Vous êtes sur la page 1sur 53

Accepted Manuscript

Durability performance and long-term prediction models of sand-coated basalt FRP


bars

Ahmed H. Ali, Hamdy M. Mohamed, Brahim Benmokrane, Adel ElSafty, Omar


Chaallal

PII: S1359-8368(18)32267-4
DOI: 10.1016/j.compositesb.2018.08.065
Reference: JCOMB 5895

To appear in: Composites Part B

Received Date: 20 July 2018

Accepted Date: 20 August 2018

Please cite this article as: Ali AH, Mohamed HM, Benmokrane B, ElSafty A, Chaallal O, Durability
performance and long-term prediction models of sand-coated basalt FRP bars, Composites Part B
(2018), doi: 10.1016/j.compositesb.2018.08.065.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. 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.
ACCEPTED MANUSCRIPT

1 Durability Performance and Long-Term Prediction Models of Sand-Coated Basalt FRP

2 Bars

3 Ahmed H. Ali,1,3 Hamdy M. Mohamed,2,3 Brahim Benmokrane,2 Adel ElSafty,4 and Omar

4 Chaallal1

PT
5
1

RI
6 Department of Construction Engineering, École de Technologie Supérieure, Montreal, Canada.
2
7 Department of Civil Engineering, University of Sherbrooke, Sherbrooke, Quebec, Canada.

SC
3
8 Department of Civil Engineering, Helwn University, Cairo, Egypt.
4
9 Civil Engineering, School of Engineering, University of North Florida, Jacksonville, Florida,

10 USA.
U
AN
11 Adel.el-safty@unf.edu
M

12

13
D

14
TE

15

16
EP

17

18
C

19
AC

20

21

22

23

1
ACCEPTED MANUSCRIPT

24 ABSTRACT

25 Basalt-fiber-reinforced polymer (BFRP) bars are expected to provide benefits that are

26 comparable or superior to other types of FRP while being significantly cost-effective. However,

27 extensive investigations are needed to evaluate the long-term characteristics and durability

PT
28 performance of these bars. This article presents an experimental study that investigated the

RI
29 physical, mechanical, microstructural, and durability characteristics of newly developed basalt-

30 fiber-reinforced polymer (BFRP) bars. The physical, mechanical properties and microstructural

SC
31 characteristics were evaluated first on the unconditioned BFRP bars. The durability performance

32 of the BFRP bars was then assessed by conducting the mechanical tests, such as transverse-shear

33
U
test, flexural test, and interlaminar-shear test, on the specimens after different exposure periods
AN
34 (1,000; 3,000; and 5,000 h) at 60oC. Thereafter, the BFRP bar properties were assessed and

compared with the values obtained on the unconditioned specimens. The test parameter was
M

35

36 conditioning time (1,000; 3,000; and 5,000 h). The test results revealed that the unconditioned
D

37 BFRP bars had the best physical properties. On the other hand, the long-term durability
TE

38 performance revealed that the transverse shear-strength, flexural-strength, and interlaminar

39 shear-strength retention were decreased by 12%, 19%, and 21%, respectively.


EP

40

Keywords: Basalt-FRP; tensile strength; alkaline solution, flexural strength, transverse shear-
C

41
AC

42 strength, interlaminar shear-strength, durability.

43

44

45

46

2
ACCEPTED MANUSCRIPT

47 INTRODUCTION

48 The durability and life performance of reinforced concrete structures have become a major

49 concern for the infrastructure systems (Wang et al. 2016a; Ali et al. 2018a). One of the main

50 factors reducing durability and service life of reinforced concrete structures is the corrosion of

PT
51 embedded steel reinforcement bars, particularly where deicing salts are routinely used, such as in

RI
52 concrete deck slabs and parking garages (Wang et al. 2015; Ali et al. 2018b). Recently, fiber-

53 reinforced-polymer (FRP) materials have emerged as an acceptable construction material, for

SC
54 both new constructions and rehabilitation and strengthening of existing structures, because of

55 their noncorrodible nature, higher tensile strength, and lower weight relative to conventional

56
U
steel reinforcing bars (ACI 440.1R-15; Arias et al. 2012; Benmokrane et al. 2018, 2016a; Wang
AN
57 et al. 2015; Ali et al.2017a). FRPs are available with a wide range of mechanical properties

(tensile strength, bond strength, and modulus of elasticity) and are made with high-tensile-
M

58

59 strength fibers such as glass, carbon, and aramid embedded in polymer matrices as polyesters,
D

60 vinyl esters, or epoxies. Moreover, FRPs can be produced as bars, ropes, tendons, and grids in a
TE

61 wide variety of shapes and surface configurations as well as with varied characteristics (ACI

62 440.1R-15; Ali et al. 2017b).


EP

63 Recent developments in FRP technology and innovate, new types of fibers—such as basalt fibers

64 which is made from basalt rock —are being introduced to manufacture basalt fiber-reinforced
C

65 polymers (BFRPs). BFRP is the most recently FRP composite, appearing within the last decade
AC

66 (Wang et al. 2016b, 2014; Shen et al. 2016; Benmokrane et al. 2015). BFRPs are non-corrosive,

67 non-magnetic, posses high resistance against low and high temperatures, good chemical

68 resistance, good environmental friendliness, high performance in terms of strength, fire

69 resistance and durability, as well as the lower potential cost with respect to many other FRP

3
ACCEPTED MANUSCRIPT

70 materials. Consequently, the basalt fibers are ideally appropriate for applications involving high

71 temperature, chemical resistance, durability, mechanical strength, and low water absorption

72 (INFOMINE 2007). In comparison to others FRP types, BFRP fibers lie between glass and

73 carbon for both stiffness and strength. BFRP has higher strength and modulus, similar cost, and

PT
74 greater chemical stability compared with E-glass FRP. Furthermore, it offers over five times the

RI
75 strength and one-third the density of commonly used low-carbon steel bars (Arias et al. 2012;

76 Zhishen et al. 2012). In the past decade, numerous studies on the durability of glass and carbon

SC
77 FRP bars have been conducted which mainly comprises their performances under alkali, acid and

78 salt environments [Ng and Lee 2002; Tȁljsten et al. 2003; Davalos et al. 2011; Won and Park

79
U
(2006); Chen et al. 2006 and 2007; Karbhari et al. 2007; Robert et al. 2010; Davalos et al. 2012,
AN
80 Benmokrane et al. 2016 (a, b), 2012, and 2014]. The majority of these studies has highlighted on

the short-term performance of FRP bars. Moreover, most of these studies have investigated the
M

81

82 durability characteristics of glass FRP bars for reinforced concrete structures. Few studies,
D

83 however, have assessed the physical, mechanical and durability characteristics of basalt-FRP, in
TE

84 order to understand their behavior and generate higher confidence in these newly developed FRP

85 materials, under alkali and salt environments (Wang et al. 2016, 2014; Shen et al. 2016;;
EP

86 Benmokrane et al. 2015, Elgabbas et al. 2015, Zhishen et al. 2012). Sim et al. (2005)

87 investigated the applicability of the basalt, glass, and carbon fibers as a strengthening material
C

88 for structural concrete members through various experimental works for durability, mechanical
AC

89 properties, and flexural strengthening at elevated temperatures. They reported that immersing

90 basalt and glass fibers in an alkaline solution led to volume and strength loss through a surface

91 reaction, whereas carbon fibers did not show any significant strength reduction. The basalt fibers,

92 however, retained about 90% of their strength at ambient temperature after exposure at 600°C for

4
ACCEPTED MANUSCRIPT

93 2 h. Aging results indicate that the interfacial region in basalt composites might be more

94 vulnerable to environmental damage than in glass composites. The basalt-epoxy interface,

95 however, might also be more durable than the glass-epoxy interface in tension-tension fatigue,

96 because basalt composites have a longer fatigue life.

PT
97 An experimental study was conducted, by Parnas et al. 2007, to determine if BFRP composites

RI
98 were feasible, practical, and a beneficial alternative for transportation applications. The

99 researchers of the mentioned study concluded that the chemical composition of basalt fibers is

SC
100 close to glass fibers, except that the basalt contains a high ratio of iron oxide, conferring its

101 brown color. Aging results indicate that the interfacial region in basalt composites might be more

102
U
vulnerable to environmental damage than in glass composites. The basalt-epoxy interface,
AN
103 however, might also be more durable than the glass-epoxy interface in tension-tension fatigue,

because basalt composites have a longer fatigue life. In 2008, Wang et al. investigated the
M

104

105 chemical durability and mechanical properties of a kind of alkali-proof basalt fiber and its
D

106 reinforced epoxy resin matrix composites in alkaline solution for 3 months. The basalt fiber was
TE

107 boiled in distilled water, sodium hydroxide and hydrochloric acid, respectively. Then the mass

108 loss and strength change of the fibers were studied showing that the alkali resistance of the basalt
EP

109 fiber is better than acid resistance. The experimental results also showed that, after exposure, the

110 modulus of the BFRP was unaffected, but its strength decreased by 40%.
C

111 In the last two years, an extensive research project has been conducted at the Department of Civil
AC

112 Engineering at the University of Sherbrooke to assess the short-term and long-term

113 characteristics of newly developed BFRP bars as a preliminary step in introducing these new

114 materials into pilot projects and FRP design codes and material specification (Benmokrane et al.

115 2015; Elgabbas et al. 2015). Benmokrane et al. 2015 presented an experimental study that

5
ACCEPTED MANUSCRIPT

116 investigated the physical, mechanical, and durability characteristics of three different types of

117 fiber-reinforced polymer (FRP) bars made of basalt and glass fibers with vinylester and epoxy

118 resins. They found that the glass/vinylester composite had the best physical and mechanical

119 properties and lowest degradation rate after conditioning in alkaline solution. The basalt/epoxy

PT
120 composite ranked second, while the basalt/vinylester composite evidenced the lowest physical

RI
121 and mechanical properties and exhibited significant degradation of its physical and mechanical

122 properties after conditioning. Elgabbas et al. 2015 investigated the durability and long-term

SC
123 performance of BFRP bars in an alkaline solution (up to 3000 h, at 60oC) to determine their

124 suitability as internal reinforcement for concrete elements. It was noticed that the BFRP bars had

125
U
good mechanical behavior and could be placed in the same category as grade II and grade III
AN
126 GFRP bars (according to tensile modulus of elasticity).

This paper represents an experimental investigation to determine the physical, mechanical, and
M

127

128 durability characteristics of newly developed BFRP bars, manufactured by a Canadian company
D

129 (Pultrall Inc.), as corrosion-resistant reinforcing material for bridge deck, girder, and pile
TE

130 applications in marine and aggressive environment. Since BFRPs have not been included in

131 design standards and specifications yet, the results of this experimental study will contribute to
EP

132 integrating BFRP into FRP standards and guides.

133 EXPERIMENTAL PROGRAM OUTLINE


C

134 Material Properties


AC

135 Sand coated basalt FRP bars manufactured by a Canadian company (Pultrall Inc. 2015) were

136 used in this study as shown in Fig. 1. The BFRP bars manufactured from continuous basalt fibers

137 impregnated in vinylester resins using the pultrusion process. The nominal cross-sectional area

138 of the bar is 284 mm2, as reported by the manufacturer, while the actual (immersion) cross-

6
ACCEPTED MANUSCRIPT

139 sectional area is 362 mm2, as specified in CSA-S806-12 (2012). The mechanical properties

140 reported herein were calculated using the nominal cross-sectional area.

141 Test Specimens

142 For this study, tensile [ASTM D7205 (ASTM 2011)], transverse shear [ASTM D7617 (ASTM

PT
143 2011)], flexural (ASTM D4476 (ASTM 2009)], and short-beam shear testing [ASTM D4475

RI
144 (ASTM 2008)] were conducted as the appropriate mechanical tests. The BFRP bars were

145 provided into 2,400; 350; and 300 mm lengths for tensile strength, transverse shear strength, and

SC
146 flexural properties, respectively. In addition, some specimens were cut into 120 mm lengths so

147 that the short-beam shear test could be performed according to ASTM D4475 (ASTM 2008). For

148
U
tensile testing, Fig. 2b shows the dimensions of the typical tensile test specimens as specified in
AN
149 ASTM D7205.

Conditioning of BFRP Bar Specimens


M

150

151 The basalt-FRP specimens were completely immersed in alkaline solution inside stainless steel
D

152 containers specially manufactured for the study (Fig. 3). The alkaline solution was prepared
TE

153 using calcium hydroxide, potassium hydroxide, and sodium hydroxide (118.5 g of Ca(OH)2 +

154 0.9 g of NaOH + 4.2 g of KOH in 1 L of deionized water) according to CSA S806 and ACI
EP

155 440.3R. The pH of alkali solution was 12.8. It is worth noting that the alkaline environment in

156 concrete has a pH above 12 (ACI 440.4R-04). The BFRP specimens and stainless steel
C

157 containers were kept at 22°C and 60°C for 1,000; 3,000; and 5,000 h of conditioning. The
AC

158 stainless steel containers were covered with polyethylene sheeting to avoid water evaporation

159 during conditioning. Furthermore, the water level was kept constant throughout the study to

160 avoid a pH increase that could result from decreased water level and a significant increase of

161 alkaline ions in the solution. The immersion temperatures were chosen to accelerate the

7
ACCEPTED MANUSCRIPT

162 degradation effect of aging, but they were not high enough to produce any thermal-degradation

163 mechanisms. The BFRP specimens were removed from the alkaline solution after 1,000, 3,000,

164 and 5,000 h and tested under transverse shear, flexural, and short-beam shear to compare their

165 strength retention values to those of the reference specimens.

PT
166 PHYSICAL CHARACTERIZATION OF BFRP BARS

RI
167 The Physical and microstructural analyses were conducted on BFRP specimens. Physical

168 properties included: (1) cross section area (2) basalt-fiber content, (3) moisture absorption, (4)

SC
169 cure ratio, (5) glass-transition temperature, (6) transverse coefficient of thermal expansion, and

170 (7) wicking. These properties were determined according to CSA S807. Moreover, optical

171
U
microscopy (OM), and scanning electronic microscopy (SEM) analyses were performed to
AN
172 investigate the material’s microstructure.

Cross-Sectional Area
M

173

174 Three 280 mm long specimens were prepared and tested to determine the actual cross-sectional
D

175 area of the BFRP bars as specified in CSA-S806-12 (2012), Annex A. All BFRP specimens
TE

176 were kept in the test environment for 24 hours prior to weighing and measuring. The actual

177 cross-sectional area was calculated according to the following equation:


EP

178 Cross-sectional area = 1000⋅(V0 – V1)/L (1)

179 where, L is specimen length; V0 is the volume of water added without BFRP specimen; V1 is the
C

180 volume of water added with BFRP specimen.


AC

181 Fiber Content

182 Basalt fiber content was determined by thermogravimetry according to ASTM E1131. A very

183 small piece of material (a few tenths milligrams) was cut from the center of the bar, placed in

184 platinum crucible and then heated up to 550°C under inert atmosphere. The weight loss (WL) has

8
ACCEPTED MANUSCRIPT

185 been recorded at a temperature equal to 550°C. Since the material only contains basalt fibers and

186 resin, fiber content by weight was then calculated according to the following equation:

187 Fiber content by weight = 100⋅(WT – WL)/WT (2)

188 where, WT is the total weight before burn off.

PT
189 Water-Immersion Test

RI
190 The moisture uptake at saturation of BFRP bars was determined according to ASTM D570,

191 except that the immersions were performed in tap water instead of distilled water. Three 50-mm-

SC
192 long specimens were cut, dried, and weighed prior to immersion in water at 50°C for three

193 weeks. The samples were removed from the water after three weeks, surface dried, and weighed.

194
U
The water content at saturation in weight percent (Ws) was calculated using Equation 3.
AN
195 Ws = 100 · (Ps – Pd)/Pd (3)
M

196 where Ps and Pd are the sample weights in saturated and dry states, respectively.

197 Cure ratio


D

198 Cure ratio was determined according to ASTM D5028. The enthalpy of polymerization of the
TE

199 sample is measured by DSC and compared to the enthalpy of polymerization of pure resin,

200 taking into account the weight percentage of resin in the matrix. Thirty to fifty milligrams of
EP

201 sample were accurately weighed and placed in aluminum crucible. The samples were then heated

202 from room temperature to 200°C at a heating rate of 20°C/min and the area of the peak of
C

203 polymerization was calculated. The measurement was carried out on 3 specimens.
AC

204 Glass transition temperature Tg

205 Glass transition temperature, Tg, was determined for 12.5 mm BFRP specimens by Differential

206 Scanning Calorimetry (DSC) using ASTM E 1131 test method. Thirty to forty milligrams of

207 composite sample were weighed and placed in an aluminum pan. The sample was then heated up

9
ACCEPTED MANUSCRIPT

208 to 200°C under nitrogen at a heating rate of 20°C/min. The value of Tg was taken at the mid-

209 height of the heat capacity (Cp) jump.

210 MECHANICAL CHARACTERIZATION OF THE BFRP BARS

211 Tensile Test

PT
212 All BFRP bars were tested under tension according to the ASTM D7205. Each specimen was

RI
213 instrumented with a linear variable differential transformer (LVDT) to capture the elongation

214 during testing. The test was carried out using a Baldwin testing machine and the load was

SC
215 increased until failure. Before test, steel pipes were attached to the BFRP test specimens

216 according to CSA S806-12, Annex B “Anchor for testing FRP specimens under monotonic,

217
U
sustained, and cyclic tension”. Fig. 2 shows the test setup, dimensions of test specimen, and
AN
218 overview of the BFRP specimens attached with steel tubes. The test specimens were

instrumented with one linear variable differential transformer (LVDT, 200 mm length) to capture
M

219

220 specimen elongation during testing. For each tensile test, the specimen was mounted in the
D

221 tensile machine with the steel pipe anchors gripped by the wedges of the upper and the lower
TE

222 jaws of the machine. The average loading rate ranged between 250 to 560 MPa/min. The applied

223 load and bar elongation were recorded during the test using a data acquisition system monitored
EP

224 by a computer. Due to the brittle nature of FRP, no yielding occurs and the stress-strain behavior

225 was linear.


C

226 Transverse-Shear Test Procedure


AC

227 Transverse shear tests were performed according to ASTM D7617 (ASTM 2011) to characterize

228 the BFRP bars. The transverse shear strengths of BFRP bar specimens were measured on

229 samples subjected to elevated temperature (60oC) for 1,000; 3,000; and 5,000 h of immersion in

230 alkaline solution. This property was investigated to provide some indications on the retention of

10
ACCEPTED MANUSCRIPT

231 the matrix related mechanical properties of BFRP bars subjected to high temperatures and

232 alkaline environment. As shown in Fig. 4, the test setup consisted of a 230 × 100 × 110 mm steel

233 base equipped with lower blades spaced at 50 mm face to face, allowing for the double

234 transverse-shear failure of the specimen caused by an upper blade. All specimens were cut at a

PT
235 length of 350 mm and tested in shear according to ASTM D7617. Six unconditioned and

RI
236 conditioned specimens 350 mm in length were tested under laboratory conditions with an MTS

237 810 testing machine equipped with a 500-kN load cell. A displacement-controlled rate of 1.3

SC
238 mm/min was used during the test, which yielded between 30 and 60 MPa/min until specimen

239 failure. The loading was done without subjecting the test specimens to any shock. The

240
U
transverse-shear strength of BFRP bars, τ u , was calculated as τ u = Ps / 2 A , where Ps is the failure
AN
241 load (N); A is the cross-sectional area (mm2).
M

242 Flexural Test Procedure

243 Flexural tests were conducted on unconditioned and conditioned BFRP specimens for 1,000;
D

244 3,000; and 5,000 h, at 60oC (ASTM 4476). All BFRP specimens were cut at a length of 300 mm
TE

245 and an overhang of 10% of the supported span was allowed at each support as shown in Fig. 5.

246 Six unconditioned and conditioned specimens were tested under laboratory conditions using
EP

247 MTS 810 testing machine equipped with 500 kN load cell. The specimens were loaded at the

248 midspan with a circular nose; the specimen ends rested on two circular supports that allowed the
C

specimens to bend. A displacement-controlled rate of 3.0 mm/min was used during the test. The
AC

249

250 rate of loading was done without subjecting the test specimen to any shock. The applied load and

251 deflection were recorded during the test using a data acquisition system monitored by a

252 computer. The flexural strength of the BFRP bars, fu, was determined as fu = PLC/(4I), where P is

11
ACCEPTED MANUSCRIPT

253 the failure load, L is the clear span, C is the distance to the centroid of the extreme-most fibers,

254 and I is the moment of inertia.

255 Short-Beam Shear Test Procedure (Interlaminar Shear Test)

256 In pultruded FRP bars, fibers are arranged unidirectionally and bonded with the polymer matrix,

PT
257 the horizontal stresses would be more conducive to inducing interface degradation than

RI
258 transverse-shear stresses (Park et al. 2008). The shear test was conducted on six specimens of

259 BFRP bar according to ASTM D4475 (ASTM 2008) in order to calculate the interlaminar shear

SC
260 strength, which is governed by the fiber/matrix interface. The tests were carried out with a 500-

261 kN MTS 810 testing machine. The distance between the shear planes was set to 6 times the

262
U
nominal diameter of the FRP bars. Fig. 6 shows the test setup and typical modes of failure of the
AN
263 tested specimens. A displacement-controlled rate of 1.3 mm/min was employed during the test.

The applied load was recorded with a computer-monitored data acquisition system. The
M

264

265 interlaminar shear strength, Su, of the BFRP bars was determined as S u = 0.849 P / t 2 , where Su
D

266 interlaminar-shear strength (MPa); P = shear failure load (N); and t = thickness of BFRP
TE

267 specimen (mm).

268 TESTS RESULTS AND DISCUSSION


EP

269 Physical Properties and SEM Results

270 Table 1 presents the physical properties of the unconditioned basalt-FRP bars, where the glass
C

transition temperature (Tg) was determined with differential scanning calorimetry (DSC) [ASTM
AC

271

272 D3418 (ASTM 2012b)] (see Fig.7). From Table 1, the test results indicated that the basalt-fiber

273 content in weight of basalt-FRP bars was 81.0%. The mass percentages of water uptake after 24

274 h, 7 days and at saturation were found to be 0.04%, 0.18, and 0.25% on average. The water-

275 absorption values obtained not exceed the limits specified in CSA S807 and ACI 440.6M-08

12
ACCEPTED MANUSCRIPT

276 (1%). The material’s cure ratio is high (100%). Table 1 presents, also, the Tg values for the first

277 and second heating of unconditioned samples (reference samples). Note that, for the

278 unconditioned specimen, the Tg corresponding to the second heating run was close to the Tg

279 corresponding to the first scan. These results confirm the high cure ratio of more than 99%

PT
280 measured by DSC and shown in Table 1.

RI
281 Scanning electron microscopy (SEM) observations and image analysis were performed to

282 observe the microstructure of BFRP specimens. All specimens observed in the SEM were first

SC
283 cut, polished, and coated with a thin layer of gold-palladium by a vapor-deposit process. After

284 coating the surfaces, microstructural observations were performed on a JEOL JSM-840A SEM.

285
U
These observations were conducted to see any defect in polymer matrix, basalt fibers, or
AN
286 interfaces, if any. A general view of basalt-FRP bar was obtained as shown in Fig.8. No pore or

void was found in a SEM micrograph at the interfaces between the fibers and matrix. It has to be
M

287

288 noted that the material contains fibers with various diameters as shown in Fig. 9. SEM
D

289 micrograph of interface between basalt and resin matrix in Fig. 10 clearly shows that no defect in
TE

290 polymer matrix, basalt fibers, or interfaces was observed. The bonding between the basalt fibers

291 and the thermosetting resin is good since there were no free gaps at the interface.
C EP
AC

13
ACCEPTED MANUSCRIPT

292 Mechanical Properties Results

293 Tensile Strength, Modulus of Elasticity, and Strain

294 The tensile test of unconditioned BFRP specimens showed an approximately linear behavior up

295 to failure. Specimens failed through rupture of fibers. The failure was accompanied by

PT
296 delamination of fibers and resin, as shown in Fig. 11. Table 2 presents the experimental results

RI
297 obtained during the tensile tests concerning the ultimate tensile strength, modulus of elasticity,

298 and ultimate strain of unconditioned BFRP bars. As shown in Table 2, the tensile strength,

SC
299 modulus of elasticity, and ultimate strain of unconditioned BFRP bars were equal to 1,646±12

300 MPa; 69.7 GPa; and 2.4%, while these values, as reported by the manufacturer, are 1,636 MPa;

301
U
64 GPa; and 2.5%, respectively. Fig. 11 shows the mode of failure during the tensile tests for
AN
302 unconditioned (reference) BFRP bars.

Effect of Conditioning on Transverse-Shear Strength


M

303

304 Table 2 shows that the transverse-shear strength of the unconditioned basalt-FRP bars was
D

305 241±2.1 MPa, while Table 3 shows the transverse-shear strength and strength-retention ratio of
TE

306 the tested BFRP bars after 1,000, 3,000, and 5,000 h of immersion in the alkaline solution at

307 60°C. Table 3 indicates that the BFRP bars were affected by accelerated aging with a transverse-
EP

308 shear strength reduction of 12% after 5,000 h of immersion. Fig. 12 shows the effect of the

309 alkaline solution on the transverse shear strength after different exposure times. The BFRP bars
C

310 exhibited no significant reductions in the early stages (less than 3,000 h).
AC

311 Effect of Conditioning on Flexural Strength

312 The three-point flexural strength, flexural modulus of elasticity, and ultimate outer-fiber strain of

313 the tested unconditioned BFRP bars, are presented in Table 2. The flexural strength and flexural

314 modulus of elasticity were 653±15 MPa and 41.0 GPa, respectively. Table 3 and Fig. 13a show

14
ACCEPTED MANUSCRIPT

315 the experimental results obtained during the flexural tests of aged BFRP bars tested after

316 immersion in alkaline solution. Fig. 13a shows flexural strength loss observed with BFRP

317 reinforcing bars conditioned in alkaline solution during 5000 h (208 days). The BFRP bars had

318 flexural-strength reductions of 2% and 10 % after 1,000 h, and 3,000 h, while the bars showed a

PT
319 lower reduction of 19 % after 5,000 h. These observations confirm that the bond between the

RI
320 BFRP fibers and resin after conditioning was lower than that before conditioning. Consequently,

321 debonding occurring at the fiber–matrix interface caused the fibers to separate from the resin.

SC
322 Fig. 13a shows the effect of the alkaline solution on flexural strength. Fig. 5b shows the mode of

323 failure of failure during the flexural tests for reference and specimen aged at 60oC during 5,000

324
U
h. It can be seen in Fig. 5a that the tested specimen was loaded until rupture occurs in the
AN
325 extreme tensile fibers and that aged and reference BFRP bars have similar mode of failure. Note

that the extreme compressive fibers did not show any premature compression failure. Progressive
M

326

327 damage propagation occurred through matrix cracking, fiber breaking, and interfacial debonding
D

328 of the extreme tensile fibers side of specimens.


TE

329 Table 3 gives, also, the flexural modulus of elasticity and the retention ratio of the tested BFRP

330 bars after 1,000, 3,000, and 5,000 h of immersion. The bars had no significant differences in
EP

331 flexural modulus of elasticity after 5,000 h. The reduction ranged from 3% to 11% in comparison

332 to the references. Fig. 13b illustrates the effect of the alkaline solution at 60oC on the flexural
C

333 modulus of elasticity, with all bar specimens exhibiting a steady reduction rate.
AC

334 Effect of Conditioning on Interlaminar-shear strength

335 The short-beam shear test revealed that the unconditioned BFRP bars had the interlaminar-shear

336 strength of 30.2±2.7 MPa, (Table 2). It is worth mentioning that the high values of the

15
ACCEPTED MANUSCRIPT

337 interlaminar-shear strength reveal a strong interface between the resins and reinforcing fibers,

338 which was clarified in the SEM analysis of the unconditioned BFRP specimen.

339 Table 3 also shows the apparent horizontal shear (interlaminar shear) strength and strength-

340 retention ratios of the tested FRP bars after 1,000, 3,000, and 5,000 h of immersion in alkaline

PT
341 solution. The reduction ratio for the BFRP bars after 5,000 h was 21%. Fig. 14 shows the effect

RI
342 of the alkaline solution on the interlaminar-shear strength, with the BFRP bars, for different time

343 of immersions.

SC
344 PREDICTION OF LONG-TERM BEHAVIOR AND SERVICE LIFE FOR

345 TRANSVERSE SHEAR STRENGTH

346 Arrhenius model


U
AN
347 Several researchers have been employed the Arrhenius concept for predicting the long-term

performance of FRP (Bank et al. 2003; Robert et al. 2013, 2009; Benmokrane et al. 2016, Chen
M

348

349 et al. 2006, 2007). Eq. (4) expresses the Arrhenius relation-ship, in terms of the degradation rate
D

350 (Nelson 1990).


TE

 − Ea 
 
351 k= Ae  RT  (4)
EP

352 where k = degradation rate; A = constant relative to the material and degradation process; Ea =

353 reaction’s energy of activation; R = universal gas constant; and T = temperature in °C. The
C

354 primary assumption with this model is that only one dominant degradation mechanism of the
AC

355 material operates during the reaction and that this mechanism will not change with time and

356 temperature during the exposure (Chen et al. 2007). Only the rate of degradation will be

357 accelerated as the temperature increases.

358 Dejke (2001) used another approach to generate a relative TSF. He proposed using the TSF to

359 transform the time in the accelerated test to actual service lives for GFRP reinforcement.

16
ACCEPTED MANUSCRIPT

360 Because the time for a certain reaction to take place must be proportional to the inverse of the

361 rate of reaction, Dejke (2001) proposed determining the TSF in accordance with Eq. (5).

1 1
B( − )
t
362
TSF = 1 = e T1 + 273.15 T2 + 273.15 (5)
t2

PT
363 where T1 and T2 = exposure temperatures (°C); and t1 and t2 = times required to obtain a certain

RI
364 level of decrease in mechanical property at temperatures T1 and T2, respectively. The TSF is

sensitive to activation energy, and a good estimate of Ea is needed to generate a reasonable TSF.

SC
365

366 According to the procedure proposed by Bank et al. (2003), the natural logarithm of time to

U
367 reach a set of levels of normalized performances versus 1/T, expressed as the inverse of absolute
AN
368 temperature (1,000/K), was used to predict the service life of the BFRP bars at mean annual

369 temperatures (MAT) of 10°C. The temperature of 10°C corresponds well to the mean average
M

370 temperature of northern regions, where deicing salts are often used. A coefficient of

371 determination (R2) value close to 1 was desired. ASTM procedures, however, recommend a
D

372 minimum value of 0.80 for acceptability; the obtained R2 values were between 0.96 and 0.99.
TE

373 The service-life time necessary to reach the established tensile-strength retention levels (PR) can

374 be extrapolated for any temperature from the Arrhenius plot. Consequently, predictions were
EP

375 made for transverse shear-strength retention as a function of time for immersion at 10°C, and the

general relation between the PR and the predicted service life at the average temperature of 10oC
C

376
AC

377 can be drawn (Fig. 16). The predicted time to reach the determined tensile-strength retention

378 level (PR) for the BFRP bars aged in an alkaline solution simulating the concrete environments

379 at an isotherm temperature of 10oC is approximately 150 years for a PR of more than 80%.

380 Moreover, the predicted service life of the BFRP bars aged in the alkaline solution at an isotherm

381 temperature of 10oC with a PR of less than 80% can be estimated as being infinity. These

17
ACCEPTED MANUSCRIPT

382 predictions show that the BFRP bars tested in this study are a high durability in a concrete

383 environment. On the other hand, the prediction curves for the BFRP bars predicted transverse

384 shear-strength retention of 82% after a service life of 75 years. Table 4 presents the transverse

385 shear-strength retention after 10, 25, 50, 75, and 100 years of service at MATs of 10°C and

PT
386 30°C. Table 4 indicates that, even after a service life of 100 years, which corresponds to the

RI
387 maximum expected service life, the shear-strength retention was still 81.3 %, for BFRP bars. At

388 30°C, the shear-strength retention of BFRP bars was 76.1, 77.0, 78.1, and 79.0% after a service

SC
389 life of 200, 150, 100, 75 years, respectively, (Table 4).

390 Fib Bulletin (40) method

391
U
In 2007, FIB Bulletin (40) Task Group 9.3 proposed durability design approach for FRP bars by
AN
392 incorporating factors of relative humidity (RH), exposed mean average temperature (MAT), and

services life (FIB 2007). The long-term design transverse shear-strength is given by Eq. (6)
M

393

f fko
f fd =
D

394 (η env ,t .γ f ) (6)


TE

395 where ffd is a long-term design value of transverse shear-strength for BFRP bars; ffko is the

396 characteristic value of transverse shear-strength (short term), which the composite bars can resist
EP

397 after exposure to a practical test environment for 1,000 h and this value can be expressed as a

percentage of the transverse shear-strength or as absolute value; ɣf is material factor; ηenv,t is


C

398
AC

399 environmental strength reduction factor which is obtained by Eq. (7)

1
η env ,t = (7)
400
[(100 − R10 ) / 100]n

401 where R10 is the standard reduction in transverse shear-strength in percent per decade

402 (logarithmic decade) due to environmental effects, which can be extrapolated from each

18
ACCEPTED MANUSCRIPT

403 individual degradation line (Fig. 15). The exponent n is the sum of four influence terms as shown

404 in Eq. (8)

405 n = nmo + nT + nSL + nd (8)

PT
406 where nmo, nT, nSL, nd are the influence terms for moisture condition, temperature, desired service

407 life, and diameter influence, respectively.

RI
408 It can be seen that an exponential approach is used in the design strength equation, for which the

409 reason is because the deterioration is described best by the kinetics of the chemical and physical

SC
410 processes (Weber 2006). This assumption itself is reasonable; however, as can be seen, the base

U
411 of the exponential in expressions (6) and (7) is the strength retention (1−R10), other than the
AN
412 strength reduction R10. As known, the kinetics of the chemical and physical processes is linked to

413 the outer layer of BFRP bar that degrades by the environmental attacks. Hence, the power terms
M

414 n, including the effects of moisture, temperature, time, and diameter, should be a link to the outer

415 degraded part of BFRP bar reflected as R10. Thus, the design strength equation in the FIB (2007)
D

416 approach is of low confidence also.


TE

417 The values of the environmental influence parameter R10 of the BFRP specimens conditioned at

418 22°C, 40°C, and 60°C were 7.3%, 7.7%, and 9.5%. From the curve fitting, the shear-strength
EP

419 retentions after 100 years were 78%, 76%, and 68% at 22°C, 40°C, and 60°C, respectively.
C

420 According to fib Bulletin 40 (2007), for instance, nmo= 1 and nSL = 3.0 at a service life of 100
AC

421 years, assuming a moisture-saturated condition. As adopted by Serbescu et al. (2014), nT is equal

422 to 0.5, 1.5, and 2.5 at 22°C, 40°C, and 60°C, respectively. The value R10 for all the environments

423 tested can be determined by using the average slope of the individual degradation lines,

424 assuming that the degradation rate is similar regardless of environment (Serbescu et al. 2014).

425 Thus, R10 is equal to 8.17%. The estimated shear-strength retention (1/ ηenv,b) is equal 68.1%,

19
ACCEPTED MANUSCRIPT

426 62.6%, and 57.5% at 22°C, 40°C, and 60°C, respectively. Noticeable differences between the

427 two methods were observed for each environmental conditioning. The differences might be

428 attributed to increased concrete strength resulting from immersion. This is not considered in the

429 equation, nor are the effects of moisture diffusion on the degradation mechanism. Table 5 reports

PT
430 the (1/ηenv,b) predications at different moisture-saturated conditions and mean annual

RI
431 temperatures (MATs) after 100 years of service life. The shear-strength retentions after 100 years

432 of service life in dry, moist, and moisture saturated environments and MATs varied from 65% to

SC
433 88%, as reported on Table 5. In order to validate the FIB Bulletin (40) model’s reliability, further

434 work is needed with different accelerated moist environments, different service life influence,

435 and longer exposure times.


U
AN
436 SUMMARY AND CONCLUSIONS

In this research study, new basalt-FRP bars were exposed to an alkaline solution simulating a
M

437

438 concrete environment to determine the bras’ suitability as internal reinforcement for concrete
D

439 elements. Physical, mechanical, microstructural analyses and durability characterization were
TE

440 conducted. The BFRP specimens were immersed in alkaline solution (1,000; 3,000; and 5,000 h)

441 and subjected to elevated exposure temperature (60°C) to simulate the alkaline effect of concrete.
EP

442 In addition, differential-scanning calorimetry (DSC) and scanning electron microscopy (SEM)

443 were used to characterize the physical properties of BFRP bar specimens. Based on the results of
C

444 this study, the following conclusions may be drawn on the tested products:
AC

445 1- The test observation indicated that the basalt fiber content is 81% by weight and the

446 water uptake at saturation is equal to 0.25%. The cure ratio of the material is very high

447 (close to 100%) but its glass transition temperature is116oC (H) by DSC.

20
ACCEPTED MANUSCRIPT

448 2- Optical microscopy and electronic scanning microscopy (SEM) analysis of the

449 unconditioned BFRP bars showed that no defect in polymer matrix, basalt fibers, or

450 interfaces was observed. The bonding between the basalt fibers and the thermosetting

451 resin was good since there was no free gap at the interface.

PT
452 3- The results indicate that the transverse-shear strength of the BFRP specimens was slightly

RI
453 affected by increasing the immersion duration at higher temperature levels. After 5,000 h

454 of immersion in the alkaline solution at 60oC, test result indicated that 12 % degradation

SC
455 in the transverse-shear occurred.

456 4- The flexural strength of the BFRP bars was significantly affected by accelerated aging

457 (19% reduction after 5,000 h, at 60oC).


U
AN
458 5- The interlaminar-shear strength of the BFRP bars was highly affected by accelerated

aging (21% reduction after 5,000 h, at 60oC). The fiber–resin interface plays a significant
M

459

460 role in controlling the degradation due to conditioning.


D

461 6- According to the long-term predictions, the transverse shear-strength retention of the
TE

462 BFRP bars immersed in the alkaline solution will decrease by 19.8% and 23.0% after

463 150 years at an isotherm temperature of 10°C and 30°C, respectively. It was shown that
EP

464 the BFRP bars service life with a transverse shear-strength retention of less than 79.6 and

465 76.1% at 10oC and 30oC, respectively, should be infinite.


C

466 ACKNOWLEDGMENTS
AC

467 This study was conducted with funding from the Fonds de recherche du Quebec en nature et

468 technologie (FRQNT), the Natural Sciences and Engineering Research Council of Canada

469 (NSERC) Research Chair in Innovative FRP Reinforcement for Concrete Infrastructure, and the

470 service des matériaux d'infrastructures of the Ministry of Transportation of Quebec. The authors

21
ACCEPTED MANUSCRIPT

471 wish to express their gratitude and appreciation to Pultrall Inc., Thetford Mines, Quebec, for

472 material support. The technical assistance from the staff of the Structural Laboratory in the

473 Department of Civil Engineering, Faculty of Engineering at the University of Sherbrooke is also

474 acknowledged.

PT
475 REFERENCES

RI
476 ACI (American Concrete Institute). (2004). “Guide test methods for fiberreinforced polymers

477 (FRPs) for reinforcing or strengthening concrete structures.” ACI 440.3R-04, Farmington Hills,

SC
478 MI.

479 ACI (American Concrete Institute). (2004). “Prestressing concrete structures with FRP tendons.”

480 ACI 440.4R-04, Farmington Hills, MI.


U
AN
481 ACI (American Concrete Institute). (2008). “Specification for carbon and glass fiber-reinforced

polymer bar materials for concrete reinforcement.” ACI 440.6M-08, Farmington Hills, MI.
M

482

483 ACI (American Concrete Institute). (2015). “Guide for the design and construction of structural
D

484 concrete reinforced with FRP bars.” ACI 440.1R-15, Farmington Hills, MI.
TE

485 Ali, A. H., Benmokrane, B., Mohamed, H.M., Manalo, A., and El-Safty, A. (2018b). “Statistical

486 Analysis and Theoretical Predictions of the Tensile Strength Retention of GFRP Bars based on
EP

487 Resin Degradation.” Journal of Composite Materials, 52(21), 1-20,

488 https://doi.org/10.1177/0021998318755866.
C

489 Ali, A. H., Mohamed, Benmokrane, B., and El-Safty, A. (2018a). “Effect of Applied Sustained
AC

490 Load and Severe Environments on Durability Performance of Carbon-Fiber Composite Cables

491 (CFCCs).” Journal of Composite Materials, 1-16, https://doi.org/ 10.1177/0021998318789742.

22
ACCEPTED MANUSCRIPT

492 Ali, A. H., Mohamed, H. M., and Benmokrane, B. (2017a). “Shear strength of circular concrete

493 beams reinforced with glass fiber-reinforced polymer bars and spirals.” ACI Structural Journal,

494 114(1), 39-49.

495 Ali, A. H., Mohamed, H. M., Benmokrane, B., and ElSafty, A. (2017b). “Influence of Resin

PT
496 Type on Physical, Mechanical and Durability Performance of Glass-FRP Bars in Concrete

RI
497 Environment at Elevated Temperature.” Fifth International Conference on Durability of Fibre

498 Reinforced Polymer (FRP) Composites for Construction and Rehabilitation of Structures (CDCC

SC
499 2017), 19-21th July, 2017, Sherbrooke, Qc, Canada.

500 Ali, A. H., Mohamed, H.M., Chaallal, O., Benmokrane, B., and Ghrib, F. (2018c). “Shear

501
U
Resistance of RC Circular Members with FRP Discrete Hoops versus Spirals.” Engineering
AN
502 Structures Journal, Vol. 174, 688-700.

Ali, N.M., Wang, X., Wu, Z., and Hassanein, A.Y. (2015). “Basalt fiber reinforced polymer grids
M

503

504 as an external reinforcement for reinforced concrete structures.” Journal of Reinforced Plastics
D

505 Composites, 34(19), 1615-1627.


TE

506 Arias, J.P.M., Vazquez, A., and Escobar, M. E. (2012). “Use of sand coating to improve bonding

507 between GFRP bars and concrete.” Journal of composite materials, 46(18), 2271-2278.
EP

508 ASTM. (2003). “Standard test method for compositional analysis by thermogravimetry.” ASTM

509 E1131, West Conshohocken, PA.


C

510 ASTM. (2003). “Standard test method for curing properties of pultrusion resins by thermal
AC

511 analysis.” ASTM D5028, West Conshohocken, PA.

512 ASTM. (2008). “Standard test method for apparent horizontal shear strength of pultruded

513 reinforced plastic rods by the short beam method.” ASTM D4475, West Conshohocken, PA.

23
ACCEPTED MANUSCRIPT

514 ASTM. (2009). “Standard test method for flexural properties of fiber reinforced pultruded plastic

515 rods.” ASTM D4476, West Conshohocken, PA.

516 ASTM. (2010). “Water absorption of plastics.” ASTM D570, West Conshohocken, PA.

517 ASTM. (2011). “Standard test method for transverse shear strength of fiber-reinforced polymer

PT
518 matrix composite bars.” ASTM D7617, West Conshohocken, PA.

RI
519 ASTM. (2011). “Tensile properties of fiber reinforced polymer matrix composite bars.” ASTM

520 D7205, West Conshohocken, PA.

SC
521 ASTM. (2012b). “Standard test method for transition temperatures and enthalpies of fusion and

522 crystallization of polymers by differential scanning calorimetry.” ASTM D3418, West

523 Conshohocken, PA.


U
AN
524 Benmokrane, B. Ali, A. H., Mohamed, H. M. and Safty, A. (2014). “Long-Term Tensile

Properties of Carbon FRP Cable.” 15th International European Bridge Conference, London, UK
M

525

526 8-10th, July 2014.


D

527 Benmokrane, B., Ali, A. H., and Mohamed, H. M. (2012). “Durability of glass FRP solid and
TE

528 hollow bars (rock bolts) for the application in ground control of the jurong rock caverns in

529 Singapore.” Univ. of Sherbrooke, QC, Canada.


EP

530 Benmokrane, B., Ali, A. H., Mohamed, H. M., ElSafty, A., Manalo, A. (2017). “Laboratory

531 Assessment and Durability Performance of Vinyl-Ester, Polyester, and Epoxy Glass-FRP Bars
C

532 for Concrete Structures.” Journal Composites: Part B, Vol. 114, 163-174.
AC

533 Benmokrane, B., Ali, A. H., Mohamed, H., Robert, M., and ElSafty, A. (2016a). “Durability

534 Performance and Service Life of CFCC Tendons Exposed to Elevated Temperature and Alkaline

535 Environment.” J. Compos. Constr., 10.1061/(ASCE)CC.1943-5614.0000606, 04015043.

24
ACCEPTED MANUSCRIPT

536 Benmokrane, B., Elgabbas, F., Ahmed, E., and Cousin, P. (2015). “Characterization and

537 Comparative Durability Study of Glass/Vinylester, Basalt/Vinylester, and Basalt/Epoxy FRP

538 Bars.” J. Compos. Constr., 10.1061/(ASCE)CC.1943-5614.0000564, 04015008.

539 Benmokrane, B., Robert, M., Mohamed, H., Ali, A. H., and Cousin, P. (2016b). “Durability

PT
540 Assessment of Glass FRP Solid and Hollow Bars (Rock Bolts) for Application in Ground

RI
541 Control of Jurong Rock Caverns in Singapore.” J. Compos. Constr., 10.1061/(ASCE)CC.1943-

542 5614.0000775 , 06016002.

SC
543 Benmokrane,B., Mohamed, H. M., and Ali, A. H. (2018). “Service-Life-Prediction and Field

544 Application of GFRP Tubular and Solid Bolts based on Laboratory Physical and Mechanical

545 Assessment.” Journal of


U Composite Materials, 1-15.
AN
546 https://doi.org/10.1177/0021998318764806.

Chen, Y., Davalos, J. F., and Ray, I. (2006). “Durability prediction for GFRP bars using short-
M

547

548 term data of accelerated aging tests.” J. Compos. Constr., 10(4), 279-286.
D

549 Chen, Y., Davalos, J. F., Ray, I., and Kim, H. Y. (2007). “Accelerated aging tests for evaluation
TE

550 of durability performance of FRP reinforcing bars reinforcing bars for concrete structures.”

551 Compos. Struct., 78(1), 101-111.


EP

552 CSA (Canadian Standards Association). (2010). “Specification for fibre reinforced polymers.”

553 CAN/CSA S807-10, Toronto.


C

554 CSA (Canadian Standards Association). (2012). “Design and construction of building structures
AC

555 with fiber reinforced polymers.” CAN/CSA S806-12, Toronto.

556 Davalos, J .F., Chen, Y., and Ray, I. (2012) “Long-term durability prediction models for GFRP

557 bars in concrete environment.” J. Comp. Mater., 46: 1899-1914.

25
ACCEPTED MANUSCRIPT

558 Davalos, J.F, Chen, Y., and Ray, I. (2011). “Long-term durability prediction models for GFRP

559 bars in concrete environment.” Journal of composite materials, 46(16), 1899-1914.

560 Elgabbas, F., Ahmed, E., Benmokrane, B. (2015) “Physical and mechanical characteristics of

561 new basalt-FRP bars for reinforcing concrete structures” Constr. and Build. Mater., Vol. 95,

PT
562 623–635.

RI
563 Fib Bulletin 40. FRP reinforcement in concrete structures. International Federation for 440

564 Structural Concrete, Lausanne, Switzerland. 2007.

SC
565 INFOMINE Research Group, Basalt fiber-based thermal insulating materials market research in

566 Russia, Moscow, 2007.

567
U
Karbhari, V.M., and Stachowsky, C., Wu, L. (2007). “Durability of pultruded E glass/vinylester
AN
568 under combined hygrothermal exposure and sustained bending.” J. Compos. Constr., 19 (8), 665-

673.
M

569

570 Katsuki, F., and Uomoto, T. (1995). "Prediction of Deterioration of FRP Rods due to Alkali
D

571 Attack." Proceedings of the Second International RILEM Symposium (FRPRCS-2), Non-
TE

572 Metallic (FRP) Reinforcement for Concrete Structures, L. Taerwe, ed., E&FN Spon, London, pp.

573 83-89.
EP

574 Ng, SC., and Lee, S. (2002). “A study of flexural behavior of reinforced concrete beam

575 strengthened with carbon fiberreinforced plastic (CFRP).” Journal of Reinforced Plastics
C

576 Composites, 21, 919–938.


AC

577 Parnas, R., Shaw, M., and Liu, Q. (2007). “Basalt fiber reinforced polymer composites.”

578 Technical Rep. No. NETCR63, New England Transportation Consortium C/O Advanced

579 Technology & Manufacturing Center, Univ. of Massachusetts Dartmouth, Fall River, MA.

580 Pultrall. (2015). “Composite reinforcing rods technical data sheet.” Thetford Mines, Canada.

26
ACCEPTED MANUSCRIPT

581 Robert, M., Wang, P., Cousin, P., and Benmokrane, B. (2010). “Temperature as an accelerating

582 factor for long term durability testing of FRPs should there be any limitations.” J. Compos.

583 Constr., 14(4), 361-367.

584 Shen, C. H., and Springer, G. S. (1976). “Moisture Absorption and Desorption of Composite

PT
585 Materials.” Journal of Composite Materials, Vol. 10, pp. 2-20.

RI
586 Shen, D., Ojha, B., Shi, X., Zhang, H., and Shen, J. (2016). “Bond stress–slip relationship

587 between basalt fiber-reinforced polymer bars and concrete using a pull-out test.” Journal of

SC
588 Reinforced Plastics Composites, 35(9), 747-763.

589 Sim, J., Park, C., and Moon, D. Y. (2005). “Characteristics of basalt fiber as a strengthening

590
U
material for concrete structures.” Compos. Part B, 36(6–7), 504–512.
AN
591 Taljsten B., Carolin, A. and Nordin, H. (2003). “Concrete structures strengthened with near

surface mounted reinforcement of CFRP.” Advances in Structural Engineering, 6, 201–213.


M

592

593 Tannous, F. E., and Saadatmanesh, H. (1998). "Environmental Effects on the Mechanical
D

594 Properties of E-glass FRP Rebars." ACI Materials Journal, 95(2), pp. 87-100.
TE

595 Wang, J., GangaRao, H., Liang, R., and Liu, W. (2016a) “Durability and prediction models of

596 fiber-reinforced polymer composites under various environmental conditions: A critical review.”
EP

597 Journal of Reinforced Plastics Composites, 35(3), 179-211.

598 Wang, J., GangaRao, H., Liang, R., Zhou, D., Liu, W., and Fang, Y. (2015). “Durability of glass
C

599 fiber-reinforced polymer composites under the combined effects of moisture and sustained
AC

600 loads” Journal of Reinforced Plastics Composites, 34(21), 1739-1754.

601 Wang, X., Wu, G., Wu, Z. (2014). “Evaluation of prestressed basalt fiber and hybrid fiber

602 reinforced polymer tendons under marine environment.” Material Design, 64, 721–728. 9.

27
ACCEPTED MANUSCRIPT

603 Wang, X., Zhao, X., Wu, Z., Zhu, Z., and Wang, Z. (2016b). “Interlaminar shear behavior of

604 basalt FRP and hybrid FRP laminates.” Journal of composite materials, 50(8), 1073-1084.

605 Won, J.P., and Park, C.G. (2006). “Effect of Environmental Exposure on the Mechanical and

606 Bonding Properties of Hybrid FRP Reinforcing Bars for Concrete Structures.” Journal of

PT
607 composite materials, 40(12), 1063-1076.

RI
608 Zhishen, W., Xin, W., Gang, W. (2012) “Advancement of structural safety and sustainability

609 with basalt fiber reinforced polymers.” Proc. of 6th International Conference on FRP

SC
610 Composites in Civil Engineering (CICE), Rome, Italy, IIFC, 2012, p. 29.

611

612
U
AN
613
M

614

615
D

616
TE

617

618
EP

619

620
C

621
AC

622

623

624

625

28
ACCEPTED MANUSCRIPT

626 List of Tables

627 Table 1. Physical properties of the reference basalt-FRP bars.

628 Table 2. Mechanical properties of the reference basalt-FRP bars.

PT
629 Table 3. Retention of mechanical properties of the conditioned BFRP bars.

630 Table 4. Transverse shear-strength retention at different service-life periods at MATs of 10 and

RI
631 30°C based on Arrhenius model.

SC
632 Table 5. Transverse shear-strength retention predications after service life of 100 years based on
633 the method in fib Bulletin 40.

U
634
AN
635

636
M

637

638
D

639
TE

640

641
EP

642

643
C
AC

644

645

646

647

29
ACCEPTED MANUSCRIPT

648 Table 1. Physical properties of the reference basalt FRP bars


Specified limit
Property Basalt FRP bars
ACI 440.6 CSA S807-10
Fiber content by weight (%) 81 55% (by vol.) 70%(by wt.)
Cure ratio (%) 100 N/A 93 (D2); 95 (D1)
Transverse CTE, (x10-6oC-1) 22 N/A 40
Run 1 116
Glass transition temperature (Tg) 100 80 (D2); 100(D1)

PT
Run 2 117
Moisture uptake (%) 0.25 <1.0 1.0 (D2); 0.75 (D1)
Diameter (mm) 20.0 N/A N/A
Cross-sectional area (mm2) 362 N/A N/A

RI
Wicking dots 0.0 N/A N/A
649 *Note: CTE = coefficient of thermal expansion. D1 and D2 classifications can be found in CSA S807-10.
650 *The mechanical properties were calculated using the nominal cross-sectional area.
651

SC
652
653
654
655
656

U
657
658
659
AN
660
661
662
663
664
M

665
666
667
668
D

669
670
671
TE

672
673
674
675
EP

676
677
678
679
680
C

681
682
683
AC

684
685
686
687
688
689
690
691
692
693
694
695
696

30
ACCEPTED MANUSCRIPT

697 Table 2. Mechanical properties of the reference basalt FRP bars


Specified limit
Property BFRP bars
ACI 440.6 CSA S807-10
Transverse shear strength (MPa) 241±2.1 >124 MPa >160 MPa
Tensile strength (MPa) 1646 -- --
Tensile modulus of elasticity (GPa) 69.7 39.3 GPa 40.0 GPa
Ultimate strain (%) 2.4 >1.2% >1.2%

PT
Short-beam shear strength (MPa) 30±0.5 -- --
Four-point flexural strength (MPa) 653±15 N/A N/A
Flexural modulus (stiffness) (GPa) 77.2±0.3 N/A N/A
698 *The mechanical properties were calculated using the nominal cross-sectional area.

RI
699

700

SC
701

702

U
703
AN
704

705
M

706

707
D

708
TE

709

710
EP

711

712
C

713
AC

714

715

716

717

718

31
ACCEPTED MANUSCRIPT

719

720 Table 3. Retention of mechanical properties of the conditioned BFRP bars


Conditioned Temp. τu Ret. fu Ret. E Ret. Su Ret.
Fiber/resin (oC) (%) (%) (%) (%)
period (MPa) (MPa) (GPa) (MPa)
22 240 99.6 647 99.1 77.0 99.7 29.5 98.3
1,000 40 237 98.3 646 89.9 76.8 99.5 28.2 94.0

PT
60 233 96.7 642 98.3 75.8 98.2 28.0 93.3
22 239 99.1 594 90.9 76.6 99.2 27.9 93.0
Basalt-FRP
3,000 40 233 96.7 592 90.6 76.7 99.3 27.3 91.0
bars
60 230 95.4 590 90.3 73.9 95.7 27.0 90.0

RI
22 227 94.2 540 83.0 76.2 98.7 25.8 86.0
5,000 40 222 92.1 533 81.6 76.1 98.5 25.3 84.3
60 214 88.8 528 80.0 70.8 91.7 24.1 80.0

SC
721 Note: τu= Transverse shear strength (MPa); fu= Four-point flexural strength (MPa); Su = Short-beam shear strength (MPa); Ret. =
722 Retention of strength (%).

723

U
724
AN
725

726
M

727

728
D

729
TE

730

731
EP

732

733
C

734
AC

735

736

737

738

739

32
ACCEPTED MANUSCRIPT

740

741 Table 4. Transverse shear-strength retention at different service-life periods at MATs of 10 and
742 30°C based on Arrhenius model
Transverse shear-strength retention
Service life (%)
(Years)
10oC 30oC

PT
10 86.9 84.9
25 84.7 82.1
50 83.0 80.2
75 82.0 79.0

RI
100 81.3 78.1
150 80.2 77.0
200 79.6 76.1

SC
743

744

U
745
AN
746

747
M

748

749
D

750
TE

751

752
EP

753

754
C

755
AC

756

757

758

759

33
ACCEPTED MANUSCRIPT

760 Table 5. Transverse shear-strength retention predications after service life of 100 years based on
761 the method in fib Bulletin 40
Moisture
Material
condition
nmo MAT (oC) nT nSL n ηenv,b 1/ ηenv,b
<5 -0.5 3.0 1.5 1.136 87.9%
Dry 5~15 0.0 3.0 2.0 1.185 84.3%
-1
(RH = 50%) 15~25 0.5 3.0 2.5 1.237 80.8%

PT
25~35 1.0 3.0 3.0 1.291 77.4%
<5 -0.5 3.0 2.5 1.237 80.8%
Basalt-FRP
Moist 5~15 0.0 3.0 3.0 1.291 77.4%
bars (12 mm 0
(RH = 80%) 15~25 0.5 3.0 3.5 1.347 74.2%

RI
in diameter)
25~35 1.0 3.0 4.0 1.406 71.1%
<5 -0.5 3.0 3.5 1.347 74.2%
Moisture
5~15 0.0 3.0 4.0 1.406 71.1%
saturated +1

SC
15~25 0.5 3.0 4.5 1.407 68.1%
(RH = 100%)
25~35 1.0 3.0 5.0 1.531 65.3%
762 * Note that: R10=8.17% and nSL = 3.0

U
763

764
AN
765
M

766

767
D

768
TE

769

770
EP

771

772
C

773
AC

774

775

776

777

778

34
ACCEPTED MANUSCRIPT

779

780 List of Figures

781 Fig. 1. Tested basalt-FRP bars

782 Fig. 2. Tensile test: (a) test setup; (b) specimen dimensions (mm); (b) overview of the basalt-

PT
783 FRP specimens attached with steel pips

784 Fig. 3. Stainless-steel container built for aging the BFRP bar specimens in alkaline solution at

RI
785 60ºC

SC
786 Fig. 4. Transverse shear test setup and specimens at failure

787 Fig. 5. Typical flexural test setup and specimens at failure

U
788 Fig. 6. Interlaminar-shear test setup (short-beam test) and specimens at failure
AN
789 Fig. 7. DSC scans for glass transition temperature (Tg)

790 Fig. 8. Micrographs of the BFRP bars cross-section (20 mm-diameter)


M

791 Fig. 9. Micrographs of the basalt-FRP bar cross section at medium magnification
D

792 Fig. 10. Micrographs of the bars for basalt fiber/matrix interface
TE

793 Fig. 12. Effect of conditioning in the alkaline solution at 60°C on transverse-shear strength

794 Fig. 13. Effect of conditioning in the alkaline solution at 60°C on mechanical properties:
EP

795 (a) flexural strength; (b) flexural modulus of elasticity

796 Fig. 14. Effect of conditioning in the alkaline solution at 60°C on interlaminar-shear strength
C

797 Fig. 15. Plot of the transverse shear strength retention of BFRP bars as a function of time
AC

798 Fig. 16. General relation between the PR and the predicted service life at mean annual

799 temperatures of 10 and 22°C

800

801

802

35
ACCEPTED MANUSCRIPT

803

804

805

PT
20 mm

RI
U SC
806
AN
807 Fig. 1. BFRP bar 20 mm in diameter used in this investigation

808
M

809
D

810
TE

811

812
EP

813

814
C

815
AC

816

817

818

819

820

36
ACCEPTED MANUSCRIPT

821

822

823

BFRP specimen Steel tube

PT
700 1000 mm 700
20

RI
2400 mm
Di=38 mm
(LVDT, 200 mm)
(b) Do=48 mm

U SC
AN
Plastic Cover
for BFRP Bar
824
(a) (c)
825 Fig. 2. Tensile test: (a) test setup; (b) specimen dimensions (mm); (b) overview of the basalt-
M

826 FRP specimens attached with steel pips


D

827

828
TE

829

830
EP

831
C

832
AC

833

834

835

836

837

838

37
ACCEPTED MANUSCRIPT

839

840

PT
RI
U SC
841
AN
842 Fig. 3. Stainless-steel container built for aging the BFRP bar specimens in alkaline solution at

843 60ºC
M

844
845
D

846
TE

847

848
EP

849

850
C

851
AC

852

853

854

855

856

38
ACCEPTED MANUSCRIPT

857

858

PT
RI
SC
859

U
860 (a) Test setup (b) Specimens at failure
Fig. 4. Transverse shear test setup and specimens at failure.
AN
861

862
M

863

864
D

865
TE

866

867
EP

868

869
C

870
AC

871

872

873

874

875

39
ACCEPTED MANUSCRIPT

PT
RI
876

SC
877 (a) Test setup (b) Specimens at failure
878 Fig. 5. Typical flexural test setup and specimens at failure

879

880
U
AN
881

882
M

883
D

884

885
TE

886
EP

887

888
C

889
AC

890

891

892

893

894

895

40
ACCEPTED MANUSCRIPT

896

897

898

Vertical crack Loading head penetration

PT
RI
Horizontal crack

SC
Vertical cracks Horizontal cracks

U
899
900 (a) Test setup (b) Specimens at failure
AN
901 Fig. 6. Interlaminar-shear test setup (short-beam test) and specimens at failure.

902
M

903

904
D

905
TE

906

907
EP

908

909
C

910
AC

911

912

913

914

915

41
ACCEPTED MANUSCRIPT

916

917

918

0.4

PT
0.3

Tg: Half Cp Extrapolated = 128.3oC


Heat Floe (W\g)

RI
0.2

0.1

SC
0.0
Reference BFRP specimen
-0.1
40 60 80 100 120 140 160 180

U
919 Temperature (oC)
AN
920 Fig. 7. DSC scans for glass transition temperature (Tg)

921
M

922

923
D

924
TE

925

926
EP

927

928
C

929
AC

930

931

932

933

934

42
ACCEPTED MANUSCRIPT

935

PT
RI
SC
936

937 Fig. 8. Micrographs of the BFRP bars cross-section (20 mm-diameter)

U
938
AN
939

940
M

941

942
D

943
TE

944

945
EP

946

947
C

948
AC

949

950

951

43
ACCEPTED MANUSCRIPT

Variation in the basalt-fiber diameter

PT
RI
952

SC
953 Fig. 9. Micrographs of the basalt FRP bar cross section at medium magnification

954

U
955
AN
956

957
M

958

959
D

960
TE

961

962
EP

963

964
C

965
AC

966

44
ACCEPTED MANUSCRIPT

PT
RI
967

U SC
AN
M

968
D

969 Fig. 10. Micrographs of the bars for basalt fiber/matrix interface
TE

970

971
EP

972

973
C

974
AC

975

976

977

978

979

45
ACCEPTED MANUSCRIPT

PT
RI
980

SC
981 Fig. 11. Typical failure mode of tested BFRP specimens subjected to tensile test

982

U
983
AN
984

985
M

986

987
D

988
TE

989

990
EP

991
C
AC

46
ACCEPTED MANUSCRIPT

120
Transvers Shear Strength

Transvers shear strength retention


105 100
92.9
90 89
79.2
75

PT
(%) 60

45

30

RI
15

SC
0
0 1000 3000 5000
992 Exposure time (hours)

993 Fig. 12. Effect of conditioning in the alkaline solution at 60°C on transverse-shear strength

994
U
AN
995
M

996

997
D

998
TE

999

1000
EP

1001

1002
C

1003
AC

1004

1005

1006

1007

1008

47
ACCEPTED MANUSCRIPT

1009

1010

120 120
(a) Flexural strength (b) Flexural modulus of elasticity
105 100 105 100
Flexural strength retention (%)

98.3

Modulus of elasticityretention (%)


95.3

PT
90.4 90.6 88.1
90 90
81.0
75 75

RI
60 60

45 45

SC
30 30

15
15
0

U
0
0 1000 3000 5000
0 1000 3000 5000
Exposure time (hours) Exposure time (hours)
AN
1011

1012 Fig. 13. Effect of conditioning in the alkaline solution at 60°C on mechanical properties:
M

1013 (a) flexural strength; (b) flexural modulus of elasticity

1014
D

1015
TE

1016

1017
EP

1018

1019
C

1020
AC

1021

1022

48
ACCEPTED MANUSCRIPT

120

Interlaminar shear strength retention


Interlaminar Shear Strength
105 100
92.9
90 89.0
79.2
75
(%)

PT
60
45
30

RI
15
0

SC
0 1000 3000 5000
1023 Exposure time (hours)

Fig. 14. Effect of conditioning in the alkaline solution at 60°C on interlaminar-shear strength

U
1024 AN
1025

1026
M

1027

1028
D

1029
TE

1030

1031
EP

1032

1033
C

1034
AC

1035

1036

1037

1038

49
ACCEPTED MANUSCRIPT

1039

1040

1041

PT
105
y = -6.6739x + 111.08
Transverse Shear strength retention

y = -8.0951x + 112.01 22C 40C 60C

Transverse Shear strength retention


100 R² = 0.8229 100
R² = 0.8949
95
y = -9.4647x + 112.83
90

100 years
200 years
90 R² = 0.9049

RI
85
(%)

(%)
80 80
R10

SC
75
70
70 Log decade
65 (a) (b)
22C 40C 60C
60
60 1000 10000 100000 1000000

U
1.5 1.7 1.9 2.1 2.3 2.5
Log[Time (days)] Time (hours)
1042
AN
1043 Fig. 15. Plot of the transverse shear strength retention of BFRP bars as a function of time

1044
M

1045

1046
D

1047
TE

1048

1049
EP

1050

1051
C

1052
AC

1053

1054

1055

1056

1057

50
ACCEPTED MANUSCRIPT

1058

1059

105
Shear-Strength Retention (%)
100

PT
95
90
85
T = 10oC

RI
80
75
T = 22oC
70

SC
65
60
0 50 100 150 200

U
1060 Service Life (Years)
AN
1061 Fig. 16. General relation between the PR and the predicted service life at mean annual
1062 temperatures of 10 and 22°C
1063
M

1064
D
TE
C EP
AC

51
ACCEPTED MANUSCRIPT

Highlights

 Basalt FRP

 Durability of Basalt Fiber

 Long-Term performance of BFRP

PT
 Shear strength of BFRP

RI
 Tensile strength of BFRP

U SC
AN
M
D
TE
C EP
AC

Vous aimerez peut-être aussi