, Ishiwata S, Enomoto K. Mechanical properties of CFRP


1321 laminates manufactured from unidirectional prepregs using CSCNT-dispersed epoxy.
1322 Composites Part A: Applied Science and Manufacturing. 2007;38:2121-30.

1323 [99] Ashrafi B, Guan J, Mirjalili V, Zhang Y, Chun L, Hubert P, et al. Enhancement of
1324 mechanical performance of epoxy/carbon fiber laminate composites using single-walled
1325 carbon nanotubes. Composites Science and Technology. 2011;71:1569-78.

1326 [100] Landowski M, Strugała G, Budzik M, Imielińska K. Impact damage in SiO2


1327 nanoparticle enhanced epoxy–Carbon fibre composites. Composites Part B:
1328 Engineering. 2017;113:91-9.

1329 [101] Ravindran AR, Ladani RB, Kinloch AJ, Wang C-H, Mouritz AP. Improving the
1330 delamination resistance and impact damage tolerance of carbon fibre-epoxy composites
1331 using multi-scale fibre toughening. Composites Part A: Applied Science and
1332 Manufacturing. 2021;150:106624.

1333 [102] Kostopoulos V, Baltopoulos A, Karapappas P, Vavouliotis A, Paipetis A. Impact


1334 and after-impact properties of carbon fibre reinforced composites enhanced with
1335 multi-wall carbon nanotubes. Composites Science and Technology. 2010;70:553-63.

1336 [103] Li J, Bai T. The effect of CNT modification on the mechanical properties of
1337 polyimide composites with and without MoS 2. Mechanics of Composite Materials.
1338 2012;47:597-602.

1339 [104] Varley RJ. Toughening of epoxy resin systems using low ‐ viscosity additives.

41
1340 polymer International. 2004;53:78-84.

1341 [105] Kinloch A, Masania K, Taylor A, Sprenger S, Egan D. The fracture of


1342 glass-fibre-reinforced epoxy composites using nanoparticle-modified matrices. Journal
1343 of Materials Science. 2008;43:1151-4.

1344 [106] DeCarli M, Kozielski K, Tian W, Varley R. Toughening of a carbon fibre


1345 reinforced epoxy anhydride composite using an epoxy terminated hyperbranched
1346 modifier. Composites Science and Technology. 2005;65:2156-66.

1347 [107] Akangah P, Lingaiah S, Shivakumar K. Effect of Nylon-66 nano-fiber


1348 interleaving on impact damage resistance of epoxy/carbon fiber composite laminates.
1349 Composite Structures. 2010;92:1432-9.

1350 [108] Wu G, Ma L, Liu L, Wang Y, Xie F, Zhong Z, et al. Interfacially reinforced


1351 methylphenylsilicone resin composites by chemically grafting multiwall carbon
1352 nanotubes onto carbon fibers. Composites Part B: Engineering. 2015;82:50-8.

1353 [109] Zhao M, Meng L, Ma L, Ma L, Yang X, Huang Y, et al. Layer-by-layer grafting


1354 CNTs onto carbon fibers surface for enhancing the interfacial properties of epoxy resin
1355 composites. Composites Science and Technology. 2018;154:28-36.

1356 [110] Xiang Y, Fang Z, Fang YW. Single and multiple impact behavior of CFRP cables
1357 under pretension. Construction and Building Materials. 2017;140:521-33.

1358 [111] Fang Y, Fang Z, Jiang Z, Jiang R, Zhou X. Investigation on failure behavior of
1359 carbon fiber reinforced polymer wire subjected to combined tension and bending.
1360 Composite Structures. 2021;267:113927.

1361 [112] Xiang Y, Fang Z, Wang C, Zhang Y, Fang Y. Experimental investigations on


1362 impact behavior of CFRP cables under pretension. Journal of Composites for
1363 Construction. 2017;21:04016087.

1364 [113] Fang YW, Fang Z, Jiang RN, Xiang Y, Huang DB. Transverse Static and
1365 Low-Velocity Impact Behavior of CFRP Wires under Pretension. Journal of
1366 Composites for Construction. 2019;23.

1367 [114] Zhu D, Li Z, Guo S, Li S, Yi Y. Lateral impact resistance of BFRP tendon under
1368 different pretensions and impact energies. Acta Materiae Compositae Sinica.
1369 2022;39:371-80.

1370 [115] Li S, Guo S, Yi Y, Rahman MZ, Bai X, Shi C, et al. Transverse low-velocity
1371 impact performance of BFRP bars after exposure to the saline-alkaline environment.
1372 Construction and Building Materials. 2021;307:124650.

42
1373 [116] Meier U. Carbon fiber reinforced polymer cables: Why? Why not? What if?
1374 Arabian Journal for Science and Engineering. 2012;37:399-411.

1375 [117] Fang Y, Fang Z, Feng L, Xiang Y, Zhou X. Bond behavior of an ultra-high
1376 performance concrete-filled anchorage for carbon fiber-reinforced polymer tendons
1377 under static and impact loads. Engineering Structures. 2023;274:115128.

1378 [118] Hong B, Xian G, Li H. Comparative study of the durability behaviors of


1379 epoxy-and polyurethane-based CFRP plates subjected to the combined effects of
1380 sustained bending and water/seawater immersion. Polymers. 2017;9:603.

1381 [119] Hong B, Xian G, Wang Z. Durability study of pultruded carbon fiber reinforced
1382 polymer plates subjected to water immersion. Advances in Structural Engineering.
1383 2018;21:571-9.

1384 [120] Hong B, Xian G. Ageing of a thermosetting polyurethane and its pultruded carbon
1385 fiber plates subjected to seawater immersion. Construction and Building Materials.
1386 2018;165:514-22.

1387 [121] Fang Y, Fang Z, Xiang Y, Feng L, Zhou X. Charpy impact properties of
1388 uni-directional carbon fiber-reinforced polymer tendons with protective layers.
1389 Advances in Structural Engineering. 2022:13694332221119876.

1390 [122] National Standards of the People's Republic of China. Hot-extruded PE protection
1391 paralleled high strength wire cable for cable-stayed bridge. 2018;GB/T 18365-2018.

1392 [123] Vasile C, Pascu M. Practical guide to polyethylene: iSmithers Rapra Publishing,
1393 2005.

1394 [124] Sun X, Gao Z, Cao P, Zhou C, Ling Y, Wang X, et al. Fracture performance and
1395 numerical simulation of basalt fiber concrete using three-point bending test on notched
1396 beam. Construction and Building Materials. 2019;225:788-800.

1397 [125] González E, Maimí P, Camanho P, Turon A, Mayugo J. Simulation of


1398 drop-weight impact and compression after impact tests on composite laminates.
1399 Composite Structures. 2012;94:3364-78.

1400 [126] Feng D, Aymerich F. Damage prediction in composite sandwich panels subjected
1401 to low-velocity impact. Composites Part A: Applied Science and Manufacturing.
1402 2013;52:12-22.

1403 [127] Dlugosch M, Fritsch J, Lukaszewicz D, Hiermaier S. Experimental investigation


1404 and evaluation of numerical modeling approaches for hybrid-FRP-steel sections under
1405 impact loading for the application in automotive crash-structures. Composite Structures.

43
1406 2017;174:338-47.

1407 [128] Davies G, Olsson R. Impact on composite structures. The Aeronautical Journal.
1408 2004;108:541-63.

1409 [129] Matthews FL, Davies G, Hitchings D, Soutis C. Finite element modelling of
1410 composite materials and structures: Elsevier, 2000.

1411 [130] Donadon M, Iannucci L, Falzon BG, Hodgkinson J, de Almeida SF. A


1412 progressive failure model for composite laminates subjected to low velocity impact
1413 damage. Computers & Structures. 2008;86:1232-52.

1414 [131] Iannucci L, Ankersen J. An energy based damage model for thin laminated
1415 composites. Composites Science and Technology. 2006;66:934-51.

1416 [132] Zhou J, Wen P, Wang S. Finite element analysis of a modified progressive
1417 damage model for composite laminates under low-velocity impact. Composite
1418 Structures. 2019;225:111113.

1419 [133] Tsai SW, Wu EM. A general theory of strength for anisotropic materials. Journal
1420 of composite materials. 1971;5:58-80.

1421 [134] Hill R. A theory of the yielding and plastic flow of anisotropic metals.
1422 Proceedings of the Royal Society of London Series A Mathematical and Physical
1423 Sciences. 1948;193:281-97.

1424 [135] Hill R. The mathematical theory of plasticity: Oxford university press, 1998.

1425 [136] Cepero-Mejías F, Curiel-Sosa J, Blázquez A, Yu T, Kerrigan K, Phadnis V.


1426 Review of recent developments and induced damage assessment in the modelling of the
1427 machining of long fibre reinforced polymer composites. Composite Structures.
1428 2020;240:112006.

1429 [137] Puck A, Schürmann H. Failure analysis of FRP laminates by means of physically
1430 based phenomenological models. Composites Science and Technology.
1431 2002;62:1633-62.

1432 [138] Pinho ST, Dávila CG, Camanho PP, Iannucci L, Robinson P. Failure models and
1433 criteria for FRP under in-plane or three-dimensional stress states including shear
1434 non-linearity. 2005.

1435 [139] Kober M, Kühhorn A. Comparison of different failure criteria for fiber-reinforced
1436 plastics in terms of fracture curves for arbitrary stress combinations. Composites
1437 Science and Technology. 2012;72:1941-51.

44
1438 [140] Hinton M, Kaddour A. The background to the second world-wide failure exercise.
1439 Journal of composite materials. 2012;46:2283-94.

1440 [141] Kaddour A-S, Hinton MJ. Maturity of 3D failure criteria for fibre-reinforced
1441 composites: Comparison between theories and experiments: Part B of WWFE-II.
1442 Journal of composite materials. 2013;47:925-66.

1443 [142] Pinho S, Darvizeh R, Robinson P, Schuecker C, Camanho P. Material and


1444 structural response of polymer-matrix fibre-reinforced composites. Journal of composite
1445 materials. 2012;46:2313-41.

1446 [143] Carrere N, Laurin F, Maire J. Micromechanical-based hybrid mesoscopic 3D


1447 approach for non-linear progressive failure analysis of composite structures. Journal of
1448 composite materials. 2012;46:2389-415.

1449 [144] Cuntze R, Freund A. The predictive capability of failure mode concept-based
1450 strength criteria for multidirectional laminates. Composites Science and Technology.
1451 2004;64:343-77.

1452 [145] Zhang J, Zhang X. Simulating low-velocity impact induced delamination in


1453 composites by a quasi-static load model with surface-based cohesive contact.
1454 Composite Structures. 2015;125:51-7.

1455 [146] de Freitas M, Silva A, Reis L. Numerical evaluation of failure mechanisms on


1456 composite specimens subjected to impact loading. Composites Part B: Engineering.
1457 2000;31:199-207.

1458 [147] Naik NK, Meduri S. Polymer-matrix composites subjected to low-velocity


1459 impact: effect of laminate configuration. Composites Science and Technology.
1460 2001;61:1429-36.

1461 [148] Menna C, Asprone D, Caprino G, Lopresto V, Prota A. Numerical simulation of


1462 impact tests on GFRP composite laminates. International Journal of Impact
1463 Engineering. 2011;38:677-85.

1464 [149] Singh H, Namala KK, Mahajan P. A damage evolution study of E-glass/epoxy
1465 composite under low velocity impact. Composites Part B: Engineering. 2015;76:235-48.

1466 [150] Liu P, Liao B, Jia L, Peng X. Finite element analysis of dynamic progressive
1467 failure of carbon fiber composite laminates under low velocity impact. Composite
1468 Structures. 2016;149:408-22.

1469 [151] Wang Z, Zhao J, Zhang X. Finite element analysis of composite laminates
1470 subjected to low-velocity impact based on multiple failure criteria. Materials Research

45
1471 Express. 2018;5:065320.

1472 [152] Shao J, Liu N, Zheng Z. Numerical comparison between Hashin and
1473 Chang-Chang failure criteria in terms of inter-laminar damage behavior of laminated
1474 composite. Materials Research Express. 2021;8:085602.

1475 [153] Li X, Ma D, Liu H, Tan W, Gong X, Zhang C, et al. Assessment of failure criteria
1476 and damage evolution methods for composite laminates under low-velocity impact.
1477 Composite Structures. 2019;207:727-39.

1478 [154] Lopes CS, Camanho PP, Gürdal Z, Maimí P, González EV. Low-velocity impact
1479 damage on dispersed stacking sequence laminates. Part II: Numerical simulations.
1480 Composites Science and Technology. 2009;69:937-47.

1481 [155] Zhou J, Wen P, Wang S. Numerical investigation on the repeated low-velocity
1482 impact behavior of composite laminates. Composites Part B: Engineering.
1483 2020;185:107771.

1484 [156] Miao H, Wu Z, Ying Z, Hu X. The numerical and experimental investigation on


1485 low-velocity impact response of composite panels: Effect of fabric architecture.
1486 Composite Structures. 2019;227:111343.

1487 [157] Zhang F, Liu K, Wan Y, Jin L, Gu B, Sun B. Experimental and numerical
1488 analyses of the mechanical behaviors of three-dimensional orthogonal woven
1489 composites under compressive loadings with different strain rates. International journal
1490 of damage mechanics. 2014;23:636-60.

1491 [158] Muñoz R, Seltzer R, Sket F, González C, Llorca J. Influence of hybridisation on


1492 energy absorption of 3D woven composites under low-velocity impact loading.
1493 Modelling and experimental validation. International Journal of Impact Engineering.
1494 2022;165:104229.

1495 [159] Davila CG, Camanho PP, Rose CA. Failure criteria for FRP laminates. Journal of
1496 composite materials. 2005;39:323-45.

1497 [160] Shah SZH, Megat-Yusoff PSM, Karuppanan S, Choudhry RS, Sajid Z. Multiscale
1498 damage modelling of 3D woven composites under static and impact loads. Composites
1499 Part A: Applied Science and Manufacturing. 2021;151:106659.

1500 [161] Liu H, Liu J, Ding Y, Zheng J, Kong X, Zhou J, et al. The behaviour of
1501 thermoplastic and thermoset carbon fibre composites subjected to low-velocity and
1502 high-velocity impact. Journal of Materials Science. 2020;55:15741-68.

1503 [162] Cui H, Thomson D, Eskandari S, Petrinic N. A critical study on impact damage

46
1504 simulation of IM7/8552 composite laminate plate. International Journal of Impact
1505 Engineering. 2019;127:100-9.

1506 [163] Khan AS, Huang S. Continuum theory of plasticity: John Wiley & Sons, 1995.

1507 [164] Luccioni B, Oller S, Danesi R. Coupled plastic-damaged model. Computer


1508 methods in applied mechanics and engineering. 1996;129:81-9.

1509 [165] Barbero EJ, Lonetti P. An inelastic damage model for fiber reinforced laminates.
1510 Journal of composite materials. 2002;36:941-62.

1511 [166] Ladeveze P, LeDantec E. Damage modelling of the elementary ply for laminated
1512 composites. Composites Science and Technology. 1992;43:257-67.

1513 [167] Van Paepegem W, De Baere I, Degrieck J. Modelling the nonlinear shear stress–
1514 strain response of glass fibre-reinforced composites. Part II: Model development and
1515 finite element simulations. Composites Science and Technology. 2006;66:1465-78.

1516 [168] Tan W, Falzon BG. Modelling the nonlinear behaviour and fracture process of
1517 AS4/PEKK thermoplastic composite under shear loading. Composites Science and
1518 Technology. 2016;126:60-77.

1519 [169] Chiu LNS, Falzon BG, Boman R, Chen B, Yan W. Finite element modelling of
1520 composite structures under crushing load. Composite Structures. 2015;131:215-28.

1521 [170] Tan W, Falzon BG, Chiu LNS, Price M. Predicting low velocity impact damage
1522 and Compression-After-Impact (CAI) behaviour of composite laminates. Composites
1523 Part A: Applied Science and Manufacturing. 2015;71:212-26.

1524 [171] Falzon BG, Apruzzese P. Numerical analysis of intralaminar failure mechanisms
1525 in composite structures. Part I: FE implementation. Composite Structures.
1526 2011;93:1039-46.

1527 [172] Faggiani A, Falzon BG. Predicting low-velocity impact damage on a stiffened
1528 composite panel. Composites Part A: Applied Science and Manufacturing.
1529 2010;41:737-49.

1530 [173] Shi Y, Pinna C, Soutis C. Modelling impact damage in composite laminates: A
1531 simulation of intra- and inter-laminar cracking. Composite Structures. 2014;114:10-9.

1532 [174] Berbinau P, Soutis C, Goutas P, Curtis PT. Effect of off-axis ply orientation on
1533 0°-fibre microbuckling. Composites Part A: Applied Science and Manufacturing.
1534 1999;30:1197-207.

47
1535 [175] Berbinau P, Soutis C, Guz IA. Compressive failure of 0° unidirectional
1536 carbon-fibre-reinforced plastic (CFRP) laminates by fibre microbuckling. Composites
1537 Science and Technology. 1999;59:1451-5.

1538 [176] Liu H, Falzon BG, Tan W. Experimental and numerical studies on the impact
1539 response of damage-tolerant hybrid unidirectional/woven carbon-fibre reinforced
1540 composite laminates. Composites Part B: Engineering. 2018;136:101-18.

1541 [177] Rezasefat M, Gonzalez-Jimenez A, Giglio M, Manes A. Numerical study on the


1542 dynamic progressive failure due to low-velocity repeated impacts in thin CFRP
1543 laminated composite plates. Thin-Walled Structures. 2021;167:108220.

1544 [178] Feng D, Aymerich F. Finite element modelling of damage induced by


1545 low-velocity impact on composite laminates. Composite Structures. 2014;108:161-71.

1546 [179] Sun C, Chen J. A simple flow rule for characterizing nonlinear behavior of fiber
1547 composites. Journal of composite materials. 1989;23:1009-20.

1548 [180] Chen JF, Morozov EV, Shankar K. A combined elastoplastic damage model for
1549 progressive failure analysis of composite materials and structures. Composite
1550 Structures. 2012;94:3478-89.

1551 [181] Liao BB, Liu PF. Finite element analysis of dynamic progressive failure of plastic
1552 composite laminates under low velocity impact. Composite Structures.
1553 2017;159:567-78.

1554 [182] Singh H, Mahajan P. Modeling damage induced plasticity for low velocity impact
1555 simulation of three dimensional fiber reinforced composite. Composite Structures.
1556 2015;131:290-303.

1557 [183] Liu H, Liu J, Ding Y, Hall ZE, Kong X, Zhou J, et al. A three-dimensional
1558 elastic-plastic damage model for predicting the impact behaviour of fibre-reinforced
1559 polymer-matrix composites. Composites Part B: Engineering. 2020;201:108389.

1560 [184] Daniel IM, Luo J-J, Schubel PM. Three-dimensional characterization of textile
1561 composites. Composites Part B: Engineering. 2008;39:13-9.

1562 [185] Daniel IM, Luo J-J, Schubel PM, Werner BT. Interfiber/interlaminar failure of
1563 composites under multi-axial states of stress. Composites Science and Technology.
1564 2009;69:764-71.

1565 [186] Daniel IM. Failure of Composite Materials. Strain. 2007;43:4-12.

1566 [187] Camanho PP, Dávila CG. Mixed-mode decohesion finite elements for the

48
1567 simulation of delamination in composite materials. 2002.

1568 [188] Brewer JC, Lagace PA. Quadratic stress criterion for initiation of delamination.
1569 Journal of composite materials. 1988;22:1141-55.

1570 [189] Benzeggagh ML, Kenane M. Measurement of mixed-mode delamination fracture


1571 toughness of unidirectional glass/epoxy composites with mixed-mode bending
1572 apparatus. Composites Science and Technology. 1996;56:439-49.

1573 [190] Lim AS, An Q, Chou T-W, Thostenson ET. Mechanical and electrical response of
1574 carbon nanotube-based fabric composites to Hopkinson bar loading. Composites
1575 Science and Technology. 2011;71:616-21.

1576 [191] Namala KK, Mahajan P, Bhatnagar N. Digital image correlation of low-velocity
1577 impact on a glass/epoxy composite. International Journal for Computational Methods in
1578 Engineering Science and Mechanics. 2014;15:203-17.

49
Fig. 1 Vehicle collisions on the cable-supported bridges in the last decade: (a) Lions
Gate Bridge, Canada [31]; (b) Mackinac Bridge, USA [32]; (c) Chengdu Fuhe Bridge,
China [33]; (d) The Second Yangtes River Bridge of Wanzhou, China [34]; (e) Ravenel
Bridge, USA [35]; (f) Zakim Bridge, USA [36].

50
Fig. 2 Drop-weight impact test instrument and the fixing system [46].

Fig. 3 Pendulum impact tests: (a) Charpy and (b) Izod impact tests [51, 53].

Fig. 4 Instruments of the high-strain rate impact test: (a) HSPB [190] and (b) gas gun
[57].

51
Fig. 5 Classification of different elastic waves: (a) dilatational waves; (b) flexural and
shear governed waves; (c) quasi-static waves [62].

(a) (b)

Fig. 6 Target responses under impact loads: (a) boundary-controlled impact; (b)
wave-controlled impact [63].

52
Fig. 7 Failure mechanisms of FRP composites under (a) low, (b) intermediate, and (c)
high energy impacts.

(a) (b)

Fig. 8 Typical mechanical responses of FRP composites under various energy impacts:
(a) load-time and load-displacement curves, (b) energy- and velocity-time curves.

53
Fig. 9 The three main hybrid patterns for FRP composite: (a) interlayer or
layer-by-layer, (b) intralayer or yarn-by-yarn, and (c) intrayarn or fiber-by-fiber.

Fig. 10 Interface system reinforcement methods: (a) CNT dispersion in matrix, (b) sheet
interleaving between fiber layers, and (c) grafting CNTs directly on the fibers.

54
Fig. 11 The mechanism of loading transfer in FRP composite after grafting [109].

Fig. 12 Different forms of FRP cables: (a) parallel wire cable, (b) twisted wire cable, (c)
plate cable, and (d) sheet cable [30].

Fig. 13 Anchorage systems: (a) gradient anchorage by filling bonding material, (b)
modular clamp anchorage [37].

55
Fig. 14 Impact experiment and the impact response of CFRP composite: (a) CFRP
strand with twisted wire form, (b) drop-weigh impact test, (c) impact response of CFRP
strand under different impact energy events, and (d) impact response of CFRP wire with
different pretension loads [110, 113].

Fig. 15 Intralaminar fracture considered in the damage initiation criteria.

56
Fig. 16 Ranking of the theories according to their scores of Grades A, B, and C (Grade
A: when the prediction lies within ±10% of the experimental value; Grade B: when
the prediction lies between ±10% and ±50% of the experimental value; Grade C:
when the prediction lies below ±50% and above ±150% of the experimental value)
[141].

Fig. 17 CDM methods using (a) linear and (b) exponential stiffness degradations [136].

57
Fig. 18 Nonlinear shear damage model [172].

(a) (b)

Fig. A.1 Schematic representation of the stress tensor in (a) the main axis and (b) the
inclination angle of the FRP composite cell.

58
Table A.1 Equivalent displacement/strain and stress of each failure mode.

Failure Equivalent displacement or strain Equivalent stress


modes

Fiber
lc   11 11   1212   1313 
2
 ft ,eq  lc 11  122  132
tension1  ft ,eq 
 ft ,eq

2
 ft ,eq  11  122  132

Fiber
 fc ,eq  lc 11 lc   11 11 
compression  fc ,eq 
 fc ,eq
1

 fc ,eq  11

Matrix 2 2
 mt ,eq 
lc (  22  22   33  33  1212   23 23  1313 )
 mt ,eq  lc  22   33  122   23
2
 132  mt ,eq
tension1

59
2 2
 mt ,eq   22   33  122   23
2
 132

Matrix 2 2
 mc ,eq 
lc (  22  22   33  33   1212   23 23   1313 )
 mc , eq  lc  22   33   122   23
2
  132  mc ,eq
compression
1
2 2
 mc ,eq   22   33  122   23
2
 132

Fiber
lc   11 11   1212   1313 
2
 ft ,eq  lc 11  122  132
tension2  ft ,eq 
 ft ,eq

2
 ft ,eq  11  122  132

Fiber
lc   11 11   1212   1313 
2
 fc ,eq  lc 11  122  132
compression  fc ,eq 
 ft ,eq
2

2
 fc ,eq  11  122  132

Matrix
lc   22  22   1212   23 23 
2
 mt ,eq  lc  22  122   23
2

tension2  mt ,eq 
 mt ,eq

2
 mt ,eq   22  122   23
2

Matrix
lc   22  22   1212 
2
 mc ,eq  lc  22  122
compression  mc ,eq 
 mc ,eq
2

2
 mc ,eq   22  122

Matrix
lc   22  22   1212 
2
 mt ,eq  lc  22  122
tension3  mt ,eq 
 mt ,eq

2
 mt ,eq   22  122

60
Matrix
lc   22  22   1212 
2
 mc ,eq  lc  22  122
compression  mc ,eq 
 mc ,eq
3

2
 mc ,eq   22  122

Matrix
 mt ,eq  lc  n2   nl2   nt2  mt ,eq   n2   nl2   nt2
tension4

 mt ,eq   n2   nl2   nt2

Matrix
 mc ,eq  lc  nl2   nt2  mc ,eq   nl2   nt2
compression
4

 mc ,eq   nl2   nt2

Note: Superscripts 1, 2, 3, and 4 correspond to the Hashin, Hou, Chang-Chang, and


Puck & Schürmann criteria, respectively; the symbol  is the Macaulay bracket

(   (   ) 2 ).

Table A.2 Properties of the unidirectional ply of the FRP composites.

E-glass/e
Paramete T700GC/M HS300/ET2 IM7/85 HS160/RE AS4/85 AS4/PEE poxy
rs 21 [153] 23 [178] 52 [162] M [145] 52 [154] K [183] [149,
191]

61
E11
130 122 162 93.7 135 127 40
(GPa)

E22, 33
7.7 6.2 9.7 7.45 9.6 10.3 10
(GPa)

G12, 13
4.8 4.4 4.7 3.97 5.3 5.7 3.15
(GPa)

G23
3.8 3.7 - 3.97 - 5.7 4.32
(GPa)

μ12,13 0.33 0.35 0.36 0.26 0.32 0.3 0.3

μ23 0.35 0.5 - 0.26 0.49 0.3 0.21

XT
2080 1850 2625 - 2207 2070 988
(MPa)

XC
1250 1470 1451 - 1531 1360 1432
(MPa)

YT (MPa) 60 29 60.2 30 80.7 85 44

YC (MPa) 140 290 400 - 199.8 276 285

S12, 13
110 65 112.7 80 114.5 186 60.6
(MPa)

S23
110 65 - 80 - 186 22
(MPa)

Gft
133 92 82 - 81.5 218 -
(N/mm)

62
Gfc
40 80 102 - 106.3 104 -
(N/mm)

Gmt
0.6 0.52 - - - 1.7 -
(N/mm)

Gmc
2.1 1.61 - - - 2.0 -
(N/mm)

Kn,s
(GPa/m 5 120n,43s - 120n,43s 9.6 640 -
m)

N (MPa) 30 30 - 30 80.7 43 -

S, T
30 80 - 80 114.5 50 -
(MPa)

GnC
0.6 0.52 0.22 0.52 0.28 1.7 -
(N/mm)

GsC
2.1 0.92 0.66 0.97 0.97 2.0 -
(N/mm)

Note: Kn,s means the cohesive stiffness in the normal and shear directions.

63
Declaration of interests

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.

64

You might also like