Sei sulla pagina 1di 28

Construction and Building Materials 232 (2020) 117152

Contents lists available at ScienceDirect

Construction and Building Materials


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

Review

Performance evaluation of Ultrahigh performance fibre reinforced


concrete – A review
Faiz Uddin Ahmed Shaikh a, Salmabanu Luhar b,⇑, Hasan Sß ahan Arel c, Ismail Luhar d
a
School of Civil and Mechanical Engineering, Curtin University, Perth, Australia
b
Institute of Mineral Resources Engineering, National Taipei University of Technology, Taipei, Taiwan
c _
Faculty of Architecture, Izmir _
University, Izmir, Turkey
d
Shri Jagdishprasad Jhabarmal Tibrewala University, Rajasthan, India

h i g h l i g h t s

 Fibers might be wielded for the fabrication of Ultrahigh performance fibre reinforced concrete.
 UHPFRC unveiling a higher compressive, tensile and flexural strengths.
 UHPFRC is having higher ductility as well as admirable durability.
 The article reviewed the mechanical properties, specimens size effect, as well as loading rate effect.
 The article also reveals the effects of fıbre– type, geometry, length, volume and orientation on concrete attribute.

a r t i c l e i n f o a b s t r a c t

Article history: Ultrahigh performance fibre reinforced concrete (UHPFRC) is a newfangled cement-based material
Received 17 June 2019 unveiling a compressive strength of more than 150 MPa, higher tensile and flexural strengths, ductility
Received in revised form 22 September as well as admirable durability. This review of the past researches on diverse influential features is vital
2019
in obtaining fundamental data for its viability for the application. Consequently, this study aimed to
Accepted 3 October 2019
review the available literatures on UHPFRCs that examine the attributes of ideal UHPFRC, such as
mechanical properties viz., compressive, tensile and flexural strengths; mixing proportions and environ-
ment friendliness; curing regimes; effect of specimens size on its compressive, flexural and tensile
Keywords:
Ultrahigh performance fibre reinforced
strength as well as loading rate; effects of fıbre properties – type, geometry, length, and volume fractions,
concretes (UHPFRC) orıentatıon and quantıty attrıbutes of UHPFRC is endeavoured along with relevant discussion.
Compressive strength Subsequently, the best UHPFRC mixture was determined to be obtainable with 2% to 3% of steel fibre con-
Impact strength tent and water/cement ratio of <0.2. Additionally, the UHPFRCs that were subjected to curing at 90 °C
Bond behavior yielded compressive, tensile, and flexural strengths that were 49% better than the samples cured at
Steel fibre 20 °C. The review elucidates the key points of producing environmental friendly UHPFRC material for
Tensile and flexural strengths future applications as the current UHPFRC contains about twice the amount of cement compared to ordi-
Supplementary cementitious materials
nary concrete.
Ó 2019 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2. Ideal UHPFRC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
3. Mix Proportion of UHPFRC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
4. Sustainable UHPFRC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
5. Curing Regimes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
6. UHPFRC Mechanical Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
6.1. Compressive Strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

⇑ Corresponding author.
E-mail address: ersalmabanu.mnit@gmail.com (S. Luhar).

https://doi.org/10.1016/j.conbuildmat.2019.117152
0950-0618/Ó 2019 Elsevier Ltd. All rights reserved.
2 F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152

6.2. Tensile and Flexural Strengths . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10


6.3. Impact Strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
7. Effects of specimen size on mechanical properties of UHPFRC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
8. Effects of fıbre geometry, length, and volume fractions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
9. Effects of fıber orıentatıon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
10. Effect of fıbre type . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
11. Effect of quantity of fıbres . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
12. Applications and prospective recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
13. Conclusions and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
Declaration of Competing Interest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

1. Introduction sile strength of >7 MPa (Fig. 2 illustrations the tensile stress-strain
curve), a maximum aggregate diameter of less than 1 mm, superior
Concrete is the utmost expansively employed building material ductility in tension and bending, high fracture toughness and low
owing to its widespread application and rewards despite its short- maintenance. Despite its low water/concrete ratio, it has high
comings, such as its moderately lower tensile strength or brittle- workability (200 mm) and retains its pump ability [14–17].
ness [1–4] and comprehensively employed construction UHPFRC has been an attractive alternative material for High eleva-
materials on the earth which is an impetus of constructional and tion buildings, pre-stressed girders as well as long span bridges
infrastructural development of any nation since Romans had initi- because of its superior compressive and tensile strengths and
ated its use as ‘‘Opus Caementicium” [5,6]. However, an exigency toughness performance [18,19]. Fig. 3 shows the SEM images of sil-
of various performances by concrete has escalated with the ica fume-substituted normal concrete (a), fibre-reinforced concrete
increasing demands of today’s construction and infrastructure (b), HPC (c), and ultra-HPFRC (d).
industries and has attained levels that traditional concrete cannot
satisfy [7,8]. High-performance concrete (HPC) is preferential to
2. Ideal UHPFRC
conventional concrete on account of high strength and durability
in engineering applications [9]. However, although it has greater
The strength of UHPFRC depends on three characteristics:
strength, its brittleness is known be higher than that of traditional
cement paste pore structure, quality of the aggregate, and struc-
concrete. The use of HPC dates to the 1970s. HPCs with 62-MPa
ture of the aggregate–matrix and fibre-matrix interfaces. The
compressive strength were used for the Water Tower Place in Chi-
weakest among these characteristics is the interface area, which
cago in 1974, HPCs with 69-MPa compressive strength were uti-
could be enhanced by reducing the water to cement ratio as well
lized for the construction of Taipei 101 in Taiwan in 2004, and
as custody the aggregate diameter below 1 mm. However, both
HPCs with a compressive resistance of 80 MPa were employed
methods have upper limits. Cementitious materials, such as fly
for the 126 floors of the Burj Khalifa in the United Arab Emirates,
ash, slag, and silica fume, not only lower the cost of production
Dubai in 2009 [10]. Despite its frequent use, several disadvantages
but also increase the strength [20,21]. The durability of high-
stand out. For instance, if high-performances and conventional
strength concrete is determined as functions of full volume and
concretes are subject to heat, high-performance concrete’s
the pore s. The quality of the aggregate directly affects the high-
strength loss is higher than conventional concrete [11]. Ultra-
strength concrete [22–27]. UHPFRCs are not generally used when
high-performance fibre-reinforced concrete (UHPFRC) is produced
traditional concrete is productive because of its elevated manufac-
by combining HPC and fibre at 3 vol% or above [12,13]. UHPFRC is a
turing price, providing the required output. In this regard,
form of concrete with a high compression intensity (150–200 MPa)
researchers pointed out three main factors that can be reduced
(Fig. 1 demonstrations a compressive stress-strain curve), a lower
to lower the cost without compromising the quality of the product,
water to concrete ratio of 0.2 or less, high bending strength, a ten-
namely, the high-strength fibre quantity added to the mix, the
powder quantity, and the curing [28–30]. For the production of

Fig. 1. Typical compressive strength – strain curve [44]. Fig. 2. Uni-axial direct tensile stress – elongation curve [45].
F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152 3

Fig. 3. SEM images of NC, FRC, HPC and UHPFRC concretes – a) [135], b) [136], c) [137,138], d) [139].

UHPFRC the mean particle size is frequently less than 1 mm, but with a grain size of 150 to 600 mm, (2) increasing the density by
aggregates are up to 8 mm to 15 mm. [31]. The reactive powder adding cement and silica fume, (3) improving ductility by adding
concrete (RPC) word can be used if the maximum total size is steel fibres, and (4) implementing temperature treatment (90 °C)
roughly 0.5 mm. The aggregate must be mechanically strength to to improve the UHPFRC microstructure and increase early age
ensure that the concrete is not fragile [32]. When coarse aggregates durability [34,40,41].
are added to the mortar, the cement content is decreased. coarse
aggregates are not often used in UHPFRC, but can be used if the
structural component size is greater than the aggregates thickness 3. Mix Proportion of UHPFRC
[33]. In UHPFRC, costly silica sands are usually used. However, nat-
ural sand can usually be substituted for silica sand and retain The raw materials of UHPFRCs and concretes with normal fibres
excellent mechanical efficiency and ductility. The use of natural are the same, except the superplasticizer, aggregates and the min-
sand does not have a significant impact on UHPFRC strength [34]. eral additives with different binding characteristics. The hydration
Depending on the concrete type and application, the fiber content process of the paste phase and the microstructure of the hydrated
can differ considerably. When it is important to workability attri- paste demonstrate variances because the water/cement or water/
bute, the fibers in the concrete are usually a mixture of a large binder ratios are low and high-range water reducers are employed
number of short fibers and a few lengthy fibres [35]. The fibers in UHPFRCs. To achieve the required UHPFRC output: (1) the
may be less than 12 mm and the complete quantity can be 11 per- cement fineness should be high; (2) the quantity of C3S should
cent [36]. Approximately 2.5% steel fibers with aspect ratio of 40 to be high; (3) C3A and C4AF quantities should be low; and (4) it
60 have however been demonstrated to produce the highest out- should have a cubic polymorph structure, and the amount of alkali
comes in terms of fresh and hardened concrete characteristics. and alkaline sulfates should be limited. The UHPFRCs collapse
The length of the fibers should be set to the maximum aggregate when aggregates crack. Thus, an aggregate with low durability
diameter in order to ensure low porosity. At least ten times the would cause the UHPFRC to have low durability as well. In
maximum overall diameter should be the length of the fiber. UHPFRCs, the high-range water reducer rapport with the cement
Unless high-performance plasticizers are added, the UHPFRC can- is highly important. Sufficient workability should be provided
not workable, i.e. up to 5 percent by weight of a cement. The aspect and not be exhausted in a short period. The purpose behind adding
ratio of fibers has a major impact on the fibers ’ facility to mix into steel fibres to the cement is to establish a crack control and resis-
the concrete and on the workability of the concrete. Workability tance to the pull-out force that the cement matrix cannot carry for
reduces in particular with the increase in the aspect ratio [37]. a long time after the peak load. Bending strength addresses the fra-
UHPFRCs are sometimes undergoing heat therapy in order to gility of the steel fibre, and the fibre amounts are the utmost
enhance the concrete quicker (compressive and tensile resistance), imperative features in determining the characteristics of UHPFRCs.
decrease shrinking time and creep impacts and considerably Preferably, the water/binder and water/cement ratios are main-
enhance the concrete’s durability [38]. Heat treatment begins to tained at approximately 0.16–0.2 in UHPFRCs [42,43]. The content
produce more hydrates, leading to enhanced features [39]. The req- of cement usually varies from 600 to 1000 kg/m3 having 3000 and
uisites to obtain maximum performance from UHPFRCs can be 4500 cm2/kg of fineness shall be used. Ordinary Portland cement
summarized into four: (1) improving the homogeneity using sand with low C3A content is suggested to use owing to its low water
Table 1

4
Research studies that examined the ideal mixture ratios of UHPFRCs at a wide range.

References Mix design (kg/m3) Fiber


[46] 875 CEM I 52.5 R, 44 micro silica, 218.7 Microsand (0–1 mm), 202.1 water, 45.9 Superplasticizer, 1054.7 Sand (0–2 mm), 0.23 w/c, Short straight steel fibers (length of 13 mm and diameter of 0.2 mm), 2.5
0.19 w/b (vol.%), Specific density 7800 kg/m3
[47] Proportions by weight 1.0 Portland cement, 0.3 Filler (crushed quartz with an average diameter of 10 lm and a density of 2600 kg/ steel fibers, Diameter of 0.2 mm, lengths of 16, Proportions by
m3), 1.1 Fine aggregate (sand with a diameter of < 0.5 mm), 0.02 Water-reducing admixture, 0.2 w/b Diameter of 0.2 mm, lengths of 19 (Used together at weight 0.25
the same time) (volume of 1.5%)
[55] (Proportions by weight) 1.00 CEM I 52.5R, 0.25 Silica fume, 0.25 Glass powder, 0.22 Water, 0.031 high-range water reducer (Sika Steel fiber, 13 mm long with a diameter of 0.15 mm 0.2 (Proportions by
SVC 20 Gold) , 0.019 high-range water reducer (Sika ViscoCrete 20He), 0.42 Fine sand (0.1/0.6 mm), 0.8 Fine sand (0.3/0.8 mm), and tensile strength of 2800 MPa weight)
0.176 w/b
[61] 800 CEM I 52.5R, 176 water, 40 high range water reducers, 336 fine sand (0.1/0.6 mm), 640 fine sand (0.3/0.8 mm), 200 silica Steel fiber Short high strength 13  0.15 mm 160 kg (%2 by
fume, 200 glass power volume)
[145] (proportions by weight) 1.00 Cement I 52.5 R, 0.22 water, 0.05 0.95 Fine sand (0.3/0.8 mm) (proportions by weight) Steel Fibers (13  0.15 mm) 0.05
high range water reducer, 0.42 Fine sand (0.1/0.6 mm), 0.25 Silica 0.90 Fine sand (0.3/0.8 mm), 0.1
fume, 0.22 w/c 0.85 Fine sand (0.3/0.8 mm) 0.15
0.80 Fine sand (0.3/0.8 mm) 0.2

F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152
0.75 Fine sand (0.3/0.8 mm) 0.25
0.70 Fine sand (0.3/0.8 mm) 0.3
[67] 657 CEM I 52.5 N, 418 Ground Granulated Blast Furnace Slag, 119 Silica fume, 1051 Silica sand, 59 Superplasticizers, 185 Water, steel fibers, 13 mm length and 0.16 mm diameter 3%
0.15 w/b
[146] 1050 CEM I 52.5 N, 730 Quartz sand (Diameter < 500 lm), 275 Silica Fume (spec. surf. 12 m2/g), 35 Superplasticizer, 0.14 w/b steel fibers, straight (10 mm, 0.2 mm) 470
[147] 768 OPC (Type 1, 42.5R), 1140 Mining sand (Diameter < 1180 lm), 192 Silica fume (23.7 m2/g), 40 Superplasticizer (PCE-based), Steel fiber (Lf = 10 mm, df = 0.2 mm), 157
144 Water, 0.15 w/b
[148] 825 OPC (Type 1, 42.5R), 1140 Mining sand (Diameter < 1180 lm), 181 Silica fume (23.7 m2/g), 45 Superplasticizer (PCE-based), Steel fiber (Lf = 10 mm, df = 0.2 mm), 90
174 Water, 0.17 w/b
[45] 1011 Blast furnace slag cement III/40, 58 Silicate fume, 79 Sílica flour (18 lm), 60 Aggregates (150–300 lm), 823 Aggregates (425– Steel fibers 12 mm long with diameter of 0.18 mm 158
600 lm), 76 Micro fibers de wollastonite, 50 Plasticizer PA, 162 Water, 3.75 Air (%), 0.19 w/b
[149] 657 CEM I 52.5 N, 418 Ground Granulated Blast Furnace Slag, 119 Silica fume, 1051 Silica sand (average size 0.27 mm), 40 steel fiber 13 mm length and 0.2 mm diameter 157
Superplasticizers, 185 Water, 0.15 w/b
[150] 657 CEM1 52.5 N, 418 Ground Granulated Blast Furnace Slag, 119 Silica fume, 1051 Silica sand (average size 0.27 mm), 40 straight high carbon steel with a tensile strength of %2 By volume
Superplasticizers, 185 Water 2000 MPa, 13 mm in length and 0.20 mm in diameter
[144] (Proportions by weight) 1.00 Cement PC Type I (3930 cm2/g Blaine value), 0.25 Silica fume (very low carbon content (0.3%) and a Steel fibers straight, lf/df = 13 mm/0.20 mm Proportions by
median particle size of 1.2 lm), 0.25 Silica powder (median particle size 1.7 lm) , 0.22 Water, 0.0054 Superplasticizer (Solid weight 0.25
content), 0.26 Fine sand 1 (max. grain size 0.2 mm), 1.03 Fine sand 2 (max. grain size 0.8 mm), 0.22 w/b
[88] 960 CEM II/A-L 42.5 R, 240 Silica fume (specific surface area of about 18 m2/g), 24 Superplasticizer, 960 (Sand well-graded very The steel fibers, 13 mm long and 0.18 mm thick with 192
fine natural sand was used with particle size up to 100 lm), 227 Water, 0.24 w/c, 0.19 w/b an aspect ratio of 72
[141] (Proportion of materials by weight ratio), 1.0 ordinary Portland cement, 0.25 silica fume, 0.3 filler, 1.1 fine aggregate steel fibers with a diameter of 0.2 mm and a length of 2.0%
(diameters<0.5 mm), crushed quartz with an average diameter of 10 lm, 0.02 Water-reducing admixture, 0.2 w/c 13 mm. yield strength of 2500 MPa.
[151] 1051 CEM I 52.5, 273 silica fume, 733 Sand (dmax = 0.5 mm), 165 Water, 35 Superplasticizer, Air content 4% of total volume, 0.14 Steel fibers (1/d = 50) 468
w/b
[152] 2349 Premix EIFFAGE B1M2, 195 Water, 50 Superplasticizer Glenium G51, 1.9 void ratio (%) Fibers metal 195
[44] 657 cement, 418 Ground Granulated Blast Furnace Slag, 119 Microsilica (silica fume), 1051 Silica sand, 40 Superplasticizers, 185 Steel fiber, fibers are of type OL13 with 2000 MPa 157 (2%)
Water tensile strength, 13 mm length and 0.2 mm diameter
[153] Proportions by weight, 1.0 Cement PC Type I, 0.25 Silica Fume, 0.25 Quartz Powder, 0.22 Water, 0.0054 Superplasticizer (solid Steel, straight, lf/df = 13 mm/0.2 mm 0.25, (2.5% by
content), 0.26 Fine sand A (max. grain size 0.2 mm), 1.03 Fine sand B (max. grain size 0.8 mm), volume)
[154] 1410.2 CEM I 52.5, 366 Silica Fume, 80.4 Fine Sand (maximum size of 0.5 mm), 200.1 Water, 44.1 Superplasticizer, 0.13 w/b microfibers and straight steel fibers (10 mm/0.2 mm) 706.5
[143] 2195 (Premix = Proprietary mixture designs, including inert and cementitious constituents), 30 High-range water-reducing 13 mm steel fibers, 0.2 mm in diameter (%2 Steel 156
admixture, 130 water fiber volumetric percentage)
2161 (Premix = Proprietary mixture designs, including inert and cementitious constituents), 29 High-range water-reducing 13 mm steel fibers, 0.2 mm in diameter (2.5% Steel 195
admixture, 128 water fiber volumetric percentage)
2296 (Premix = Proprietary mixture designs, including inert and cementitious constituents), 50 High-range water-reducing 20 mm steel fibers, 0.3 mm in diameter (2.5% Steel 195
admixture, 190 water fiber volumetric percentage)
[26] 1277.4 CEM III/B (66–80% high percentage of blast furnace slag), 95.8 Silica fume, 664.6 Sand, 42.3 Superplasticizer, 198 Water, W/ 235.5 Steel fibers (13/0.16 mm)
C = 0.155
[60] (proportions by weight), 1.00 Cement, 0.25 Silica fume, 0.25 Glass powder (Median particles size is 1.7 lm), 0.19 Water, 0.011 0.27 (proportions by weight) Fiber, Straight steel fiber,
Superplasticizer, 0.28 Sand A (Maximum grain size is 0.2 mm), 0.64 Sand B (Maximum grain size is 0.8 mm), length/diameter = 13 mm/0.20 mm
Table 1 (continued)

References Mix design (kg/m3) Fiber


[155] (Proportion by weight), 1.00 Cement 52.5 ordinary Portland cement, 0.41 high strength fine aggregate (size ranging from 1 to short straight steel fiber, 6 mm long, diameter of approximately 0.2 mm (The
3 mm), 0.3 Silica flour (size between 0.1 mm and 0.3 mm and surface area ranging from 20 to 28 m2/g), 0.32 Silica fume (average tensile strength and Young’s modulus for the steel fibers were reported to be
size of 0.04 mm), 0.19 water, 0.19 W/C 1.2 GPa and 210 GPa)
[68] 795.4 CEM I 52,5R HS-NA Cement, 168.6 sika silicoll uncompact, 24.1 sika viscocrete, 198.4 fine quartz, 971 sand quartz (0.125/ 79.31 steel fiber (1%)
0.5), 0.255 w/c, 0.210 w/b 160.25 steel fiber (2%)
[156] 1.00 Type 1 Portland cement, 0.25 Water, 0.25 Silica fume, 1.10 Sand (grain size lower than 0.5 mm), 0.30 Silica four (2 lm 2% Steel Smooth fiber, 0.2 Diameter, 13 Length, 2500 Tensile strength (MPa)
diameter, including 98% SiO2), 0.018 Superplasticizer, 0.2 w/b
[62] 1.00 Cement, 0.25 Water, 0.25 Silica fume, 1.10 Sand (grain size smaller than 0.5 mm), 0.30 Silica flour (2 lm diameter, including 2% Steel Smooth fiber, 0.2 Diameter, 13 Length, 200 Tensile strength (GPa)
98% SiO2), 0.018 Superplasticizer, 0.2 w/b
[157] 855 cement, 470 quartz sand (9–300 lm), 470 quartz sand (250–600 lm), 214 microsilica, 188 water, 28 Superplasticizer, 0.22 (390 kg, 0.16 mm diameter, 6 mm long), (78 kg, 0.16 mm diameter, 13 mm
water/cement, 0.18 water/binder used together at the same time
[64] Percentage by Weight (%),28.6 Cement, 9.3 Silica Fume, 8.5 Ground Steel Fibers, diameter of 0.2 mm and a nominal length of Percentage by Weight (%) 6.4
Quartz, 41.1 Fine Sand, 0.5 Superplasticizer, 5.6 Water 12.7 mm. Its yielding stress is 3150 MPa and ultimate stress
is 3250 MPa.
[158] Relative weight ratios to cement, 1.00 Type I Portland cement, 0.25 Water, 0.25 Silica fume (200,000 cm2/g Specific surface), 1.10 0.2 mm Diameter, 13.0 mm Length, 7.8 g/cm3 1%
Sand (grain size smaller than 0.5 mm and 2-lm-diameter), 0.30 Silica flour (98% SiO2), 0.016 Superplasticizer, 0.2 w/b Density, 2500 MPa Tensile strength 2%

F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152
3%
4%
[159] Ductal Premix (consist of Portland cement, silica fume, quartz sand with a maximum particle diameter of 1.2 mm and very fine Short straight steel fiber, diameter 0.2 mm, 15 mm 157 kg/m3, (%2)
powder composed mainly of quartz as the mineral admixture), w/c 0.22, 155 water, 25 superplasticizer long
[160] DuctalÒ (712 kg/m3 Portland Cement, 1020 Fine Sand, 231 Silica Fume, 211 Ground Quartz, 30.7 high-range water-reducing Short straight steel fiber 1%, 2%, 3%
admixture, 30 Accelerator, 109 Water), 0.22 w/c
[59] (Relative weight ratios to cement), 1.00 Type I Portland cement, 0.25 water, 0.25 silica fume, 1.10 sand (grain size smaller than 2% Steel fiber, 0.2 mm Diameter, 13.0 mm Length, 2500 MPa Tensile strength
0.5 mm), 0.30 silica flour (2 lm-diameter, 98% SiO2) , 0.016 superplasticizer, 0.2 w/b
[142] 654 type I Portland cement, 109 silica fume (specific surface area 2200 m2/kg), 109 fly ash (specific surface area 686 m2/kg), 218 Cold drawn low carbon steel fiber, 13 mm length, 2%
slag (specific surface area 766 m2/kg), 1090 sand, 32.7 superplastizer, 0.16 w/b 0.175 mm diameter, tensile strength 1800 MPa 3%
4%
[66] 880 Standard Portland pozzolan cement with a strength category of 52.5 N/mm2, 220 Undensified silica fume with 97% purity in 401 Brass Coated Steel Fibers 6 mm and 0.16 mm diameter, 80 Brass Coated
SiO2, 475 Calcareous Sand 125–250 lm, 358 Calcareous Sand 250–500 lm, 172 Water, 67 Polycarboxylate polymer based Steel Fibers 13 mm and 0.16 mm diameter used together at the same time
superplasticizer, 0.20 Water/cement, 0.16 Water/binder

5
6 F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152

Table 2
Curing regimes.

References Curing type


[46] The prisms are demolded approximately 24 h after casting and then cured in water at about 21 °C.
[54] The samples were de-molded after 24 h. Thereafter, the recommended curing treatment for the UHPFRC (steaming at 90 °C and 100% relative humidity for
48 h) was applied. The samples were then cured in water at 27 °C ± 2 °C until the testing day.
[34] The specimens were then covered with damp hessian and polythene sheeting. After one day, they were de-molded and cured in water either at 20 °C or at
90 °C. The heat cured specimens were stored in a hot water bath from an age of 1 day until 7 days. These specimens were then stored in air at room
temperature until testing. The 20 °C cured specimens were kept in a standard curing tank until testing.
[56] Thermal treatment for 5 days with a constant temperature of 90 °C, water curing for 5 days with a constant temperature of 20 ± 2 °C, Subsequently the
specimens were kept in the temperature 20 ± 2 °C and relative humidity 60 ± 5% until testing
[57] The concrete samples underwent 5 days of moist curing followed by dry curing under ambient conditions, with the concrete properties determined on the
day of the column tests
[47] The cylindrical test specimens were cured by steam curing at a temperature above 90 °C for 48 h, and then they cured at room temperature for 60 days
until testing
[145] Specimens were removed from molds after 24 h and subsequently placed in water tanks for another 27 days
[67] The specimens were cured in a steam curing tank at 90 °C for 3 days.
[146] The evolution of the relative humidity in the material was recorded in a sealed specimen under isothermal conditions at T = 20 °C. The relative humidity
was constant at 100% until 32 h after the addition of water and then decreased monotonically
[45] The specimens were covered with a damp cloth and a polythene sheet in their molds for 24 h before being demolded and cured for 28 days in a fog room
with a controlled environment of 100% relative humidity and 23 ± 1 °C.
[149] Specimens mixes were covered with damp hessian and polythene sheets, and kept at laboratory temperature 20 °C for 24 h. After all the specimens were
checked for initial set, demolding took place at approximately 24 h. All the specimens were placed in a curing tank in an elevated temperature of 90 °C for
the next 48 h and finally kept at laboratory temperature until the testing day
[150] All the samples were cured after casting under approximately 20 °C for the first 24 h, followed by a hot water curing tank at 90 °C for the next 48 h and
finally dry cured at laboratory temperatures until testing
[144] After casting, the specimens were covered with plastic sheets and stored at room temperature for 24 h. Then they were taken out of their molds and stored
in a water tank at 20 °C for an additional 25 days. All specimens were tested at the age of 28 days after being removed from water and left to dry in the
laboratory environment for 48 h.
[88] Three prismatic specimens (40 by 40 by 160 mm) were manufactured for each mixture and for each curing time in order to evaluate mechanical behavior
of the seven UHPFRC mixtures. they were soft cast in steel forms (vibrated for 30 s after casting), then wet cured at 20 °C (standard curing) for compressive
strength measurements.
[151] Cast in 50  200  500 mm3 molds. No thermal treatment was applied. After 3 days in sealed conditions, the specimens were removed from molds and
exposed to wet curing until 28 days (Table 2). Afterwards, specimens were stored in a room at 20 ◦C and 50% relative humidity. The specimens were
submitted to tensile tests after at least 1 month of storage and brought back to the room.
[105] No thermal treatment was applied. After 3 days under sealed conditions, the specimens were removed from molds and exposed to wet curing until
28 days.
[161] All test specimens were covered with plastic sheets just after concrete casting and cured at room temperature for the first 48 h prior to demolding. After
demolding, heat curing (90 ± 2 °C) was carried out for 3 days, and then the specimens were stored in the laboratory with room temperature until testing.
[162] The temperature evolution of the UHPFRC, extrapolated from semi-adiabatic tests, showed a dormant period of more than 24 h. The final adiabatic
temperature was 115 °C.
[147] The specimens were cured by steam curing for 48 h at a temperature of 90 °C and 100% relative humidity, and subsequently cured in water maintained at
room temperature of 27 ± 2C until the testing day.
[148] The specimens were left for one day after casting without curing, apply curing treatment includes steaming the UHPFC at 90 °C for 60 h.
[44] Specimens were then cast in molds under a frequency vibrator for<1 min. They were then covered with a polyethylene sheet and allowed to harden at
laboratory temperature (20 °C) for one day. All the specimens were then taken out of the molds and were placed in a special curing tank at 90 °C for 2 days.
[153] After casting, the specimens were covered with plastic sheets and stored at room temperature for 24 h. Then they were taken out of their molds and stored
in a water tank at 20 °C for an additional 25 days. All specimens were tested at the age of 28 days.
[143] Steam treatment curing at 90 °C (194°F) and 95% humidity for 48 h, Laboratory curing
[155] The specimens underwent a curing regimen. Placed the specimens in a constant 22 °C temperature and 100% humidity room, and performed demold 24 h
later. 48 h later, specimens were submerged in a water bath maintaining 90 °C for 24 h. Finally, put the concrete specimens to the autoclave, raised the
autoclave temperature from the room temperature to 200 °C within 2.5 h. Kept 1.1 MPa autoclave pressure and 200 °C temperature for a moment. Then,
naturally cooled down the specimens to the room temperature.
[156] After demolding, the specimens were steam cured at a high temperature of 90 ± 2 °C for 3 days and then stored at room temperature again until testing.
[62] All test specimens were demolded after 24 h and immediately sealed with aluminum adhesive tape. The shrinkage measurement began just after UHPFRC
casting and was performed in a room with a temperature of 23 ± 1 °C and a humidity of 60 ± 5%.
[65] After casting, the specimens are covered and stored at room temperature for 24 h. Afterward they are removed from their molds and stored in a water tank
at 20 °C for additional 25 days. All specimens are then tested at 28 days. About 48 h prior to testing the specimens are removed from water and left to dry
in the laboratory environment. About 24 h prior to testing, a spray coating is applied on the surface of the middle portion of each specimen for better crack
detection.
[163] The specimens were covered with plastic sheets and stored at room temperature for 48 h prior to demolding. Water curing with high temperature
(90 ± 2 °C) for 3 days after demolding was carried out. All specimens were tested in dry conditions for 21 days. Two to three layers of polyurethane were
sprayed on all surfaces of the specimens after drying to facilitate crack detection.
[157] Specimens were heated to a maximum temperature of 90 °C from the room temperature of about 25 °C. Each 24-h thermal cycle consisted of a gradual
temperature rise period of 30 min, followed by a dwell period of 8 h and ending with a slow temperature drop. This thermal cycle mimics a summer’s day
in arid hot climate when the temperature of exposed concrete surface is known to reach the chosen maximum temperature. The number of thermal cycles
varied from 0 to 90 cycles. After the requisite number of thermal cycles, the specimens were tested at room temperature.
[140] The mixes can be cured at ambient temperature for 28 days or at the elevated temperature of 90 °C for just 7 days with no noticeable difference in the
mechanical properties.
[64] Ice cubes are added during the casting to control the mixing temperature and working properties and are necessary if the environmental temperature is
higher than 25 ◦C.
[158] The test specimens were cured at room temperature for the first 48 h prior to demolding. After demolding, heat curing (90 ± 2 °C) was performed for
3 days.
[159] After their removal from the molds, all specimens were cured at 90 °C for 2 days
[142] The specimens were stored in the condition of 20 °C and 95% relative humidity for 24 h, and then removed from the molds and cured in the same condition
for 90 days. After curing, the hardened UHPCC were cut and ground into smooth cylinder specimens with thickness of 35 mm and diameter of 75 mm.
F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152 7

consumption. This is preferably due to the very small water/bind- UHPFRC has also been decreased because palm oil fuel ash has
ing proportion of UHPFRC. Silica fumes (SF) consist of very lower been added as part of a cement substitution by Aldahdooh et al.
particles (around 1/100) than cement particles. Its tiny size makes [47]. In their research, the palm oil fuel ash was used to replace
SF very effective as filler and improves the density of packing. A cement as 30%. Yu et al. [48] also used as a partial substitute of
large amount of silica fumes of about 10 to 30 percent of the cement as municipal waste bottom ash (20 wt%) in UHPFRC. Lam-
cement mass are required to fill the voids between cement propoulos et al. [49] developed UHPFRC containing 35% slag, 10%
particles. silica fume, 55% cement and 3% steel fibres which unveiled com-
pressive strength-164 MPa, tensile strength as well as strain capac-
ities are 12 MPa and 0.35%, respectively. Hasan et al. [50] and
4. Sustainable UHPFRC Hasan and Jones [51] also developed UHPFRC using similar mix
proportions of Lampropoulos et al. [49] however, with steel fibres
The mixture ratios, fibre characteristics, and material attributes content of 2% which revealed analogous compressive strength-
obtained from various studies of UHPFRCs are outlined in Table 1. 164 MPa but considerable inferior tensile strength of 7.4 MPa.
The table shows that silica fume is mixed with immense amounts Makita and Bruhwiler [26] reported fatigue behaviour of UHPFRC
of cement in most study (>80 percent) in the blend that makes made by low heat cement containing about 66–80% slag and
UHPFRC carbon footprint and its price very high. This high- 8.5% silica fume as addition to cement in their composite. Shaikh
performance material, which reduces its extensive implementation et al. [52] recently evaluated the impact of the different fly ash con-
in many nations in infrastructure, has the disadvantages of high tent of 20% to 50% as part of a cement substitution on the UHPFRC
price and high carbon footprint. A number of researchers have compressive strength and tensile stress strain behaviour. Their
soughed to assess the effects of different additional cementitious outcome showed approximately 4–14% less compressive strength,
materials (SCM), for instance, slag, palm oil ash, fly ash, as a approximately 14–19% less ultimate tensile strength, about 10–
replacement of cement in the UHPFRC binder to tackle the high 37% fewer ultimate tensile strain and about 22–48% less energy
carbon footprint and high costs of the UHPFRC. Mahmud et al. consumption during strain hardening than the control UHPFRC.
[44] A research indicated where slag was used to substitute cement Their findings indicate that their tensile strength is lower. Their
with 35 percent by mass. In this research, compression strength findings have shown that UHPFRC with 40 percent fly ash has a sig-
with approximately 150 MPa and uniaxial tensile strength with nificantly greater ultimate tensile strain, higher ultimate tensile
approximately 9 MPa are noted. Toledo-Filho et al. [45] reported strength, and energy absorption capacities than the one with 30
that a compressive strength of 160 MPa, 11 MPa tensile strength percent fly ash whereas the former has a slightly reduced CO2
and an ultimate tensile strain of 0.25%, in which 52% of blast fur- emission about 15% than the latter.
nace slag mixed to cement to produce UHPFRC. The effects of slag,
fly ash and limestone powder as partial (30% each) substitution of
Cement on mechanical characteristics of the UHPFRC were 5. Curing Regimes
assessed in other research by Yu et al. [46]. The results indicate a
decrease in UHPFRC compressive strength with 30% slag relative UHPFRCs might not display significant changes in terms of
to UHPFRC control and less decline in fly ash and limestone powder mechanical characteristics after 28 days. Although the lengthening
composites. In the event of flexural resistance similar findings are of the curing process yields a certain amount of increase in the
also noted. In another research the compressive strength of mechanical characteristics, the increase in the mechanical charac-

Fig. 4. Studies with compressive strength between 150 and 170 MPa.
8 F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152

Fig. 5. Studies with compressive strength between 170 and 190 MPa.

Fig. 6. Studies with compressive strength between 190 and 300 MPa.

teristics after one year is not significantly different from those of paced where the temperature gradually increased to around 85°
the 28-day-old samples. In fact, the samples that were cured in C to 90° C. The cement is maintained at relative moisture of almost
air and contained silica fume displayed time-dependent strength 100% over 1 to 2 days at this temperature. This therapy is imple-
retrogression. Two types of heat treatments are applied in mented after the concrete is settled [38]. The characteristics of
UHPFRCs [39]. (1) Autoclaving takes place at a moderate 65 °C the concrete material are affected by this kind of therapy owing
temperature because greater temperatures affect the threat of to more hydrates in the concrete. Subsequent to the treatment, cer-
delayed ettringites formation at elevated humidity. In particular, tain aspects of the concrete’s durability are improved, the long-
this method decreases the early setting time. (2) Concrete was term performance is better; shrinkage as well as creep are reduced.
F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152 9

Table 3
Compressive strengths and specimen sizes.

References Specimen size and Comp. Strength at 28 day (MPa)


[46] UHPFRC is cast in molds with the size of 40 mm  40 mm  160 mm and compacted on a vibrating table, 156 MPa
[34] The compressive strengths of the samples that were cured below 20 °C were not noted due to the yielded strengths being below 150 MPa
For the samples that were cured at 90 °C; 50 mm cubes, when using silica sand 170 MPa, fine ordinary sand type I 177 MPa, Fine ordinary sand II 175 MPa,
recycled glass cullet 153 MPa
[54] 100 mm concrete cubes (With the 637.9 cement, 173 MPa), (With the 700 cement, 175 microsilica fume, 175 MPa), (With the 700 cement, 250 microsilica
fume 176.9 MPa), (With the 850 cement, 265.5 microsilica fume, 160 MPa), (With the 850 cement, 212.5 microsilica fume, 173 MPa), (With the 850
cement, 159.5 microsilica fume, 170 MPa), (With the 850 cement, 175 microsilica fume, 150 MPa), With the (720.49 cement, 181.41 MPa)
[56] 70  70  70 mm cubes, thermal treated 181.2
[57] When 100 mm  200 mm cylinders, 4% steel fiber is used 153Mpa. When 6% steel fiber is used 165 MPa
[47] Cylindrical test specimens with a diameter of 100 mm and a height of 200 mm 166.9
[55] Flexural strength was evaluated on 40  40  160 mm prisms and compressive strength was determined on the halves of these prisms 151.7
following
[61] Cylinders with 100 mm diameter and height of 200 mm 151.7
[145] Cylinders were 200 mm high with diameter of 100 mm. When 0.5% fiber is used 154 MPa, when 1% fiber is used 149 MPa, when 1.5% is used 149 MPa,
when 2% fiber is used 164 MPa, when 2.5% fiber is used 159 MPa, when 3% fiber is used 148 MPa
[67] Cube compressive tests (100 mm side) 164
[146] Compressive strength was determined on cylindrical specimens (£ 11 cm, length 22 cm) 168
[147] 100 mm cube 170
[148] 100 mm cube 172
[45] 50  100 mm cylindrical specimens 156
[149] 100 mm cube 151
[150] 100 mm cube 164.3
[144] 100 mm cube, Steel fibers straight, lf/df = 13 mm/0.20 mm, Proportions by weight steel fiber 0.25 201
[88] prismatic specimens (40 by 40 by 160 mm) 156
[141] cylindrical specimens with a diameter of 100 mm and a height of 200 mm 194
[151] cast in 50  200  500 mm3 molds 168
[152] Nine standard 110  220 mm control cylinders were cast 170
[164] The compressive strengths were determined on 11/22 cm cylinders. 192.4
[105] Cast horizontally in 50  200  500 mm3 molds. 168
[161] Cylindrical specimens with a dimension of u 100  200 mm were used. Relative weight ratios to cement is 0.
When shrinkage reducing admixture was used 200.14 MPa, when 0.01 shrinkage reducing admixture was used 186.75 MPa, when 0.02 shrinkage reducing
admixture was used 194.85 MPa
[162] Cylinders (16 cm, length = 32 cm) 168
[44] Cylindrical, 50 mm diameter  100 mm 150
[153] Prism shaped specimens with a slenderness of 2 (50  50  100 mm) 201
[154] 11/22 cm cylinders 190
[143] All of these values were calculated from tests on cast cylinders, with cylinders with a diameter of 76 and 110 mm, The result of steam curing at 2195 kg
premix was 220 MPa, the result of lab curing at 2195 kg premix was 192 MPa, the result of steam curing at 2161 kg premix was 212 MPa, the result of
2296 kg premix was 213 MPa
[26] prisms with a section of 40 mm  40 mm and a length of 160 mm were cast 56 daily
217
[60] 250
[155] 150 mm cube 300
[68] 1% fiber, a diameter of 100 mm and a height of 100 mm (cube) 176.27
1% fiber, a diameter of 100 mm and a height of 200 mm (cylinder) 178.28
2% fiber, 100 mm  100 mm  100 mm (cube) 178.03
2% fiber, 100 mm  100 mm  100 mm (cylinder) 178.35
[165] The length and width of the tested tunnel lining segments were 1000 mm and 500 mm respectively, while their thickness was 100 mm. The
crown height of the tested segment was 100 mm from the horizontal surface. This was intended to represent a segment taken from a tunnel
having a diameter of 1.9 m.
When 8 mm length fiber is used 1% result is 156 MPa, when it is used 3% result is 164 MPa, when it is used 6% result is 171 MPa
When 12 mm length fiber is used 1% result is 158 MPa, when it is used 3% result is 166 MPa, when it is used 6% result is 173 MPa
When 16 mm length fiber is used 1% result is 159 MPa, when it is used 3% result is 165 MPa, when it is used 6% result is 170 MPa
[63] three cylindrical specimens of diameter 100 mm and length 200 mm were prepared and tested according to ASTM C 39 197.3
[156] Three cylindrical specimens with a dimension of u 100 200 mm 201.8
[62] cylinder u 100  200 mm 196.4
[157] 0 thermal cycle 200 MPa, 30 thermal cycle 224 MPa, 90 thermal cycle 225 MPa
[140] When 0.5% of 6 mm and 2.5% of 30 mm fibers are used together 167.1 MPa, when 1.0% of 6 mm and 2.5% of 30 mm fibers are used together 171.7 MPa,
when 1.5% of 6 mm and 2.5% of 30 mm fibers are used together 176.9 MPa, when 2.0% of 6 mm and 2.5% of 30 mm fibers are used together 182.4 MPa
[58] The compressive test was carried – out on cube specimens (100 * 100 * 100 mm) and cylinders with (100 mm diameter and 200 mm height) after 28 days
curing (1%, length of 30 mm and a diameter of 1.0 mm and steel fibers, cube 151 MPa, cylinder 137 MPa), (1%, length of 50 mm and a diameter of 1.0 mm
and steel fibers, cube 154 MPa, cylinder 141 MPa), (2%, length of 50 mm and a diameter of 1.0 mm and steel fibers, cube 163 MPa, cylinder 147 MPa), (2%,
length of 30 mm and a diameter of 1.0 mm and steel fibers, cube 160 MPa, cylinder 145 MPa), (3%, length of 50 mm and a diameter of 1.0 mm and steel
fibers, cube 169 MPa, cylinder 154 MPa), (3%, length of 30 mm and a diameter of 1.0 mm and steel fibers, cube 166 MPa, cylinder 151 MPa)
[64] The compressive tests were conducted on 7 cm diameter, 14 cm long cylinders with both ends ground. Mean strength of 228 MPa was recorded, giving a
characteristic value of 197 MPa with 95% confidence. The design strength is set at 180 MPa accordingly.
[158] Dimension of 100  200 mm, with 1% fiber 195 MPa, with 2% fiber 200 MPa, with 3% fiber 209 MPa, with 4% fiber 183 MPa
[159] 214.7 MPa
[160] Cylinders 100  200 mm. With 1% fiber 154.8 MPa, with 2% fiber 162.4 MPa, with 3% fiber 158.7 MPa
[156] Dimension of 100  200 mm, 152 MPa
[142] Dimension of 100 mm  100 mm  100 mm. With 2% fiber 151 MPa, with 3% fiber 177 MPa, with 4% fiber 194 MPa
[66] 173.0 MPa
10 F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152

be between 700 and 1050 kg/m3, and that of superplasticizer


should be between 30 and 50 kg/m3. The water/binder ratio should
be < 0.18 and the water/cement ratio < 0.2. The samples should be
subjected to 90 °C heat curing. Quartz should be used as the sand
type (specific surface area, 20,000–25,000 m2/kg), and the steel
fibre volume should be between 2% and 4%. The preferred sand
type should have a homogenous matrix. Thus, internal structure
and geometric shape are the determinants of compressive strength
and are necessary to avoid aggregate types, such as bottom ash or
recycled glass.
Yang et al. [34] reported that fine sand aggregate is an impor-
tant factor in compressive strength of UHPFRC and recycled glass
cullet due to its weak internal structure and its decreasing effect
on the compressive strength compared to ordinary fine sand. Yu
et al. [47] reported that when approximately 20% of quartz powder
is used instead of cement, the hydration of the cement and the
compressive strength improves due to the nucleation effect of
the fine particles. Aldahdooh et al. [54] discovered that the capil-
lary porosity increases with the amount of binder. They reported
that compressive strength can decrease due to the bleeding and
segregation that occur with the usage of an over the optimum dose
of superplasticizer. Lim and Hong [55] reported that the compres-
sive strength after 28 days of curing without using fibre is 2.5% and
6% higher than those using polypropylene and steel fibres, respec-
tively. The crack that occurs due to the compressive strength is
shown in Fig. 7.
Fig. 7. UHPFRC sample after the compressive strength experiment– [39,140]. Magureanu et al. [56] informed that the highest compressive
strength is obtained through the samples made of hybrid fibres
and 4% nano-silica mixture. They also suggested that polypropy-
The heat curing is an expensive and energy-consuming approach to lene fibre do not contribute to the compressive strength that
improve the material characteristics of UHPFRC. It restricts causes a statistically significant change. Aoude et al. [57] noted
UHPFRC manufacturing to precast sector, and thus restricts the that the application of steel fibers has an impact on elastic modulus
use in the cast in situ implementation of this material [34]. Devel- and compressive strength pre-cracking values, but has an signifi-
oping UHPFRCs without heat treatment or pressure treatment cant impact on post-cracking and failure mechanisms. Eldin et al.
would encourage use of the product, but because of all the influ- [58] described that compressive strength is higher in the samples
encing parameters it was a long time quite difficult [53,54]. Table 2 containing 3% fibre than those with 1% and 2%; the samples that
represents the curing regimes that were applied in various are produced with aspect ratio of 50, have higher compressive
UHPFRCs in the literature. It can be seen that about 2/3rd of the strength than those with 30. Othman and Marzouk [59] reported
studies presented in that table applied heat treatment during first that increasing the fibre content of UHPFRCs to 3% further enriches
few days of their curing periods. the impact strength compared to 1% and 2%.

6.2. Tensile and Flexural Strengths


6. UHPFRC Mechanical Properties
For UHPFRCs, tensile and flexural strengths are also essential
6.1. Compressive Strength mechanical characteristics along with the compressive strength
[33]. The compatibility of the fibres with the matrix and their sizes
According to Association Française de Génie Civil (AFGC)’s rec- influence tensile behavior as shown in Fig. 26. Table 4 shows the
ommendations [39], cylinders with sizes of 7  14 cm or tensile strength values in accordance with the sample sizes, fibre
11  22 cm should use to determine UHPFRC compressive content, and type in the studies reviewed. Table 5 shows the flex-
strength. In accordance with Eurocode 2, the compressive strength ural strength values.
can also be measured on cubes, provided that the coefficient for Steel fibres have an important role in the flexural and tensile
switching from cylinders to cubes is validated in design or test strengths of UHPFRC. While samples type ‘‘dog-bone” were mostly
suitability. The compressive strength as well as ration of water to used to measure the tensile strength (Fig. 8 shows the samples of
cement relations of the studies that yielded values of 150– the tensile strength experiments of Graybeal and Baby [60]), differ-
170 MPa, 170–190 MPa, and 190–300 MPa are depicted in Figs. 4– ent sizes of prisms were used in flexural strength measurement
6, respectively. The relationship between the specimen size and (Fig. 9 shows the 40  40  160 mm prismatic sample used in
the compressive strength after curing of 28 days is shown in the flexural strength experiment of Magureanu et al. [56]). When
Table 3. the studies were evaluated, the highest tensile strengths were
Table 3 and Figs. 4–6 demonstrate that with the decrease of the observed and showed the following hierarchy: twisted fibres > -
ratio of water to cement, the compressive strength increases. The long smooth fibres > hooked fibres > straight fibres. The common
increase of the cement content with the compressive strength does data show that the samples of twisted fibres with 2% to 3% cured
not yield a positive trend. The main reason behind this situation is at 90 °C yielded higher flexural and tensile strengths compared
assumed to be the increase in capillary porosity. This situation is to the other fibres. The addition of 1% to 3% percent of shrinkage
not dependent on one variable when the studies reach >170 MPa reducing additive to the mixture increased the flexural strength
of compressive strength. The total amount of binder should be values further. Flexural strength is directly affected by
between (cement + SF) 850 and 1000 kg/m3, that of sand should the homogenous spread of the aggregate in the mixture and the
F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152 11

Table 4
Tensile strengths and specimen sizes.

References Specimen size Tensile Strength at 28 day (MPa)


[56] 100  100  100 mm cube samples (thermal treatment, 5 days, 90 °C, 20.4 MPa), (water curing regime, 5 days,
water, temperature 20 ± 2 °C, 12.6 MPa)
[65] Shape dog-bone, Length 178 mm, width 51 mm and depth of constant area (when there is 22% Straight steel 15 MPa), (when there is 2.5% Straight steel
25 mm 16.5 MPa), (when there is 3% Straight steel 17.8 MPa), (when there is 1.5%
Hooked steel 12.4 MPa), (when there is 2% Hooked steel 14.7 MPa), (when
there is 3% Hooked steel 19.3 MPa), (when there is 1.5 Twisted steel
11.1 MPa), (when there is 2% Twisted steel 14.2 MPa), (when there is 3%
Twisted steel 19.6 MPa)
[163] The section of tensile specimens used is 50  100 mm, the gage length of When long smooth, steel is used 0.5% 11.419 MPa, when used 1.0%
the specimens is 175 mm 13.305 MPa, when used 1.5% 13.217 MPa
When hooked A uses 0.5% 10.895 MPa, when uses 1.0% 12.249 MPa, when
uses 1.5% 13.842 MPa
When hooked B, uses 0.5% 10.310 MPa, when uses 1.0% 11.331 MPa, when
uses 1.5% 12.014 MPa
When twisted uses 0.5% 13.498 MPa, when uses 1.0% 14.772 MPa, when
uses 1.5% 18.560 MPa

[144] It is pointed out that the small dimensions of the specimen in height When straight steel, diameter 0.2 mm, length 13 mm, tensile strength
(25.4 mm) and width (50.4 mm) 2600 MPa, uses 1.5% 8.3 MPa, when uses 2% 11.3 MPa, when uses 2.5%
14.2 MPa
When hooked steel, diameter 0.38 mm, length 30 mm, tensile strength
2900 MPa, uses 1% 9.4 MPa, when uses 1.5% 11.7 MPa, when uses 2%
14.0 MPa
When high twisted steel, diameter 0.3 mm, length 30 mm, tensile strength
2100 MPa uses 1% 8.0 MPa, when uses 1.5% 11.6 MPa, when uses 2%
14.9 MPa
When low twisted steel, diameter 0.3 mm, length 30 mm, tensile strength
3100 MPa uses 1.5% 3.3 MPa
[157] – Indirect tensile strength, 0 thermal cycle 27 MPa, 30 thermal cycle 31 MPa,
90 thermal cycle 31 MPa
[46] – When 0.5% of 6 mm and 2.5% of 30 mm fibers used together 21.1 MPa, 1.0%
of 6 mm and 2.5% of 30 mm fibers used together 23 MPa, 1.5% of 6 mm and
2.5% of 30 mm fibers used together 23.7 MPa, 2.0% of 6 mm and 2.5% of
30 mm fibers used together 26.6 MPa
[58] Cylindrical specimens with diameter 100 mm and height 200 mm were cast (1%, length of 30 mm and a diameter of 1.0 mm and steel fibers, 12.73)
(1%, length of 50 mm and a diameter of 1.0 mm and steel fibers, 15.4 MPa)
(2%, length of 50 mm and a diameter of 1.0 mm and steel fibers, 17.45 MPa)
(2%, length of 30 mm and a diameter of 1.0 mm and steel fibers, 15.59 MPa)
(3%, length of 50 mm and a diameter of 1.0 mm and steel fibers, 19.10 MPa)
(3%, length of 30 mm and a diameter of 1.0 mm and steel fibers, 17.82 MPa)
[164] Un-notched dog-bone specimens of 70 cm length and 10  5 cm2 cross Age 90, Uniaxial tensile strength 14 MPa
section
[146] The tensile properties were determined with a uniaxial tensile test on 11 MPa
notched plates having a cross-section of 20*5 cm2 and a reduced cross-
section of 16*5 cm2 at the notches
[143] The limit of the hydraulic wedge grip mouth opening led to the selection of 2% steel fiber at Long specimen and steam curing is 11.20 MPa,
a prismatic specimen with a 50.8 mm square cross section for all tests. The 2% steel fiber at Short specimen and steam curing is 10.29 MPa,
tapered aluminum plates affixed to two sides of each end of each specimen 2% steel fiber at Long specimen and lab. curing is 9.18 MPa,
were nominally 4.76 mm thick and linearly tapered to 1.0 mm thick over a 2% steel fiber at Short specimen and lab. curing is 8.56 MPa,
50.8 mm length. Two different specimen lengths, with corresponding 2.5% steel fiber at Long specimen and steam curing is 11.56 MPa,
changes in instrumented gauge lengths, aluminum plate dimensions, and 2.5% steel fiber at Short specimen and steam curing is 11.36 MPa,
grip lengths, were tested within the program. ‘‘Long” refers to a 431.8 mm 2.5% steel fiber at Short specimen and lab. uring is 10.53 MPa
total length prism, while ‘‘short” refers to a 304.8 mm total length prism. In
all cases, the specimens were single-point cast in prismatic molds, allowing
the UHPFRC to flow along the length of the form.
[64] 196.4 cm  4 cm  16 cm prisms The corrected characteristic tensile strength is 8.1 MPa
[54] To measure the direct tensile strength, dog bone-shaped specimens with a In the one with 720.49 cement, on the 28th day, the average strength still
testing section length of 80 mm and cross section of 16 mm  30 mm were increased by up to 12.49 MPa.
used.
[47] Test specimens had an overall width of 125 mm, a height of 300 mm, and a 11.5 MPa
thickness of 25 mm, but an effective width and a height are 75 and 150 mm,
respectively.
[56] The splitting tensile strength was measured on 100  100  100 mm cubes Thermal treated (5 days, 90 °C) 20.4 MPa, water curing regime (5 days,
water, temperature 20 ± 2 °C) 12.6 MPa
[55] Direct tensile tests were carried out on dog-bone shaped specimens 10.4 MPa
without a notch. The length of the specimens was 330 mm and the cross-
section of the narrowed part was 30  30 mm.
[145] Uniaxial tensile strength was determined on dog-bone shaped specimens When 0.5% fiber is used 7.5 MPa, when 1% fiber is used 7.8 MPa, when 1.5%
where the cross-section of the gauge section had a square shape with a side is used 9.9 MPa, when 2% fiber is used 164 MPa, when 2.5% fiber is used
length of 30 mm. Length of the gauge section was 80 mm and total length of 8.9 MPa, when 3% fiber is used 10.9 MPa
the dog-bone specimen was 330 mm.

(continued on next page)


12 F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152

Table 4 (continued)

References Specimen size Tensile Strength at 28 day (MPa)


[67] direct tensile tests of 6 dog-bone specimens with dimensions of 40 mm, The experimental results indicate a variation of the tensile strength
150 mm and 13 mm were carried out between 11.74 MPa and 14.20 MPa. An average stress–strain curve was
calculated and the average strength was found equal to 12 MPa

[45] for the uni-axial tensile behavior 50  12x 200 mm prisms An average first-crack stress of 9.2 MPa and was reached at a deflection of
about 0.026 mm. The maximum post-cracking stress was about 11.1 MPa
and was achieved at an elongation of about 0.213 mm. Note that the
elongation at maximum stress is about 8 times higher than that observed at
first crack.
[149] two unnotched dog-bone specimens of slightly different geometries, each 9.07 MPa
with an overall length of 200 mm, were prepared. The cross section of the
specimens starts with 50  50 mm and changes to a prismatic shape of
26  50 mm after either 25 or 50 mm away from each ends of the
specimens

[162] cast in 50  200  500 mm3 moulds 11 MPa


[158] Relative weight ratios to cement, when 0 shrinkage reducing admixture is
used is 13.38 MPa, when 0.01 shrinkage reducing admixture is used is
12.5 MPa, when 0.02 shrinkage reducing admixture is used is 11.2 MPa

[162] A uniaxial tensile test is developed for UHPFRC. Prismatic specimens with a In tension the UHPFRC was characterized by a nearly linear-elastic stress
cross-section of 20  5 cm2 are built-in the testing machine by applying the increase until the first cracking strength 9.1 MPa (at 28 days), followed by
principle ‘‘gluing without adherence” strain-hardening until a strain of 0.28% at the tensile strength 11.0 MPa (at
28 days).
[44] 9 MPa
[60] 18 MPa

[165] The length and width of the tested tunnel lining segments were 1000 mm When 8 mm length fiber is used 1% 17.6 MPa, when used 3% 21.9 MPa,
and 500 mm respectively, while their thickness was 100 mm. The crown when used 6% 39.8 MPa
height of the tested segment was 100 mm from the horizontal surface. This When 12 mm length fiber is used 1% 15.8 MPa, when used 3% 16.6 MPa,
was intended to represent a segment taken from a tunnel having a diameter when used 6% 17.3 MPa
of 1.9 m. When 16 mm length fiber is used 1% 13.9 MPa, when used 3% 18.9 MPa,
when used 6% 33.8 MPa
F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152 13

Table 4 (continued)

References Specimen size Tensile Strength at 28 day (MPa)


[63] 11.4 MPa

[62] Direct tensile test, A dog-bone shaped specimen was fabricated with a cross 7.4 MPa
section of 50  100 mm in the middle

[160] Cylinders 100  200 mm When there is 1% fiber 7.3 Pa, when there is 2% fiber 11.1 MPa, when there
is 3%fiber 14.0 MPa

Table 5
Flexural Strengths of Studies.

Researchers Specimen size Flexural Strength at 28 day (MPa)


[54] For the bending tensile strength measurement, four-point bending tests In the one with 720.49 kg cement 30.31 MPa
were conducted. The beams used to measure the bending tensile strength
had a length of 279 mm, cross section of 51 mm  51 mm, upper loading
span of 76 mm, lower support span length of 229 mm, and crosshead
loading rates of 0.1 mm/min.
[56] The flexural strength was investigated by performing a 3-point bending In the one with dimensions of 40  40  160 mm 22.30 MPa, in the one
test using 40  40  160 mm and 100  100  300 mm prismatic with dimensions of 100  100  300 mm 16.6 MPa
specimens. The specimens were thermal treated
[55] Flexural strength was evaluated on 40  40  160 mm prisms 40.1
[45] 100  12x 400 mm plate specimens From the results we obtained, a mean bending tensile strength at the first
crack of 18.2 MPa at a deflection of about 0.08 mm. The maximum post-
cracking strength reached at a maximum bending strength level of
23.0 MPa at a deflection of about 0.4 mm. Furthermore, a four-point
bending test for 12 mm thick specimen and span of 300 mm showed a peak
capacity of 35.0 MPa and for a specimen thickness of 100 mm the capacity
was 23.0 MPa.
[143,144] lf/df = 13 mm/0.20 mm, Proportions by weight steel fiber 0.25, 30 MPa
[34] 50  50  200 mm prisms The compressive strengths of the samples that were cured below 20 °C
were not noted due to the yielded strengths being below 150 MPa. Because
it does not provide an opportunity to compare the flexural strength values
of the samples that were cured at 20 °C’ were not written
In the samples that were cured at 90 °C’ de; when using silica sand 24 MPa,
fine ordinary sand type I 23 MPa, Fine ordinary sand II 25 MPa, recycled
glass cullet 21.5 MPa
[88] Three prismatic specimens (40 by 40 by 160 mm) 38 MPa
[141] A three-point bending test, The prism specimen had a height of 100 mm, a 30.5 MPa
width of 100 mm, a span of 300 mm and a length of 400 mm
[161] three-point flexure test was performed using 100  100  400 mm sized Relative weight ratios to cement, when using 0 shrinkage reducing
beam specimens with a 10-mm notch at mid-length admixture is 36.72 MPa, when using 0.01 shrinkage reducing admixture is
38.39 MPa, when using 0.02 shrinkage reducing admixture is 31.88 MPa
[44] Fifteen beams of similar geometry with depth varying from 30 mm to In the samples with a depth of 30 mm it is 22.32 MPa, with a depth of
150 mm are tested under 3-point bending, width b = 150 mm and span 60 mm it is 21.00 MPa, with a depth 90 mm it is 20.48 MPa, with a depth of
l = 500 mm 120 mm it is 20.11 MPa, with a depth of 150 mm it is19.76 MPa
[63] 100  100 mm and length of 400 mm were fabricated by placing concrete 31.6 MPa
at the corner of the beam. three-point flexure tests were performed as per
JCI-S-002–2003 for the beams with a 30-mm notch at the mid-length
[156] three-point bending test was performed according to a previous study. 32.6 MPa
Three 100  100  400-mm-sized beams with a 10-mm notch at the mid-
length were fabricated and tested.
[48,157] 100  100  500 mm beam up to 30 MPa
[58] 100  100  500 mm prism (1%, length of 30 mm and a diameter of 1.0 mm and steel fibers, 16.8 MPa)
(1%, length of 50 mm and a diameter of 1.0 mm and steel fibers, 18.4 MPa)
(2%, length of 50 mm and a diameter of 1.0 mm and steel fibers, 20 MPa)
(2%, length of 30 mm and a diameter of 1.0 mm and steel fibers, 19.2 MPa)
(3%, length of 50 mm and a diameter of 1.0 mm and steel fibers, 22.4 MPa)
(3%, length of 30 mm and a diameter of 1.0 mm and steel fibers, 21.6 MPa)
[156] Three-point flexure test, 100  100  400 mm sized beam specimens with When there is 1% fiber 23 MPa, when there is 2% fiber 33 MPa, when there
a 10 mm notch at mid-length were fabricated and tested. is 3% fiber 41 MPa, when there is 4% fiber 44 MPa
[159] 40 MPa
[160] Three point flexural strength test, 100  100  400 mm prisms with a clear span of 300 mm, when there is 1% fiber is 8.5 Pa, when there is 2% fiber is
19.2 MPa, when there is 3% fiber is 28.3 MPa
14 F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152

Fig. 9. The 40*40*160 mm prismatic sample used in the studies of Magureanu et al.
[142].

Fig. 10. Flexural strength – fiber content correlation. Yo et al. [143].

compared to UHPC without steel fibre, and by 67 percent for


UHPFRC with a 3 percent stainless steel fibre fiber quantity. The
UHPFRC samples have a tensile strength 2.2 times greater than
those in the UHPC, reported by Magureanu et al. [56]. Wille et al.
[65] reported that increased use of smooth fibers partially miti-
gates the need for additional mechanical bonds in hooked and
twisted fibre.

6.3. Impact Strength

The impact strength of concrete leads to elevated localized


strain rates. High strain rates trigger enhanced tensile and com-
pressive strengths (See Table 6). UHPFRCs has high energy absorp-
tion capacity and, with their high-strain, has the ability to control
structural defects [37]. Several studies examined the behaviors of
Fig. 8. Tensile strength experiment samples of Graybeal and Baby’s studies [141].
UHPFRCs against dynamic impact effect when concrete is replaced
porosity and fibre content (as the fibre quantity increased, flexural by fly ash, silica fume and slag, as well as when the concrete is
strength also increased). Fig. 10 shows the link between fibre and mixed with different fibre contents [66]. Máca et al. [61] suggested
flexural strength. The type of sand, rather than the amount of sand that high-strength fibres improve the impact behavior in terms of
used, resulted in significant changes in the flexural strength values, penetration depth of UHPFRC compared to normal concrete. How-
and the best results were obtained with quartz sand compared to ever, they reported that fibres that are added with 1% have no sig-
fine ordinary sand type I. nificant effect. They determined that the optimum ratio is 2%, and
Magureanu et al. [56] indicated that UHPFRCs exhibit higher fibres that are added in this ratio improve the impact load resis-
flexural strength compared to UHPCs by 150% to 165%, the speci- tance. Fig. 11 shows the impact effect of slabs that were applied
men sizes have an effect on flexural strength, and smaller-sized in the examinational studies by Maca et al. [61]. Researchers
specimens yield higher flexural strengths. Máca et al. [61] reported unearthed that a bullet could not penetrate into the UHPFRC slabs
that the highest flexural strength is obtained when a 3% fibre vol- and hence, it bounces back. Sovják et al. [67] reported that having
ume is used. They also indicated that specimens exposed to 90 °C more than 2% of steel fibres in the UHPFRC mixtures have no pos-
temperature have higher flexural strength than those exposed to itive effect in terms of penetration depth. Feng et al. [68] indicated
20 °C. Yoo et al. [62] reported that bending-related deflection that short and straight fibres have a positive effect on impact resis-
decreases with the increase in shrinkage reducing admixture con- tance. Zhang et al. [66] suggested that steel fibres have a significant
tent, and the highest benefit is obtained when 1% of the shrinkage- effect on dynamic strength, deformation, and energy absorption.
reducing admixture is used. Mahmud et al. [44] indicated that until The presence of these fibres improves dynamic strength, energy
a depth of 150 mm, the beams do not have a significant effect on absorption, and deformation.
the flexural strengths. Yoo et al. [63] claimed that unlike conven-
tional concrete and FRC beams, only very thin flexural microcracks 7. Effects of specimen size on mechanical properties of UHPFRC
occur in beams of the UHPFRC up to the peak load owing to strain
hardening features. Farhat et al. [64] specified that the observed There exist two most extensively adopted rules for research
flexural failures were typically characterized by a single flexural investigations on ‘‘size effect” for researchers. A deterministic and
crack in the middle third of the beam extending upwards to the powerful law for size effect was suggested by Bazant [69] which
top fiber of the concrete between the load points as well as down- states that larger specimens release more stored energy to the front
wards via the repair material. Eldin et al. [58] reported that steel of the fracture than narrower samples. Furthermore, for the first
fibres have a very important effect on flexural strength, and sam- moment Weibull [70] implemented a Statistical Size Effect law,
ples that have steel fibres show 40% higher flexural strength than which states that bigger samples are much more likely to fail than
those without steel fibres. Also, the same authors reported further lower samples. The size impact of concrete is therefore quite impor-
that fibres have a very important effect on tensile strength and the tant if material evaluation results for the structural layout are
samples that contain fibres have 55% more tensile strength. agreed upon because the sizes of the specimen used for material
The report showed a rise in tensile strength by 34 percent in the testing and structural components are generally different. With
UHPFRC, with a 1 percent stainless steel fibre fiber quantity, that in mind, numerous researchers [43,70–75] have investigated
F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152 15

Table 6
Studies that examine impact strength of UHPFRCs.

References Impact strength


[55] Rectangular slabs with dimensions 300  400 mm and (Specimen number 1, 692 Muzzle velocity [m/s], 80 Crater diameter-front side [mm], 19
thickness of 50 mm, the weight of the projectile was 124 Penetration depth [mm], 63 Mass loss [g], 63 Spalling [g]), (Specimen number 2, 706
grains (8.04 g) and the average muzzle velocity was 710 m/ Muzzle velocity [m/s], 68 Crater diameter-front side [mm], 21 Penetration depth [mm], 44
s. In total 18 slabs were tested for impact loading, panel was Mass loss [g], 44 Spalling [g])
punched but the projectile bounced back,
[145] Rectangular slabs with dimensions of 300  400 mm and crater diameter, (deformable projectile, when 1% fiber is used is 85 mm,
thickness of 50 mm and 45 mm, The weight of the both steel (300  400 mm large and when 2% fiber is used is 74 mm, when 3% fiber is used is
jacketed projectiles was 8.04 g. The length of the deformable 50 mm thick slab) 73 mm), the panel was punched but the projectile bounced
and non-deformable projectile was 23.2 mm and 26.6 mm, back
respectively. Yield strength of the mild-steel core was penetration depth (deformable projectile, when 1% fiber is used is 20 mm,
determined by compressive tests as 550 MPa. Distance to (300  400 mm large and when 2% fiber is used is 20 mm, when1% fiber is used is
the slab was 20 m. 50 mm thick slab) 19 mm), the panel was punched but the projectile bounced
back
Debris fragment mass non-deformable projectile, when 2% fiber is used is 60 g,
(300  400 mm large and other fiber ratios were not measured
50 mm thick slab)
Debris fragment mass Non-deformable projectile, when 0.5% fiber is used is 295 g,
(300  400 mm large and when 1% fiber is used is 99 g, when 1.5% fiber is used is
45 mm thick slab) 109 g, when 2% fiber is used is 98 g, when 2.5% fiber is used
is 126 g. The projectile passed through the slab entirely
Impact velocity-Residual Non-deformable projectile, when 0.5% fiber is used is
velocity (300  400 mm 705 m/s-226 m/s, when 1% fiber is used is 703 m/s-228 m/s,
large and 45 mm thick when 1.5% fiber is used is 712 m/s-161 m/s, when 2% fiber is
slab) used is 720 m/s-71 m/s, when 2.5% fiber is used is 711 m/s-
47 m/s
[155] The ogival nose projectile was characterized with diameter Ogival nose projectile (UHPFRC cylinder thickness 51 mm, 817 m/s striking
of 10.8 mm, caliber-head-radius of 3.0 and amass of 30 g; velocity, 1.5 mm depth of penetration), (UHPFRC cylinder
while the geometry of the 44.5 g conical nose projectile was thickness 50.5 mm, 823 m/s striking velocity, 10 mm depth
featured by 6 mm diameter, 13 mm nose height and 91 mm of penetration), (UHPFRC cylinder thickness 52 mm, 816 m/
total height. The UHPFRC targets were casted as 104 mm s striking velocity, 0 mm depth of penetration)
diameter cylinder with depth of about 50 mm and 100 mm, Conical nose projectile (UHPFRC cylinder thickness 50.5 mm, 1382 m/s striking
respectively. velocity, 28 mm depth of penetration), (UHPFRC cylinder
thickness 98.4 mm, 1360 m/s striking velocity, 24 mm depth
of penetration), (UHPFRC cylinder thickness 95.7 mm,
1380 m/s striking velocity, 17 mm depth of penetration),
[159] I section, 200 mm depth, 150 mm width, 1700 mm length, 3 (Maximum load at Heights 0.8 m; 238.77 kN, ultimate load 128.17 kN, ultimate midspan
deformed bar with diameter 13 mm, effective depth deflection 16.8 mm)
170 mm, 400 kg hammer, radius 90 mm, beam was (Maximum load at Heights 1.0 m; 241.3 kN, ultimate load 128.36 kN, ultimate midspan
supported over a span of 1200 mm. deflection 16.9 mm)
(Maximum load at Heights 1.2 m; 243.4 kN, ultimate load 128.52 kN, ultimate midspan
deflection 17.0 mm)
(Maximum load at Heights 1.4 m; 254.23 kN, ultimate load 128.67 kN, ultimate midspan
deflection 17.1 mm)
(Maximum load at Heights 1.6 m; 246.83 kN, ultimate load 128.79 kN, ultimate midspan
deflection 17.1 mm)
[142] All tests were conducted under the condition that air (impact times/absorption energy/average strain rate/dynamic compressive strength)
pressure in the high-pressure steel bottle was about (when 2% fiber is used; 1/0.770/24.8/154.1, 2/0.682/20.8/156.4, 3/0.682/21.9/154.7, 4/1.1/
3.4 MPa. Under this condition, the projectile speed was 39.1/141.7, 5/1.41/85.6/111.1, 6/0.753/129.5/33.3)
roughly 8.9 m/s. Then, the specimen was impacted (when 3% fiber is used; 1/0.786/30.5/146.3, 2/0.788/27.8/148.9, 3/1.069/42.5/138.3, 4/
repeatedly by the projectile with speed of 8.9 m/s until the 1.365/95.5/88.2, 5/1.802/124.6/40.6, 6/0.594/125.8/32.1, 7/0.155/142.1/13.2)
specimen was broken into pieces. (when 4% fiber is used; 1/0.633/25.1/157.5, 2/0.663/29.6/154.2, 3/0.672/28.0/147.4, 4/
0.839/30.5/145.8, 5/1.382/45.8/126.2, 6/1.522/51.8/116.9, 7/1.640/79.7/98.2, 8/1.075/
111.4/46.1, 9/0.349/138.4/20.5)
[66] 1000  1000  50 mm, the two identical vertical walls had a (Average results for Solid Round projectiles 12.7 mm; 63.3 cm3 Crater volume, 0.17 kg
thickness of 25 cm and were reinforced on both faces with Mass loss)
dense steel mesh. The vertical walls were fixed on a 50-cm (Average results for Anti-Tank Explosive Shells 40 mm; 490 cm3 Crater volume, 1.29 kg
reinforced slab, with top and bottom reinforcement mesh. Mass loss)
Two different types of projectiles were used, namely ‘‘Solid
Round’’ projectiles 12.7 mm and ‘‘Anti-tank Explosive
Shells’’ 40 mm. (Initial velocity: 120 m/s, Maximum
(impact) velocity: 300 m/s, Total projectile mass: 2250 g,
Mass of hollow charge: 900 g, shooting distance: 90 m),
(Maximum velocity: 820 m/s, Bullet mass: 116 g., Bullet
diameter: 12.7 mm., Shooting distance: 300 m)

the effect of size for UHPFRC at both type of load i.e. quasi-static as samples of incongruent sizes possessing side lengths of 50 mm,
well as impact loads. For most part, laboratory examinations are 70.7 mm, 100 mm, as well as 150 mm were considered. The cubic
carried out by means of reduced scale, and therefore, simplifica- samples enjoying generously proportioned size have demonstrated
tions must be prepared for real structures. This is the core reason inferior compressive strengths in comparison with those of diminu-
why a number of researchers have paid attention to investigate tive size. Also, the samples lacking in fibres have got the nod on the
the size effect of elements in the context of UHPFRC [70,76–80]. A same line of action with regard to the effect of size on the compres-
research lab analysis has been put to trial for the size effect of ele- sive strength in comparison with high-strength and standard con-
ments on the compressive strength of UHPFRC employing hetero- crete. While, the quantity of fibre is escalating, the size effect is
geneous fibre dosages by An et al. [75]. In that event, four cubic also turning out to be more momentous. By dint of three and
16 F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152

improved, the uni-axial tensile and flexure strengths of UHPFRC


decreased idiosyncratically, and that a better flexural resistance
impact was controlled than for the uni-axial tension testing [79].
The mutually conflicting results can be due to the reality that
although the flexural strength after cracking is mainly affected
by the fiber bridging ability instead of the matrix strength [82],
nevertheless, when the size impact was tested, they did not indi-
cate the fiber orientation. Not only have that, the placement meth-
ods used in such research have not been described. Consequently,
Yoo et al. [83] accounted that the divergences in the outcomes of
size consequence attained from preceding researchers [44,70,73]
were the result of dissimilar fibre orientation attributes. Their
study evidenced an effect of clear size in beams prepared with
UHPFRC, despite of application of matching placement techniques
owing to various flow velocity gradients for varying sizes of spec-
Fig. 11. Impact effect of UHPFRC slab – [40]. imens that escorts to various fiber distribution attributes. Their
[83] verification indicated that when identical fiber allocation
characteristics such as the fiber orientation, and number per unit
four-points bending analyses, Mahmud et al. [44] also Wille and area were acquired, the resulting size impact on UHPFRC beams
Parra-Montesinos [70] have conducted experiments for effect of was immaterial [83]. Accordingly, different fibre-distribution char-
size of specimen on the flexural strength performance of UHPFRC. acteristics were mainly the effects of size in UHPFRC beams.
The upshots have driven them to ultimate conclusion that the effect Moreover, Yoo et al. [83], of late, carried out several flexural
of size of samples on the attribute of flexural strength in UHPFRC is testing of UHPFRC fiber distribution attributes (i.e., fiber orienta-
inconsequential which trails the yield measure as it is highly duc- tion and fiber dispersion) with distinct dimensions and performed
tile. Also, Spasojavic et al. [76], have proven that the impact on size image analysis, along with amount of fibers per unit area at the
in small flexural UHPFRC employees was not a large thing. Lepech local cracks. Parallel to the findings achieved by Nguyen et al.
and Li [81], have likewise uncovered that engineered cementitious [73], UHPFRC beams produced from the positioning of concrete
composites has put on show no high-flying modification with on one end and allowing it to flow were noticeably reduced in flex-
respect to flexural strength in accordance with the size of sample ural performance, as the sample size increased despite the fibre
owing to it’s ductility. What’s more to add, Nguyen et al. [73] have aspect ratio 65 to 100 or fiber forms [79]. Nevertheless, the diver-
portrayed about the flexural output of UHPFRC with a higher tensile gent characteristics of fiber distribution were the primary founda-
ductility is less sensitive to the size impact than UHPFRC with a tion for the size impact in UHPFRC beams. The larger beams have a
lower ductility, namely the flexural strength, normalised deflection low fiber orientation with less fibre. Yoo et al. [71] had made sure
plus toughness. Conversely, their results were near linear elastic that when analogous for all test beams with unlike dimensions,
fracture mechanics (LEFM) as opposed to the yield criterion characteristics of fiber distribution were obtained in contrast to
[44,70], that was contradictory to the end results of other preceding the dissimilar properties, which are considerably less sensitive to
investigations of Mahmud et al. [44]. the size impact on flexural strength, as given away in Figs. 12
Analogously, Kazemi and Lubell [78] have expressed that smal- and 13. In view of that, finally, it was summarized that the size
ler UHPFRC specimens were found with higher compressive and consequence in UHPFRC beams is predominantly on account of
flexural strengths, along with direct shear strengths, in comparison the different fibre distribution characteristics; as a consequence,
with bigger specimens. Conspicuously, a boosted size consequence by guaranteeing analogous fibre distribution attributes, a trivial
on the shear strength had attained with more quantity of fibres. dimension effect on the flexural resistance could be achieved for
Also, Frettlohr et al. [80] have headed that when the sample size

Fig. 12. Crack patterns on surface; (a) Beams of small size, (b) Beams of medium-sized and (c) Beams of large-sized. [71].
F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152 17

Fig. 13. Cracking effect; (a) number of cracks and (b) average crack spacing. [71].

Fig. 14. Effect of specimen size on compressive performance of concrete [75].

Fig. 15. Effect of quantity of fibres per unit area on the specimen size [72].

UHPFRC incorporating steel fibres about 2%. Size effect on com-


pressive strength properties is demonstrated in Fig. 14. Fig. 16. Torex twisted triangular and square steel fibre [86].

The impact of rate of load on the size impact and the fracture
behavior of the concrete was researched by Bazant and Gettu
[81]. The dimension effect on the concrete fracture was signifi- represented that the strength of the tensile after cracking
cantly strengthened by the time to failure escalation and by the enhanced as the sample dimension boosted, whereas the strain
declined effective process area length the fragilities augmented competence and hardening declined. The sample size did not influ-
by trimming down the loading rate. Krauthammer et al. [84] of ence the scores of cracks. In accordance with them [43], the ground
axial (compressive) impact tests using cylindrical samples further for the boosted strength of tensile post-cracking is the inertial
described the insinuation of the rate of loading for the dimension impact of the sample itself. The inferior capacity of strain as well
effect of high-strength concrete. Tran et al. [43] examined the as toughness were met with bigger samples owing to the enhanced
impact of sample size on UHPFRC dynamic tensile. Their findings sample dimension resulted in a more possibility of attaining more
18 F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152

closed by Yoo and Banthia [72]. The impact of sample size and
impact of the UHPFRC fiber content proportion are displayed in
Fig. 15.

8. Effects of fıbre geometry, length, and volume fractions

‘‘Torex fibre” is a novel kind of twisted steel fibre that has been
developed by Naaman [85] in 1990 which is portrayed in Fig. 16. It
is made out of high-strength steel cable and has a polygonal
transversal geometry, which allows for a twist on the axis. The
original thought after defining its transversal geometry was based
on the Fibre Intrinsic Efficiency Ratio (FIER). FIER is strongly allied
to the strength of composites in post-cracking stage. When the
cross-sectional area is constant, the triangular and square-shaped
fibres are 28 percent and 12 percent more efficient, respectively
at augmenting the FIER value as compared to fibre having
circular-shaped [86]. When FIER value was augmented, the bond-
ing of fibres with matrix like adhesion as well as the friction was
also improved. The twisting of the fibres has brought about
Fig. 17. Bending behaviour of torex and hooked based UHPFRC [86].
improvement in the mechanical bond. The archetypical fibre stress,
slip curves of straight, hooked-end, as well as twisted steel fibres
incorporated in the UHPFRC matrix are evaluated by Wille and
Naaman [87].
Wille and Naaman [87] articulated that twisted addition of
hooked at end steel fibre, which was discovered to be three times
greater than short straight-line fiber of steel, achieved peak fibre
stress. On the basis of enhanced fibre pull-out capability, the
strength for tensile and strain of post-cracking competence of
UHPFRC were correspondingly augmented to a large extent
through employing the deformed (twisted plus hooked-end) steel
fibres, in comparison with the short straight-line fibers of steel
[88]. The tensile strength of 2 percent of twisted steel fiber was
14.9 MPa, and its strain capability of 0.61%; the values of tensile
steel were approximately 32 and 205% greater, respectively com-
pared to those with 2 percent of brief metal fibres. Bending beha-
viour of torex and hooked end steel fibres reinforced UHPFRC are
depicted in Fig. 17 [86]. Fig. 17 shows comparison between
UHPFRC specimen with torex fibres and commercially available
hooked steel fibres. Fig. 17 clearly indicate that torex fibres exhibit
Fig. 18. Load-displacement curves with different fiber content [131].
good performance compare to hooked steel fibres. More to add,
Yoo and Yoon [82] recorded that UHPFRC beams with twisted steel
fibers have a flexural approximately 1.7 times higher than beams
with short steel fibers. The squeezing deportment, like the com-
pressive strength, elastic modulus, as well as strain capacity were
also enhanced through the addition of twisted steel fibres as com-
pared to short straight line steel fibres. However, the enhancement
was comparatively of no consequence in comparison with moni-
tored value for the tensile performance. In recent times, Yoo
et al. [89,90] proposed one additional method for increasing
UHPFRC’s flexural efficiency under uniaxial as well as biaxial stress
and its capacity of fracture by using lengthy straight steel fibres. By
increasing the fiber length, UHPFRC flexural resistance, deflection
ability and robustness were significantly improved. UHPFRC panels
showed an increase in flexural strength by only approximately 26
percent and 13 percent and an increase in deflection capabilities by
153 percent and 67 percent compared to those with medium and
short steel fibres. Not only that but also application of longer sized
steel fibres the fracture energies were almost 121 percent and 35
percent greater in comparison with medium and short steel fibres
Fig. 19. Effect of fibre geometry on pullout behaviour of UHPFRC matrix [144]. [90]. Employing longer steel fibres boosts the bond area among the
fibre as well as the matrix that escorts to higher fibre pullout load
carting competence and capacity of slip [91]. What is more to add,
substantial imperfections, piloting to poorer crack linking capabil- the number of crack surface fibers that is a major factor affecting
ity. Also UHPFRC beams, dimensions effect with a wide range of the tensile behavior post-cracking has been modified in the same
(straight and twisted) steel fibers under impact loads have dis- diameter trivially with fiber length. For illustration, the fiber count
F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152 19

Fig. 20. Fibre distribution (a) Conventional concrete, (b) High strength concrete [98].

Fig. 21. Binary images of fibre orientation at center of UHPFRC beam; (a) Lf = 13 mm, (b) Lf = 16.3 mm, (c) Lf = 19.5 mm and (d) Lf = 30 mm. [90].

per unit area for short fibres was 34,00/cm2, the medium fiber et al. [94] have found out that UHPFRC fibers have a maximum
33,12/cm2 and the long one 35,79/cm2 [90]. compressive strength of up to a fiber volume fraction of 3 per cent
The root cause of the incongruous change in the number of with 2 vol% 13 mm long steel fibers.
fibers per unit area, although the actual number of fibers incorpo- While on the other hand, Yunsheng et al. [92] described that the
rated into the mixture decreases with the same fiber longitudi- compressive resistance improved constantly by up to 4% increase
nally. The option is that fibers can increase with the fiber length in fiber quantity-a compressive force 30–50 MPa greater that with-
on crack surfaces depending on their volume content. Because out fiber showed the 4 vol% steel fiber sample. In addition, by com-
the fibers are included in the mixture. In brief, The shorter the bining up to 2 vol% of steel fiber in the matrix, the fiber extraction
length of the fibre, the faster the mixing, hence the greater the performance was improved [94]. In this order both the strength of
amount of fibers on the crack area. The post-cracking flexural the tensile fibers and the stress capacity of UHPFRC were amplified
strength and resistance parameters were more or less linearly rein- by increasing their fiber volumes from 8 to 14 MPa, and from 0.17
forced with an increase in the fiber volume fraction in the three lin- to 0.24 percent, as well as for straight steel fibers from 8 to 15 MPa,
ear (or bi-linear) tensioning curve. Otherwise, The fibre volume and in this order, from 0.33 to 0.61 percent. Nevertheless, the ten-
fraction was trivially affected by the first cracking flexural strength sile resistance of the UHPFRC increased with an accelerating fiber
and the corresponding deflection [92–94]. Approximately 64 MPa, quantity from 9 to 14 MPa for hooked end steel fibers, whereas
which is nearly seven times greater than the flexural strength the stress capacity was continuous at about 0.46% [94]. Load-
without fibers after cracking with short straight line steel fibers displacement curves concerning fibre volume ratio and impact of
at a volume fraction of 5 percent [93]. A greater number of steel fibre geometry on pull-out strength of embedded in UHPFRC
fibers and elastic modules up to a fiber volume fraction of 3 per- matrix are shown in Figs. 18 and 19.
cent negligibly improve the compressive resistance [94]. Since Fig. 18 demonstrate the load displacement curve for diverse fibre
the compressive strength has a major impact on fiber dispersion volume ratios. It is clear that, as the fiber content increases, original
homogeneity, the optimal fiber volume fraction producing a max- rigidity does not alter substantially, while the load in the softening
imum compression force is different for various scientists. Prabha portion is gradually increasing to fragile behaviour. While the struc-
20 F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152

Fig. 22. Fibre orientation and dispersion parameters; (a) fibre dispersion coefficient, (b) number of fibre in the unit area, (c) packing density and (d) fibre orientation
coefficient. [90].

tural ductility rises with the FRC content of fiber, there is no apparent on a short edge of the mould and allowing for its flow [95] showed
trend in displacement at peak load with the UHPFRC fiber volume poor flexure in comparison to the parallel beams positioned as per-
ratios. Fig. 19 illustrates fibre stress along with slip curves for hooked pendiculars to the flow direction. The perpendicular arrangement
end steel fibre, twisted steel fibre as well as straight steel fibre in of the fibers was ascribed to the beam length as shown in
UHPFRC. From the figure, it could be seen that the pull-out stress Fig. 20. Fig. 21 illustrates fibre orientation and dispersion in the
of twisted steel fibre is three times higher than straight steel fibre middle of the UHPFRC beams on the cutting plane. Likewise, the
[93]. Twisted as well as hooked end steel fibres are enhanced the pull beams paralleled with the flow direction showed a much lower
out resistance as compared to straight steel fibres [93]. load capacity in the situation where the concrete is placed in the
center (radial flow) [96] compared to the beams in other areas.
They also proved an unusual behaviour in flexural for UHPFRC,
9. Effects of fıber orıentatıon which is softening of deflections. They showed Mechanical and
structural performance under tension as well as flexure are influ-
The UHPFRC is used to manufacture structures with various enced by the fiber orientation. Consequently, a lot of investigations
positioning methods [95–97]. Boulekbache et al. studied [98] that [15,70,96–98] have been carried out to quantifiably estimate how
fiber is rotated with various flow rates of fiber-reinforced concrete the fibre orientation features affect the mechanical attributes of
that can be flowing. The fibres, which exert strength and moments UHPFRC and pencil in a few valuable upshots. Fibre orientation
on the fibres, are vertically lined up in radial flow and parallel to and dispersion parameters are shown in Fig. 22.
the flow direction in shear flow. The manufactured UHPFRC steel The effects of casting methods, namely layer casting plus
bar beams by Kang et al. [99] were putting concrete using two dif- middle-casting, casting speed and the flexural behaviour of uniax-
ferent placement methods (i) allowing for one end of form to flow ial UHPFRC beams, were investigated by Wille and Parra-
to another end or (ii) the center agreeing to flow to both ends. Their Montesinos [70]. It was noted that a snake-like flow pattern could
experiment results revealed a peak load that was about 15% be prevented when casting velocity increased and a thinner layer
greater than when cement was put at the center of the concrete could be made with a preferable fiber layer in the beam axis to
beams. Due to UHPFRC’s flow capacity, more fibers are oriented increase flexural efficiency. In addition, the beams cast in the cen-
towards the longitudinal direction of the beam. The rectangular ter show a transitional flexural resistance between the beams cast
slabs, unlike positioning processes, were also made by Ferrara in layers which have a high as well as low casting speed. Likewise,
et al [95] and Kwon et al. [96] and examined how fiber orientation Yoo et al. [97] also supervised that the center beams contributed to
affects UHPFRC flexural performance. The case of placing concrete greater flexural strength than the edge beams. The energy of the
F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152 21

Fig. 23. Compressive strength (a) and Flexural strength (b) under quasi-static loading [106].

fractures was nevertheless somewhat influenced by the casting occurred because its post-cracking characteristics are more influ-
method since the benefit of the greater strength resistance com- enced than matrix strength by the fiber bridge ability. The fiber
pensated for a rapid decrease in stress after peak. They also sub- distribution features are therefore also intended to affect the ten-
mitted image analysis and verified that more steel fiber is sile or flexural conduct at elevated load rates of UHPFRC. However,
positioned in the center for beams cast in the center (at peak time only a tiny amount of investigators [97] studied the impact of fiber
region) than the beams cast on the edge to evaluate the experi- orientation on UHPFRC impact resistance. Yoo et al. [97] have
ment results realistically. The flexural efficiency of UHPFRC panels lately experimentally studied for the first time how the rate based
under a biaxial stress condition was examined by Barnett et al. [15] UHPFRC flexural features are affected by fibre orientation. Investi-
and Yoo et al. [96]. However, various methodologies were gations were carried out successfully; [97] the fiber orientations in
employed-the method of the newly fetched biaxial flexural testing the UHPFRC beams were intentionally good and poor, with two dif-
(BFT) suggested by Zi et al.[100]. However, similar testing findings ferent dimensions. A correct fiber orientation is the case that the
had been achieved. The UHPFRC panel cast in the center showed fibers are mostly well aligned in the direction of tensile load, while
significantly greater flexing forces than the panels cast using vari- the poor fiber orientation suggests that most fibers are arbitrary or
ous placing techniques, i.e. casting around, casting arbitrarily and tend to tensile loads. Based on the image analysis, quantitatively
casting around the perimeter of the panel on a number of points. explored the characteristics of fiber distribution, together with
Yoo et al. [97] used binary images on the crack surfaces instead fiber orientation, dispersion and unit number in the unit region.
to work on picture analyses, which were changed by a high- When correct fibre orientation was authorized-fibers are aligned
resolution camera with RGB pictures. The information from the in a more tensile stress direction-the fibre orientation significantly
analyses showed that, the rise in the flow speed in the other panels affected the flexural performance in UHPFRC at impact load and
with varied positioning methods, a larger amount of stainless-steel increased impact strength, i.e. improved flexural strength and
fibers in the center panels were lined up at a correct angle to the energy absorption ability. Despite the potential energy and sample
flow direction. The aforementioned improved fiber arrangement size, UHPFRC beams with correct fiber orientation showed greater
led to a greater flexural strength and strength in the center of resistance to flexural impacts. The findings of Xu et al. [104] on sin-
the panels than in the others. The fiber movement was evaluated gle fiber pull out tests are supported by the finding that the
arithmetically by Kang and Kim [101], based on the Jeffery equa- increased fiber pull-out strength is achieved if it tends slightly in
tion [102], assuming that there were no fiber interactions. They the direction of load. Ideal alignment in the casting direction is
described that the fibers were switched more parallel to the direc- almost impossible because interaction between fiber to fibre, with
tion of the stream (for shear flow) and more vertical to the direc- the fiber concentration is generated and stress applied by a fluid-
tion of the stream (for the radial flow). The actual findings from speed trimmer in the fiber is reduced when the fiber structure
shear and radial fluxes that have been tested by Yoo et al. [97] have approaches the flow direction. Consequently, the fibre-friendly
been unsurpassed by these statistical findings. According to Yoo beam may be more fibre-friendly with a slightly tendency toward
et al. [97] and Lee et al. [103], the fiber orientation distribution the stress than its counterpart beam, with an increase in impact
Probability Function (PDF) for both UHPFRC and Engineered resistance under flexure. It has been noted as a whole that appro-
Cementitious Composites (ECC) showed completely poles apart priate fiber aligning with the tensile load requires excellent shock
from the behavior of suppositions for two and three dimensional. or blast resistance to be achieved in UHPFRC components.
Nevertheless, the use of a 2-D arbitrary fiber orientation is more
worthwhile in simulating the flexural behavior of uniaxial UHPFRC
beams than the 3-D random fiber orientation, based on analysis 10. Effect of fıbre type
based on micro-mechanics [87] it can be significant.
The findings of quite a number of previous studies [70,73,90,99] The effect of a type of fiber on the tensile behavior of UHPFRC at
show that the fiber distribution characteristics of fibre orientation quasi static and impact loads has been demonstrated by research
and number per unit region are considerably influenced by the ten- tests by Tran et al. [43]. The highest impact resistance in the
sile or flexural efficiency of UHPFRC at virtually static stresses. This postcracking strength, stain capabilities and toughness were
22 F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152

demonstrated in UHPFRC dog-bone samples with lengthy and fibre tensile strength, number of twists and fibre orientation.
straight steel strain range from 5 to 24/s. Although the use of Therefore, advance research work should be put in execution to
twisted steel fibers demonstrated maximum tensile resistance at pencil in unambiguous conclusions concerning the effectiveness
quasi-static loads, the strain capabilities and toughness at high of the use of UHPFRC twisted steel fibers for loading provisions.
tensile rates were found to be substantially below both longer The inferior quality impact resistance of the UHPFRC with the mas-
and short steel fibers due to fiber splintering. Also, analogous find- sive quantity of twisted steel fibers can indeed be allocated to the
ings of comments have been recorded by some additional investi- fact that twisted steel fiber has increased its bond strength because
gators [44,58]. the fiber splitting rate is more likely to result in tension or fiber
Wille et al. [105] pointed out that twisted steel fibers, due to pulling at elevated stress rates. Therefore, the significantly
their limited strain state resistivity tolerance, led to lower tensile decreased output energy of twisted steel fibers at greater strains
strength at elevated load rates than straight steel fibres. Not only has given the implementation of straight steel fibers a lower
that, Yoo et al. [106] have pointed out that the UHPFRC beams have impact strength.
been controlled with a greater flexure resistance and energy Yu et al. [109] have shown that UHPFRC’s impact strength in
absorption ability after cracking than those with short twisting place of the short fibers is controlled by the lengthy steel fibres.
and straight-driven steel fibers as shown in the Fig. 23. The mixing The impact resistance of UHPFRC was degraded, according to
proportions, straight stainless-steel fiber type and mechanical sta- [109], with the constant content of the fiber volume increasing
tic characteristics of the sample S65 are much closer to commercial the substitution rate for long fibers to short fibres. Nevertheless,
UHPFRC [90]. It can therefore be concluded that using long straight the improved impact strength and static flexural efficiency of
or twisted steel fibre improves commercial UHPFRC impact UHPFRC in lengthy fibres have been accomplished based on the
strength. In addition, the improved impact strength with lengthy previous trial results by Tran and Kim [104] and Yoo et al. [110].
straight steel fibers compared to twisted fibers was supervised in In consequence, fusion of long or medium-length steel fibres, both
the UHPFRC. This was allocated to the lower stress sensitivity to static and impact loads were promoted to improve the tensile or
DIF, shown by the twisted steel fibers, for postcracking strength, flexural performance of UHPFRC.
which leads to a reduced increase in strength at an accelerating Millard et al. [111] studied UHPC containing an amalgam of
stress rate. The clarifications are in line with the results collected both short as well as long steel fibres and found it displaying infe-
by Tai et al. [107] that show that twisted steel fibers have a rior strain-rate sensitivity, signalling lesser dynamic increase factor
reduced rate sensitivity to impact fiber pull out than straight steel at equal strain-rate, in comparison with that embracing merely the
fibers. Nevertheless, it is controversial that although single, twisted single short fibres. The core reason behind this was the reality that
fiber in UHPFRC displays a high bond strength and dissipating the mixture of fibres is greater effectually governs the formation of
energy capabilities at impact pull-out loads, UHPFRC-composites lateral crack development than its corresponding part. The pitiable
containing numerous twisted fibers arbitrarily oriented to the pro- fibre distribution in UHPFRC has been attained often when mal-
vision of an expanded weak effect resistance, as opposed to formed steel fibres were incorporated. Predominantly, when the
straight fibres [107]. In addition, due to the wonderful fiber bridg- hooked steel fibre was integrated with UHPFRC mix, deficient sep-
ing ability, the deflection-hardening behavior of UHPFRC has been aration of fibres was monitored on account of its bundles, escalat-
recorded at impact loads. [97]. ing crises for constructability and cracking of matrix at close to the
In addition, Pyo et al. [108] conducted an impact test for fibre end hook [112]. For that reason, a depreciated pull-out capac-
UHPFRC using a customized Strain Energy Frame Impact Machine. ity of the hooked fibre took place resulting in the weak flexural
The investigations of the UHPFRC with twisted steel fibres, though, performance of UHPFRC in comparison with one those containing
witnessed for higher impact resistance in the context of post- straight steel fibres. Long straight steel fibers demonstrated maxi-
cracking strength and ability to absorb energy in comparison with mum rate sensitivity after cracking when the short straight steel
straight steel fibres having a variety of aspect ratios. The preceding fibers showed the greatest strain and peak resistance sensitivity
testing upshots were divergent to this work creating an inconsis- [43].
tency among them. This might be owing to the variations in testing On the other hand, one more study by [113], straight steel fibres
equipment and fibre characteristics, like dispersion, aspect ratio, aspect ratio had not displayed any obvious impact on the dynamic
increase factor of the post-cracking resistance to flexural. These
scientists have significantly tracked the remaining flexure strength
and toughness significantly increased by the use of lengthy
stainless-steel fibers following impact damage in comparison with
small stainless steel fibers. Long, straight fibers of steel are there-
fore effective reinforcements to improve the strength of UHPFRC
following impact harm after cracking and residual loads. Yoo
et al. [89,106] has already demonstrated the effectiveness of using
lengthy straight steel fibers in UHPFRC in quasi-static loading
states. The collective application of micro steel as well as Basalt
fibres piloted to enhanced resistance to the impact and Trinitro-
toluene (TNT) blast [114,115], resultant of the synergistic impact
among the fibres. On the basis of the testing findings [114], little
permeation depth of UHPFRC was attained by repetitive impact
on utilizing both – steel as well as basalt fibres. On relying preced-
ing comparative pull-out behaviours of steel and polymeric fibres
entrenched in control or high-strength cement matrix, In addition,
polymeric fibres are advantageous to contribute pull-out beha-
viour as slip-hardening in comparison with steel fibres. Although,
the inferior tensile resistance of most of polymeric fibres, narrowed
Fig. 24. Pressure-deflection relationships for UHPFRC slabs obtained under static their use in UHPFRC owing to the fibre fracture prior to absolute
uniformly distributed loading. [122]. pull-out. Recently, Polyethylene (PE) fibres displaying the extre-
F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152 23

mely high tensile performance of roughly around 2700 MPa are As the steel fibres cover a momentous portion of the entire cost
developed, and they were productively employed to the cement of fabrication for UHPFRC, it is more methodically carrying great
matrix with the sky-scraping compressive strength of more than weight to put forward an optimum quantity of steel fibres, con-
150 MPa [116]. Given the fact that the polymeric fibres by and tributing adequate impact resistance, in comparison with explain
large offer enhanced pull-out behaviour than steel fibres, the use that elevated impact resistance is attained at more fibre volume
of PE fibres in UHPFRC mix may provide an enhancement in the contents, which is rather expected, in the context of being econom-
context of impact resistance. Of late, Ranade et al. [117] have com- ical. Tai [124] accounted that, while flat end projectile was utilized,
pared the resistance against impact of high strength concrete pos- harsher smash up was monitored for the UHPFRC panels lacking
sessing PE fibres and UHPFRC along with steel fibres- hooked type fibre scabbing larger area in comparison with standard strength
and reported that the PE fibres demonstrated flexural behaviour concrete panels, because of their brilliant fragility. The failure
more ductile in nature along with dispersed manifold micro size mode of UHPFRC panels with no fibre was switched from fragile
cracks than the UHPFRC on subjecting to impact loads drop- to pseudo plastic through incorporating steel fibres: on an adding
weight type. up of 1 vol% steel fibres, around 50% slighter scabbing area was
occupied [124]. Fascinatingly, Millard et al. [111] noted that the
11. Effect of quantity of fıbres increase of UHPFRC flexural strength by elevated rate of load, i.e.
loading rate sensitivity, can be reduced by increasing the fiber con-
Rong and Sun [113] performed tests with various amounts of tent between 0% and 6%. In accordance with their description
short straight steel fibers on the UHPFRC dynamic tensile beha- [111], the effects of a greater loading rate are allocated to the
viour with the help of a SHPB test machine and concluded consid- decline, because fibers bridge the cracks in low resistance areas
erably based on their test results. They specifically indicated that are reduced to lateral crack growth. Pressure deflection relations
the tensile strength of UHPFRC increased, supported by the for UHPFRC slab loaded with a uniform static load and the relation
upshots of Yunsheng et al. [118], by an rise in stress rates and fiber between the quantity of the fiber and the critical scale size of the
volume content. In contrast to UHPC without any fiber, UHPFRC’s slab are indicated in Figs. 24 and 25.
low strain rates achieved 2 and 2,5 times greater tensile strengths
with stainless steel fibres of 3 and 4 vol%. Also, the direct tensile 12. Applications and prospective recommendations
behavior of UHPFRC was calculating by Pyo et al. [119,120], with
various amounts of fiber of steel. Parallel to Rong and Sun [113], The strength and lightness of the structural elements have the
they advocated that the tensile performance, strain capacity to power over the bulk of the applications whereas the rest of the uti-
absorb energy be increased by increased twisted steel fibers lizations consider the belligerent resistance, sustainability, and
amount at high strain rates [119], while the strain hardening durability of the material to safeguard chiefly the existing or new-
response was monitored through the use of the Strain Energy fangled structures. Quite a lot of applications like anchor pre-
Frame Impact Machine. Therefore, Pyo et al. [120] and Yoo et al. stressing are erected primarily and essentially on compressive
[121] studies were used to determine UHPFRC strain and strength for columns or extremely stressed features. Also, UHPFRC
deflection-hardening behavior in respect of impact loads. Beyond has been found fitting to apply in architectural constructions. The
the debates, it is important to think that UHPFRC’s dynamic tensile erection of copious prototype structures have been made for
performance is improved in straight and deformed steel fiber UHPFRC in so many countries such as New Zealand, Japan, Germany,
instances with a greater content of fibre. USA, Canada, South Korea, Australia, France, Malaysia, etc. In 1997,
Mao et al. [122] were also along the same lines as UHPFRC pan- UHPFRC was applied for the first time ever for a UHPFRC-infilled
els improve their blast resistance, with fiber volume content steel tube composite to build footbridges in the city named Sher-
increasing up to 6%. With a rise in fiber content from two to six vol- brooke, in southern Quebec, Canada [125]. This was the era of com-
umes percent at matching explosive charges and standby distance, mencement of UHPFRC to draw attention of universal researchers,
the maximum 1/4 deflection of UHPFRC panels was significantly concrete technologists, academicians, engineers, several depart-
decreased. On the other side, the optimized fibre-volume content ments and authorities of government, etc. The eye-catching illustra-
of UHPFRC to decrease local smashing was suggested as about tion to implement UHPFRC structurally was during the period of
2 vol% on the basis of an previous research conducted by Máca 1997–98 in France, whereby the replacement of beams for Civaux
et al. [61,123]. Their experimental findings however have and Cattenom cooling towers was performed displaying how it
improved the effect strength of UHPC panels by integrating short can be employed in aggressive conditions of the environment. One
stainless steel fibres, while further increases in fiber content more example can be cited from the city of Seoul, South Korea during
beyond 1 vol% and 2 vol% showed no noticeable insinuations on the year 2002 whereby the first footbridge of 120 m was built totally
the penetration depth and crater diameter of UHPFRC panels by
effect of projectile impact accordingly. Finally, the discussion leads
to the conclusion that the massive breakdown of UHPFRC elements
is improved by an increase in the fiber content up to 6% while the
increase in the resistance to local damage by the projectile impact
is limited to the fiber content and is discovered to be around 2% of
its optimum quantity. The tensile strength of UHPFRC increases,
without a doubt, when the amount of steel fibers is increased,
because more fibers are bridging the splits and restricting the fur-
ther spreading and amplification of the splits. But increasing the
fibers has also resulted to disadvantages such as reduced down-
flow ability, greater porosity, more fiber alignment, fiber ball phe-
nomenon and so on, and therefore, a linear rise in tensile strength
with the fiber amount has not been achieved.
Also, Yoo et al. [110] noted, because of numerous shortcomings,
a non-linear increase in the post-cracking flexural strength of
UHPFRC with a fibre refractory index. Fig. 25. Fibre volume and critical scaled distance of slab relation [122].
24 F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152

of substandard bridge substructures, and blast mitigation and safe-


keeping utilizations. The elevated-capacity UHPFRC projects for
infrastructures are highlighted by accomplished projects of bridge
viz., the Seonyu Footbridge of South Korea; the Sherbrooke Foot-
bridge of Canada; Shepherds Gully Creek Bridge of Australia; the
Bourg Les Valence Bridge of France; Gärtnerplatz footbridge in
2007 at Kassel, Germany; Seonyu footbridge in the year 2002, Seoul,
Korea); and Sakata-Mirai footbridge of 2002, in Japan as well as roof-
ing panels for silo in the year 2001, Joppa, Illinois USA; Cattenom
nuclear power plant in 1997–1998, France; Shawnessy station in
2005, Calgary, Canada; and roof of Millau Viaduct tollgate during
2004, France.
Recommendations are necessary to suggest for healthier under-
standing of the potential of UHPFRC. So far, the majority of the
investigations were carried out on RC beam elements in context
of UHPFRC but studies on other structural elements such as walls
and columns are quite essential. Moreover, the durability perfor-
mances of strengthened RC members such as resistance to attack
of chemicals, namely, acid, sulphate as well as chloride ion attacks
should also be examined and studied in detail to offer the requisite
fortification against infiltration by such deteriorating factors creat-
ing a concern for durability and sustainability. Not only have that,
there is quite a significant call for study the performances of
UHPFRC under diverse conditions of environments including those
of aggressive kind.

13. Conclusions and Discussion

The following findings can be taken on the basis of a review of


previous literature from UHPFRC and conversations:

 The best mechanical characteristics in most UHPFRCs were


obtained with water to cement ratio less than 0.20, steel fibre
3 percent of volume, steel fibre length of about 13 mm hook
end type in the UHPFRC mixtures. The highest mechanical char-
acteristics were obtained with the specimens exposed to ther-
mal curing at 90 °C. The compressive performance of the
samples that were exposed to thermal curing was higher than
those exposed to water curing.
 The most imperative factors that affect the behaviour of the
UHPFRC were identified to be the fibre type, fibre geometry,
fibre volume fractions, the distribution as well as orientation
of the fibres in the concrete, and the matrix of the UHPFRC.
Twisted steel fibres reinforced UHPFRCs exhibited much higher
mechanical properties than those reinforced by hooked end
Fig. 26. Tensile behaviour of UHP-FRC using twisted steel fibres with increasing
steel fibres followed by those reinforced by smooth straight
volume content and same fibre aspect ratio [120].
steel fibres. Increasing the fibre length does not effect after
13 mm.
with UHPFRC in the world as reported by Deem [126]. Subsequently,  UHPFRCs showed about 2.2 times higher tensile and flexural
the VSL road bridge of Shepherds Gully Creek in Australia was built strengths than their counterpart ultrahigh strength concrete.
and launched in 2005 for public use. However, the pressing UHPFRC The impact resistance performance of UHPFRC was better than
beams built with standard concrete, but the dimensions of element static load, and impact resistance improved with the boost in
were significantly declined along with desirable durability and fibre content up to 2%. UHPFRC was capable of dissipating
lightness [127–129]. UHPFRC can be employed in an extensive range higher absorbed energy by impact than traditional concrete
of utilizations for highway infrastructures thereby offering a longer with fibres. Steel fibres contents of >2% showed no positive
period of design, slender overlays, shelves and claddings on account effect in terms of penetration depth of UHPFRC due to impact.
of its enhanced durability as well as higher compressive and tensile  Although the mixtures with long fibres exhibited a slight
strengths. Hitherto, the examples of highway bridges built with enhanced in flexural strength of UHPFRC compared to those
UHPFRC are a comparatively limited in quantity available in more with short fibres, however, UHPFRCs reinforced by short steel
often than not in Canada [130–132] and USA. The key investigations fibre yielded the same compressive strength values. Short fibres
have been carried out in France, with significant development on the are preferred because long fibres pointedly inferior the worka-
topic of bridges made up with UHPFRC. In Australia and Italy certain bility of the mixes.
applications are also available [133,134]. What’s more, a lot of uses  Using lengthy straight UHPFRC steel fibers was generally more
of UHPFRC are regarded as significant e.g. In the precast concrete efficient than the twisted and rounded steel fibre in terms of
piles, thin-bound overlays of degraded decks, the seismic retrofit enhancing the bond strength and energy dissipation capabilities
F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152 25

and displaying the greatest impact resistance, with post- [8] S. Luhar, S. Gourav, A review paper on self healing concrete, J. Civ. Eng. Res. 5
(3) (2015) 53–58, https://doi.org/10.5923/j.jce.20150503.01.
cracking strength and energy absorption capabilities. UHPFRC
[9] V. Kodur, R. Mcgrath, Fire endurance of high strength concrete columns, Fire
has also shown very high stress sensitivity. The twisted steel Technol. 39 (1) (2003) 73–87.
fibers were, however, the most efficient way to improve energy [10] D.E. Dixon, J.R. Prestrera, G.R. Burg, E.A. Abdun-Nur, S.G. Barton, L.W. Bell, S.J.
dissipation at quasi-static and dynamic concentrations. A Blas Jr, R.L. Carrasquillo, P.M. Carrasquillo, A.C. Carter, Standard Practice for
Selecting Proportions for Normal, Heavyweight, and Mass Concrete (ACI
decrease in diameter increased the UHPFRC straight steel fiber 211.1-91), (1991).
rate sensitivity. [11] J. Xiao, Z. Li, J. Li, Shear transfer across a crack in high-strength concrete after
 Larger UHPFRC specimens contributed to inferior mechanical elevated temperatures, Constr. Build. Mater. 71 (2014) 472–483.
[12] C. Wang, C. Yang, F. Liu, C. Wan, X. Pu, Preparation of ultra-high performance
strengths to smaller samples. In addition, a higher size impact concrete with common technology and materials, Cem. Concr. Compos. 34 (4)
for bending strength was achieved than the strengths obtained (2012) 538–544.
by uni axial test. The magnitude of the flexural resistance was [13] K. Koh, G. Ryu, S. Kang, J. Park, S. Kim, Shrinkage properties of ultra-high
performance concrete (UHPC), Adv. Sci. Lett. 4 (3) (2011) 948–952.
mainly because of discrepancies in the characteristics of fibre [14] F. Lagier, B. Massicotte, J.-P. Charron, Experimental investigation of bond
allocation, i.e. bigger samples escorted to a weak fiber direction stress distribution and bond strength in unconfined UHPFRC lap splices under
and fewer fibres. direct tension, Cem. Concr. Compos. 74 (2016) 26–38.
[15] S.J. Barnett, J.-F. Lataste, T. Parry, S.G. Millard, M.N. Soutsos, Assessment of
 At greater stress rates, a substantial improvement in UHPFRC’s fibre orientation in ultra high performance fibre reinforced concrete and its
mechanical properties was achieved. The content of the fiber effect on flexural strength, Mater. Struct. 43 (7) (2010) 1009–1023.
quantity has no significant effect on rate of strain sensitivity. [16] C.M. Tam, V.W. Tam, K.M. Ng, Assessing drying shrinkage and water
permeability of reactive powder concrete produced in Hong Kong, Constr.
The use of twisted steel fibers reduced tensile strength, com-
Build. Mater. 26 (1) (2012) 79–89.
pared to straight steel fibers, by approximately 10 percent, at [17] D.K. Thai, S.E. Kim, Failure analysis of UHPFRC panels subjected to aircraft
a elevated strain rate of 10–1/s in UHPFRC. The strength of flex- engine model impact, Eng. Fail. Anal. 57 (2015) 88–104.
ure resistance is twice and dissipated energy of three or four [18] M.A. Saleem, A. Mirmiran, J. Xia, K. Mackie, Ultra-high-performance concrete
bridge deck reinforced with high-strength steel, ACI Struct. J. 108 (5) (2011)
times more than standard fiber-reinforced concrete (FRC). 601.
 While more studies reported static and dynamic mechanical [19] B.A. Graybeal, Flexural behavior of an ultrahigh-performance concrete I-
properties of conventional UHPFRCs containing cement as well girder, J. Bridge Eng. 13 (6) (2008) 602–610.
[20] C. Shi, Z. Wu, J. Xiao, D. Wang, Z. Huang, Z. Fang, A review on ultra high
as silica fume as the binder, relatively fewer studies reported performance concrete: Part I. Raw materials and mixture design, Constr.
mostly static mechanical performance of UHPFRCs comprising Build. Mater. 101 (2015) 741–751.
various supplementary cementitious materials. The application [21] A.N. Mohammed, M.A.M. Johari, A.M. Zeyad, B.A. Tayeh, M.O. Yusuf,
Improving the engineering and fluid transport properties of ultra-high
of supplementary cementitious materials for diverse scale sub- strength concrete utilizing ultrafine palm oil fuel ash, J. Adv. Concr.
stitution of cement as well as silica fume could considerably Technol. 12 (4) (2014) 127–137.
diminish the materials cost of UHPFRC and carbon footprint of [22] S. Luhar, S. Chaudhary, I. Luhar, Thermal resistance of fly ash based
rubberized geopolymer concrete, J. Build. Eng. 19 (2018) 420–428, https://
this material. In an estimate, 40 percent fly ash in the UHPFRC doi.org/10.1016/j.jobe.2018.05.025.
shows approximate 35 percent lower CO2 emissions than the [23] S. Luhar, S. Chaudhary, U. Dave, Salmabanu Luhar, Effect of different
control UHPFRC. It is also shown that UHPFRC incorporating parameters on the compressive strength of rubberized geopolymer
concreteEffect of different parameters on the compressive strength of
suitable quantity of supplementary cementitious materials
rubberized geopolymer concrete, in: Multidisciplinary Sustainable
could attain mechanical properties which are slightly lower Engineering: Current and Future Trends, 2016, pp. 77–86, https://doi.org/
than the conventional UHPFRC. 10.1201/b20013-13.
[24] M. Yalçın, C. Tasßdemir, G.I.H. _ Ekim, M. Yerlikaya, Design of steel Fiber-
Reinforced Concrete Based on Usability and Bearing Capacity Limits, 7.
National Congress on Concrete, Deevelopments Implementations Concr.
Declaration of Competing Interest Technol. (2007) 353–362.
[25] B.A. Graybeal, Compressive behavior of ultra-high-performance fiber-
reinforced concrete, ACI Mater. J. 104 (2) (2007) 146–152.
The authors declare that they have no known competing finan- [26] T. Makita, E. Brühwiler, Tensile fatigue behaviour of Ultra-High Performance
cial interests or personal relationships that could have appeared Fibre Reinforced Concrete combined with steel rebars (R-UHPFRC), Int. J.
Fatigue 59 (2014) 145–152.
to influence the work reported in this paper.
[27] E. Brühwiler, E. Denarié, Rehabilitation and strengthening of concrete
structures using ultra-high performance fibre reinforced concrete, Struct.
References Eng. Int. 23 (4) (2013) 450–457.
[28] D.-Y. Yoo, N. Banthia, Mechanical properties of ultra-high-performance fiber-
reinforced concrete: a review, Cem. Concr. Compos. 73 (2016) 267–280.
[1] P. Rashiddadash, A.A. Ramezanianpour, M. Mahdikhani, Experimental [29] W. Wang, J. Liu, F. Agostini, C.A. Davy, F. Skoczylas, D. Corvez, Durability of an
investigation on flexural toughness of hybrid fiber reinforced concrete ultra high performance fiber reinforced concrete (UHPFRC) under progressive
(HFRC) containing metakaolin and pumice, Constr. Build. Mater. 51 (2014) aging, Cem. Concr. Res. 55 (2014) 1–13.
313–320. [30] S.-T. Kang, J.-K. Kim, The relation between fiber orientation and tensile
[2] S. Thokchom, P. Ghosh, S. Ghosh, Performance of fly ash based geopolymer behavior in an Ultra High Performance Fiber Reinforced Cementitious
mortars in sulphate solution, J. Eng. Sci. Technol. Rev. 3 (2010) 36–40. Composites (UHPFRCC), Cem. Concr. Res. 41 (10) (2011) 1001–1014.
https://doi.org/10.25103/jestr.031.07. [31] K. Habel, Structural behaviour of elements combining ultra-high performance
[3] S. Luhar, T.-W. Cheng, D. Nicolaides, I. Luhar, D. Panias, K. Sakkas, Valorisation fibre reinforced concretes (UHPFRC) and reinforced concrete, (2004).
of glass wastes for the development of geopolymer composites – durability, [32] F. Toutlemonde, J. Resplendino, Designing and Building with UHPFRC: State of
thermal and microstructural properties: a review, Constr. Build. Mater. 222 the Art and Development, ISTE2011.
(2019) 673–687, https://doi.org/10.1016/j.conbuildmat.2019.06.169. [33] E. Camacho, J.Á. López, P.S. Ros, Definition of three levels of performance for
[4] S. Luhar, T.-W. Cheng, D. Nicolaides, I. Luhar, D. Panias, K. Sakkas, Valorisation UHPFRC-VHPFRC with available materials, Ultra-High Performance Concrete
of glass waste for development of Geopolymer composites – mechanical and Nanotechnology in Construction. Proceedings of Hipermat 2012. 3rd
properties and rheological characteristics: a review, Constr. Build. Mater. 220 International Symposium on UHPC and Nanotechnology for High
(2019) 547–564, https://doi.org/10.1016/j.conbuildmat.2019.06.041. Performance Construction Materials, kassel university press GmbH, 2012, p.
[5] S. Luhar, T.-W. Cheng, I. Luhar, Incorporation of natural waste from 249.
agricultural and aquacultural farming as supplementary materials with [34] S. Yang, S. Millard, M. Soutsos, S. Barnett, T.T. Le, Influence of aggregate and
green concrete: a review, Compos. B Eng. 175 (2019), https://doi.org/ curing regime on the mechanical properties of ultra-high performance fibre
10.1016/j.compositesb.2019.107076 107076. reinforced concrete (UHPFRC), Constr. Build. Mater. 23 (6) (2009) 2291–2298.
[6] S. Luhar, P. Chaudhary, I. Luhar, Influence of Steel Crystal Powder on [35] P. Rossi, Ultra-High Performance Fiber-Reinforced Concretes, Concr. Int. 23
Performance of Recycled Aggregate Concrete 431 (2018) 102003, (12) (2001) 46–52.
https://doi.org/10.1088/1757-899x/431/10/102003. [36] P. Rossi, Ultra high performance concretes, Concr. Int. 30 (02) (2008) 31–34.
[7] S. Luhar, S. Chaudhary, I. Luhar, Development of rubberized geopolymer [37] R. Suter, L. Moreillon, C. Clergue, R. Racordon, Using UHPFRC for Complex
concrete: strength and durability studies, Constr. Build. Mater. 204 (2019) Façade Elements, Designing Build. UHPFRC (2011) 405–420.
740–753, https://doi.org/10.1016/j.conbuildmat.2019.01.185.
26 F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152

[38] A.G. de travail BFFUP, Ultra High Performance Fiber-Reinforced Concretes: [66] D. Nicolaides, A. Kanellopoulos, P. Savva, M. Petrou, Experimental field
Interim Recommendations, Scientific and Technical Committee, Association investigation of impact and blast load resistance of Ultra High
Française de Genie Civil (2002). Performance Fibre Reinforced Cementitious Composites (UHPFRCCs),
[39] A. Simon, New AFGC Recommendations on UHPFRC: Chapter 1–Mechanical Constr. Build. Mater. 95 (2015) 566–574.
Characteristics and Behavior of UHPFRC, Designing and Building with [67] A. Lampropoulos, S.A. Paschalis, O. Tsioulou, S.E. Dritsos, Strengthening of
UHPFRC (2011) 723-742. reinforced concrete beams using ultra high performance fibre reinforced
[40] B.A. Tayeh, B. Abu Bakar, M. Megat Johari, A. Zeyad, Microstructural analysis concrete (UHPFRC), Eng. Struct. 106 (2016) 370–384.
of the adhesion mechanism between old concrete substrate and UHPFC, J. [68] Y. Kusumawardaningsih, E. Fehling, M. Ismail, UHPC compressive strength
Adhes. Sci. Technol. 28 (18) (2014) 1846–1864. test specimens: cylinder or cube?, Procedia Eng 125 (2015) 1076–1080.
[41] B.A. Tayeh, B.A. Bakar, M.M. Johari, M. Ratnam, The relationship between [69] Z.P. Bazant, Size effect in blunt fracture: concrete, rock, metal, J. Eng. Mech.
substrate roughness parameters and bond strength of ultra high-performance 110 (4) (1984) 518–535.
fiber concrete, J. Adhes. Sci. Technol. 27 (16) (2013) 1790–1810. [70] K. Wille, G.J. Parra-Montesinos, Effect of beam size, casting method, and
[42] Y. Farnam, S. Mohammadi, M. Shekarchi, Experimental and numerical support conditions on flexural behavior of ultra-high-performance fibre
investigations of low velocity impact behavior of high-performance fiber- reinforced concrete, ACI Mater. J. 109 (3) (2012) 379–388.
reinforced cement based composite, Int. J. Impact Eng. 37 (2) (2010) 220–229. [71] D.Y. Yoo, N. Banthia, S.T. Kang, Y.S. Yoon, Size effect in ultra-high performance
[43] N.T. Tran, T.K. Tran, D.J. Kim, High rate response of ultra-high-performance concrete beams, Eng. Fract. Mech. 157 (2016) 86–106.
fiber-reinforced concretes under direct tension, Cem. Concr. Res. 69 (2015) [72] D.Y. Yoo, N. Banthia, Size-dependent impact resistance of ultra-high
72–87. performance fibre-reinforced concrete beams, Constr. Build. Mater. 142
[44] G.H. Mahmud, Z. Yang, A.M. Hassan, Experimental and numerical studies of (2017) 363–375.
size effects of Ultra High Performance Steel Fibre Reinforced Concrete [73] D.L. Nguyen, D.J. Kim, G.S. Ryu, K.T. Koh, Size effect on flexural behavior of
(UHPFRC) beams, Constr. Build. Mater. 48 (2013) 1027–1034. ultra-high-performance hybrid fibre-reinforced concrete, Compos Part B Eng.
[45] R. Toledo Filho, E. Koenders, S. Formagini, E. Fairbairn, Performance 45 (1) (2013) 1104–1116.
assessment of ultra high performance fiber reinforced cementitious [74] D.L. Nguyen, G.S. Ryu, K.T. Koh, D.J. Kim, Size and geometry dependent tensile
composites in view of sustainability, Mater. Des. (1980–2015) 36 (2012) behavior of ultra-high-performance fibre-reinforced concrete, Compos Part B
880–888. Eng. 58 (2014) 279–292.
[46] R. Yu, P. Spiesz, H. Brouwers, Mix design and properties assessment of ultra- [75] M.Z. An, L.J. Zhang, Q.X. Yi, Size effect on compressive strength of reactive
high performance fibre reinforced concrete (UHPFRC), Cem. Concr. Res. 56 powder concrete, J. Chin. Univ. Min. Technol. 18 (2) (2008) 279–282.
(2014) 29–39. [76] A. Spasojavic, D. ReDaelli, M. Fernandez Ruiz, A. Muttoni, Influence of tensile
[47] W.-Y. Lim, S.-G. Hong, Shear Tests for Ultra-High Performance Fiber properties of UHPFRC on size effect in bending, in: Proceedings of the 2nd
Reinforced Concrete (UHPFRC) Beams with Shear Reinforcement, Int. J. International Symposium on Ultra High Performance Concrete. Kassel,
Concr. Struct. Mater. 10 (2) (2016) 177–188. Germany, 2008, pp. 303-310.
[48] P. Fidjestol, R. Thorsteinsen, P. Svennevig, Making UHPC with local [77] M. Lepech, V.C. Li, Brittle Matrix Composites-7, Warsaw, Poland, in:
materials—the way forward, Proc. HiPerMat (2012) 207–214. Preliminary Findings on Size Effect in ECC Structural Members in Flexural,
[49] K. Wille, A.E. Naaman, G.J. Parra-Montesinos, Ultra-high performance 2003, 57-66.
concrete with compressive strength exceeding 150 MPa (22 ksi): a simpler [78] S. Kazemi, A.S. Lubell, Influence of specimen size and fibre content on
way, ACI Mater. J. 108 (1) (2011). mechanical properties of ultra-high-performance fibre-reinforced concrete,
[50] R. Yu, P. Spiesz, H.J.H. Brouwers, Development of ultrahigh performance fibre ACI Mater J. 109 (6) (2012) 675–684.
reinforced concrete: towards an efficient utilization of binders and fibres, [79] K.H. Reineck, S. Greiner, Scale effect and combined loading of thin UHPFRC
Constr Build Mater. 79 (2015) 273–282. members, Adv. Constr. Mater. (2007) 211–218.
[51] M.A.A. Aldahdooh, N.M. Bunnori, M.A.M. Johari, Development of green [80] B. Frettl€ohr, K.H. Reineck, H.W. Reinhardt, Size and shape effect of UHPFRC
ultrahigh performance fibre reinforced concrete containing ultrafine palm prisms tested under axial tension and bending, in: Proceedings of High
oil fuel ash, Constr Build Mater. 48 (2013) 379–389. Performance Fibre Reinforced Cement Composites 6, Netherlands, 2012, 365-
[52] R. Yu, P. Tang, P. Spiesz, et al., A study of multiple effects of ano silica and 372.
hybrid fibers on the properties of ultrahigh performance fibre reinforced [81] Z.P. Bazant, R. Gettu, Size effect in concrete structures and influence of
concrete incorporating waste bottom ash, Constr Build Mater. 60 (2014) 98– loading rate, in: Proc., 1st Mat. Engrg. Congr., Serviceability and Durability of
110. Construction Materials ASCE, Boston, 1990, 1113–1123.
[53] F.U.A. Shaikh, T. Nishiwaki, S. Kwon, Effect of fly ash on tensile properties of [82] D.Y. Yoo, Y.S. Yoon, Structural performance of ultra-high-performance
ultra-high performance fiber reinforced cementitious composites (UHP- concrete beams with different steel fibres, Eng. Struct. 102 (2015) 409–423.
FRCC), J. Sustainable Cement-Based Mater. (2018), https://doi.org/10.1080/ [83] D.Y. Yoo, S.T. Kang, N. Banthia, Y.S. Yoon, Nonlinear finite element analysis of
21650373.2018.1514672. ultra-high-performance fibre-reinforced concrete beams, Int. J. Damage
[54] M. Aldahdooh, N.M. Bunnori, M.M. Johari, Evaluation of ultra-high- Mech. (2016).
performance-fiber reinforced concrete binder content using the response [84] T. Krauthammer, M.M. Elfahal, J. Lim, T. Ohno, M. Beppu, G. Markeset, Size
surface method, Mater. Des. 52 (2013) 957–965. effect for high-strength concrete cylinders subjected to axial impact, Int. J.
[55] P. Máca, R. Sovják, P. Konvalinka, Mix design of UHPFRC and its response to Impact Eng. 28 (9) (2003) 1001–1016.
projectile impact, Int. J. Impact Eng. 63 (2014) 158–163. [85] A.E. Naaman, New fibre technology: cement, ceramic and polymeric
[56] C. Magureanu, I. Sosa, C. Negrutiu, B. Heghes, Physical and mechanical composites, Concr. Int. 20 (7) (1998) 57–62.
properties of ultra high strength fiber reinforced cementitious composites, [86] A.E. Naaman, Engineered steel fibres with optimal properties for
Korea Concrete Institute, Fracture Mechanics of Concrete and Concrete reinforcement of cement composites, J. Adv. Concr. Tech. 1 (3) (2003) 241–
Structures, 2010, pp. 1497–11491. 252.
[57] H. Aoude, F.P. Dagenais, R.P. Burrell, M. Saatcioglu, Behavior of ultra-high [87] K. Wille, A.E. Naaman, Pullout behavior of high-strength steel fibres
performance fiber reinforced concrete columns under blast loading, Int. J. embedded in ultra-high-performance concrete, ACI Mater J. 109 (4) (2012)
Impact Eng. 80 (2015) 185–202. 479–488.
[58] H.K.S. Eldin, H.A. Mohamed, M. Khater, S. Ahmed, Mechanical Properties [88] V. Corinaldesi, G. Moriconi, Mechanical and thermal evaluation of Ultra High
ofUltra-High Performance Fiber Reinforced Concrete, (2014). Performance Fiber Reinforced Concretes for engineering applications, Constr.
[59] D.-Y. Yoo, N. Banthia, S.-W. Kim, Y.-S. Yoon, Response of ultra-high- Build. Mater. 26 (1) (2012) 289–294.
performance fiber-reinforced concrete beams with continuous steel [89] D.Y. Yoo, G. Zi, S.T. Kang, Y.S. Yoon, Biaxial flexural behavior of ultra high
reinforcement subjected to low-velocity impact loading, Compos. Struct. performance fibre-reinforced concrete with different fibre lengths and
126 (2015) 233–245. placement methods, Cem. Concr. Compos 63 (2015) 51–66.
[60] M. Xu, K. Wille, Fracture energy of UHP-FRC under direct tensile loading [90] D.Y. Yoo, S.T. Kang, Y.S. Yoon, Effect of fibre length and placement method on
applied at low strain rates, Compos. B Eng. 80 (2015) 116–125. flexural behavior, tension-softening curve, and fibre distribution
[61] P. Máca, R. Sovják, T. Vavřiník, Experimental investigation of mechanical characteristics of UHPFRC, Constr. Build. Mater 64 (2014) 67–81.
properties of UHPFRC, Procedia Eng. 65 (2013) 14–19. [91] D.Y. Yoo, N. Banthia, G. Zi, Y.S. Yoon, Comparative biaxial flexural behavior of
[62] D.-Y. Yoo, J.-J. Park, S.-W. Kim, Y.-S. Yoon, Influence of reinforcing bar type on ultra-high-performance fibre-reinforced concrete panels using two different
autogenous shrinkage stress and bond behavior of ultra high performance test and placement methods, J. Test. Eval. (2016).
fiber reinforced concrete, Cem. Concr. Compos. 48 (2014) 150–161. [92] Z. Yunsheng, S. Wei, L. Sifeng, J. Chujie, L. Jianzhong, Preparation of C200
[63] D.-Y. Yoo, N. Banthia, Y.-S. Yoon, Flexural behavior of ultra-high-performance green reactive powder concrete and its staticedynamic behaviors, Cem.
fiber-reinforced concrete beams reinforced with GFRP and steel rebars, Eng. Concr. Compos 30 (9) (2008) 831–838.
Struct. 111 (2016) 246–262. [93] S.T. Kang, Y. Lee, Y.D. Park, J.K. Kim, Tensile fracture properties of an ultra high
[64] J. Xia, Ultra-high Performance Fiber Reinforced Concrete In Bridge Deck performance fibre reinforced concrete (UHPFRC) with steel fibre, Compos
Applications, (2011). Struct. 92 (1) (2010) 61–71.
[65] K. Wille, S. El-Tawil, A. Naaman, Properties of strain hardening ultra high [94] S.L. Prabha, J.K. Dattatreya, M. Neelamegam, M.V. Seshagirirao, Study on
performance fiber reinforced concrete (UHP-FRC) under direct tensile stress-strain properties of reactive powder concrete under uniaxial
loading, Cem. Concr. Compos. 48 (2014) 53–66. compression, Int. J. Eng. Sci. Technol. 2 (11) (2010) 6408–6416.
F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152 27

[95] L. Ferrara, N. Ozyurt, M. Di Prisco, High mechanical performance of fibre [122] L. Mao, S.J. Barnett, A. Tyas, J. Warren, G.K. Schleyer, S.S. Zaini, Response of
reinforced cementitious composites: the role of ‘‘casting-flow induced” fibre small scale ultra high performance fibre reinforced concrete slabs to blast
orientation, Mater Struct. 44 (1) (2011) 109–128. loading, Constr. Build. Mater. 93 (2015) 822–830.
[96] S.H. Kwon, S.T. Kang, B.Y. Lee, J.K. Kim, The variation of flow-dependent [123] P. Máca, R. Sovják, T. Vavriník, Resistance of ultra high performance fibre
tensile behavior in radial flow dominant placing of Ultra High Performance reinforced concrete to projectile impact, Struct. Under Shock Impact XII 126
Fibre Reinforced Cementitious Composites (UHPFRCC), Constr. Build. Mater (2013) 261–272.
33 (2012) 109–121. [124] Y.S. Tai, Flat ended projectile penetrating ultra-high strength concrete plate
[97] D.Y. Yoo, N. Banthia, S.T. Kang, Y.S. Yoon, Effect of fibre orientation on the rate target, Theor. Appl. Fract. Mech. 51 (2) (2009) 117–128.
dependent flexural behavior of ultra-high-performance fibre-reinforced [125] R. Adeline, M. Lachemi, P. Blais, (1998). Design and behavior of the
concrete, Compos Struct. (2016). Sherbrooke Footbridge. International Symposium on High-Performance and
[98] B. Boulekbache, M. Hamrat, M. Chemrouk, S. Amziane, Flowability of fibre Reactive Powder Concretes, Sherbrooke, Canada.
reinforced concrete and its effect on the mechanical properties of the [126] S. Deem, (2001). Concrete Attraction-Something New on the French Menu-
material, Constr. Build. Mater 24 (9) (2010) 1664–1671. Concrete.
[99] S.T. Kang, B.Y. Lee, J.K. Kim, Y.Y. Kim, The effect of fibre distribution [127] J. Resplendino, First recommendations for Ultra-High Performance Concretes
characteristics on the flexural strength of steel fibre-reinforced ultra high and examples of applications, Proceedings of the International Symposium
strength concrete, Constr. Build. Mater 25 (5) (2011) 2450–2457. on Ultra High Performance Concrete, pp.79-89, Kassel, Germany, 2004.
[100] G. Zi, H. Oh, S.K. Park, A novel indirect tensile test method to measure the [128] P. Acker, M. Behloul, Ductal Technology: a Large Spectrum of Properties, a
biaxial tensile strength of concretes and other quasibrittle materials, Cem. Wide Range of, Proceedings of the International Symposium on Ultra High
Concr. Res. 38 (6) (2008) 751–756. Performance Concrete, Kassel, 2004, pp. 11–23.
[101] S.T. Kang, J.K. Kim, Numerical simulation of the variation of fibre orientation [129] H. Park, E. Chuang, F.-J. Ulm, Model-Based Optimization of Ultra High
distribution during flow molding of Ultra High Performance Cementitious Performance Concrete Highway Bridge Girders , Massachusetts Institute of
Composites (UHPCC), Cem. Concr. Compos 34 (2) (2012) 208–217. Technology, Cambridge, USA, March 2003.
[102] G.B. Jeffery, The motion of ellipsoidal particles immersed in a viscous fluid, [130] A.E. Naaman, K. Chandrangsu, Innovative bridge deck system using high
Proc. Roy. Soc. Lond. 102 (715) (1922) 161–179. performance fiberreinforced cement composites, ACI Struct. J. 101 (1) (2004)
[103] B.Y. Lee, J.K. Kim, Y.Y. Kim, Prediction of ECC tensile stresse strain curves 57–64.
based on modified fibre bridging relations considering fibre distribution [131] B.A. Graybeal, J.L. Hartmann, Construction of an optimized UHPC vehicle
characteristics, Comput. Concr. 7 (5) (2010) 455–468. bridge, Seventh International Symposium on the utilization of High-Strength/
[104] N.T. Tran, D.J. Kim, High energy absorption capacity of ultra-high HIgh-Performance Concrete, ACI Special Publication SP-228, Vol. 2, pp. 1109-
performance hybrid-fibre-reinforced concretes under direct tensile impact, 1118, Washington D.C., USA, June, 2005.
in: 4th International Symposium on Ultra-High Performance Concrete and [132] E. Brühwiler, E. Denarié, K. Habel, Ultra-High Performance Fibre Reinforced
High Performance Construction Materials, Kassel, Germany, 2016, 8. Concrete for advanced rehabilitation of bridges, Proceedings Fib Symposium
[105] J.-P. Charron, E. Denarié, E. Brühwiler, Transport properties of water ‘‘Keep Concrete Attractive”, Budapest, Hungary, Editors G. L. Balasz, A.
and glycol in an ultra high performance fiber reinforced concrete Borosnyoi, pp. 951-956, Budapest, Hungary, 2005.
(UHPFRC) under high tensile deformation, Cem. Concr. Res. 38 (5) [133] B. Cavill, G. Chirgwin, The world first Reactive Powder Concrete road bridge
(2008) 689–698. at Shepherds Gully Creek, NSW, 21th Biennal Conference of the Concrete
[106] D.Y. Yoo, N. Banthia, Y.S. Yoon. Impact resistance of ultra-high-performance Institute of Australia, Brisbane, 17-19 July, 11 pp., Brisbane, Australia, July,
fibre-reinforced concrete with different steel fibres. 9th RILEM International 2003.
Symposium on Fibre Reinforced Concrete (BEFIB 2016), Pacific Gateway [134] A. Meda, G. Rosati, Design and construction of a bridge in very high
Hotel, Vancouver, BC, Canada, 2016. performance fiber-reinforced concrete, ASCE Journal of Bridge Engineering,
[107] Y.S. Tai, S. El-Tawil, T.H. Chung, Performance of deformed steel fibres Vol. 8, No. 5, pp. 281-287, September-Octobre, 2003.
embedded in ultra-high performance concrete subjected to various pullout [135] C. Ferraris, E. Garboczi, P. Stutzman, J. Winpigler, J. Clifton, Influence of silica
rates, Cem. Concr. Res. 89 (2016) 1–13. fume on the stresses generated by alkali-silica reaction, Cem., Concr.
[108] S. Pyo, S. El-Tawil, A.E. Naaman, Direct tensile behavior of ultra high Aggregates 22 (1) (2000) 73–78.
performance fibre reinforced concrete (UHP-FRC) at high strain rates, Cem. [136] Z.S. Metaxa, M.S. Konsta-Gdoutos, S.P. Shah, Mechanical properties and
Concr. Res. 88 (2016) 144–156. nanostructure of cement-based materials reinforced with carbon nanofibers
[109] R. Yu, P. Spiesz, H.J.H. Brouwers, Static properties and impact resistance of a and Polyvinyl Alcohol (PVA) microfibers, in Advances in the Material Science
green Ultra-High Performance Hybrid Fibre Reinforced Concrete (UHPHFRC): of Concrete, in: Session at the ACI Spring 2010 Convention, 2010, pp. 115–
experiments and modeling, Constr. Build. Mater. 68 (2014) 158–171. 126.
[110] D.Y. Yoo, S.W. Kim, J.J. Park, Comparative flexural behavior of ultra- [137] B.A. Tayeh, B. Abu Bakar, M. Johari, A. Zeyad, The role of silica fume in the
highperformance concrete reinforced with hybrid straight steel fibres, adhesion of concrete restoration systems, Adv. Mater. Res., Trans. Tech. Publ.
Constr. Build. Mater. 132 (2017) 219–229. (2013) 265–269.
[111] S.G. Millard, T.C.K. Molyneaux, S.J. Barnett, X. Gao, Dynamic enhancement of [138] M. Jalal, A. Pouladkhan, O.F. Harandi, D. Jafari, Comparative study on effects of
blast-resistant ultra high performance fibre-reinforced concrete under Class F fly ash, nano silica and silica fume on properties of high performance
flexural and shear loading, Int. J. Impact Eng. 37 (4) (2010) 405–413. self compacting concrete, Constr. Build. Mater. 94 (2015) 90–104.
[112] L. Flanders, T. Rushing, E. Landis, Energy dissipation mechanisms in the [139] N. González, J. Castaño, Y. Alvarado, I. Gasch, Influence of fiber volume and
fracture of fibre reinforced ultra high performance concrete, in: 4th subsequent curing on post-crack behavior of an ultra high performance
International Symposium on Ultra-High Performance Concrete and High concrete (UHPC), Revista Ingeniería de Construcción 29 (3) (2015) 220–233.
Performance Construction Materials (HiPerMat2016), Kassel, Germany, 2016, [140] B. Karihaloo, CARDIFRC–from concept to industrial application, High
1–8. Performance Fiber Reinforced Cement Composites 6, Springer, 2012, pp.
[113] Z. Rong, W. Sun, Experimental and numerical investigation on the dynamic 397–404.
tensile behavior of ultra-high performance cement based composites, Constr. [141] I.H. Yang, C. Joh, B.-S. Kim, Structural behavior of ultra high performance
Build. Mater. 31 (2012) 168–173. concrete beams subjected to bending, Eng. Struct. 32 (11) (2010) 3478–3487.
[114] D.Y. Yoo, S.T. Kang, Y.S. Yoon, Enhancing the flexural performance of [142] W. Zhang, Y. Zhang, G. Zhang, Single and multiple dynamic impacts
ultrahigh- performance concrete using long steel fibres, Compos. Struct. behaviour of ultra-high performance cementitious composite, J. Wuhan
147 (2016) 220–230. Univ. Technol.-Mater. Sci. Ed. 26 (6) (2011) 1227–1234.
[115] J. Lai, X. Guo, Y. Zhu, Repeated penetration and different depth explosion of [143] B.A. Graybeal, F. Baby, Development of direct tension test method for ultra-
ultra-high performance concrete, Int. J. Impact Eng. 84 (2015) 1–12. high-performance fiber-reinforced concrete, ACI Mater. J. 110 (2) (2013).
[116] L. Jianzhong, Z. Yaoyong, W. Huifang, Penetration and explosion of ultra-high [144] K. Wille, D.J. Kim, A.E. Naaman, Strain-hardening UHP-FRC with low fiber
performance fibre reinforced cement composite subjected to impact, in: contents, Mater. Struct. 44 (3) (2011) 583–598.
Proceedings of High Performance Fibre Reinforced Cement Composites [145] R. Sovják, T. Vavřiník, J. Zatloukal, P. Máca, T. Mičunek, M. Frydrýn, Resistance
(HPFRCC7), Stuttgart, Germany, 2015, pp. 325–332. of slim UHPFRC targets to projectile impact using in-service bullets, Int. J.
[117] R. Ranade, V.C. Li, W.F. Heard, B.A. Williams, Impact resistance of high Impact Eng. 76 (2015) 166–177.
strength-high ductility concrete, Cem. Concr. Res. 98 (2017) 24–35. [146] K. Habel, M. Viviani, E. Denarié, E. Brühwiler, Development of the mechanical
[118] Z. Yunsheng, S. Wei, L. Sifeng, J. Chujie, L. Jianzhong, Preparation of C200 properties of an ultra-high performance fiber reinforced concrete (UHPFRC),
green reactive powder concrete and its static–dynamic behaviors, Cem. Cem. Concr. Res. 36 (7) (2006) 1362–1370.
Concr. Compos. 30 (9) (2008) 831–838. [147] B.A. Tayeh, B.A. Bakar, M.M. Johari, Y.L. Voo, Mechanical and permeability
[119] S. Pyo, S. El-Tawil, Capturing the strain hardening and softening responses of properties of the interface between normal concrete substrate and ultra high
cementitious composites subjected to impact loading, Constr. Build. Mater. performance fiber concrete overlay, Constr. Build. Mater. 36 (2012) 538–548.
81 (2015) 276–283. [148] L.K. Askar, B.A. Tayeh, B. Abu Bakar, Effect of different curing conditions on
[120] S. Pyo, K. Wille, S. El-Tawil, A.E. Naaman, Strain rate dependent properties of the mechanical properties of UHPFC, Awam International Conference on Civil
ultra high performance fibre reinforced concrete (UHP-FRC) under tension, Engineering (AICCE’12) and Geohazard Information Zonation (GIZ’12), 2012,
Cem. Concr. Compos. 56 (2015) 15–24. pp. 28-30.
[121] D.Y. Yoo, N. Banthia, S.T. Kang, Y.S. Yoon, Effect of fibre orientation on the [149] A. Hassan, S. Jones, G. Mahmud, Experimental test methods to determine the
rate-dependent flexural behavior of ultra-high-performance fibre-reinforced uniaxial tensile and compressive behaviour of ultra high performance fibre
concrete, Compos. Struct. 157 (2016) 62–70. reinforced concrete (UHPFRC), Constr. Build. Mater. 37 (2012) 874–882.
28 F.U.A. Shaikh et al. / Construction and Building Materials 232 (2020) 117152

[150] A. Hassan, S. Jones, Non-destructive testing of ultra high performance fibre application to retrofitting, Eng. Fract. Mech. 74 (1) (2007) 151–167.
reinforced concrete (UHPFRC): a feasibility study for using ultrasonic and [158] D.-Y. Yoo, J.-H. Lee, Y.-S. Yoon, Effect of fiber content on mechanical and
resonant frequency testing techniques, Constr. Build. Mater. 35 (2012) 361– fracture properties of ultra high performance fiber reinforced cementitious
367. composites, Compos. Struct. 106 (2013) 742–753.
[151] J.-P. Charron, E. Denarié, E. Brühwiler, Permeability of ultra high performance [159] K. Fujikake, T. Senga, N. Ueda, T. Ohno, M. Katagiri, Study on impact response
fiber reinforced concretes (UHPFRC) under high stresses, Mater. Struct. 40 (3) of reactive powder concrete beam and its analytical model, J. Adv. Concr.
(2007) 269–277. Technol. 4 (1) (2006) 99–108.
[152] F.S. Ahmad, G. Foret, R. Le Roy, Bond between carbon fibre-reinforced [160] H. Othman, H. Marzouk, Impact response of ultra-high-performance
polymer (CFRP) bars and ultra high performance fibre reinforced concrete reinforced concrete plates, ACI Struct. J. 113 (6) (2016) 1325.
(UHPFRC): Experimental study, Constr. Build. Mater. 25 (2) (2011) 479–485. [161] D.-Y. Yoo, S.-T. Kang, J.-H. Lee, Y.-S. Yoon, Effect of shrinkage reducing
[153] K. Wille, A. Naaman, Fracture energy of UHPFRC under direct tensile loading, admixture on tensile and flexural behaviors of UHPFRC considering fiber
FraMCoS-7 international conference, Jeju, Korea, 2010. distribution characteristics, Cem. Concr. Res. 54 (2013) 180–190.
[154] A. Kamen, Time dependent behaviour of ultra high performance fibre [162] K. Habel, E. Denarié, E. Brühwiler, Time dependent behavior of elements
reinforced concrete (UHPFRC), 6th International PhD Symposium, Civ. Eng. combining ultra-high performance fiber reinforced concretes (UHPFRC) and
(2006). reinforced concrete, Mater. Struct. 39 (5) (2006) 557–569.
[155] J. Feng, W. Sun, Z. Liu, C. Cui, X. Wang, An armour-piercing projectile [163] S.H. Park, D.J. Kim, G.S. Ryu, K.T. Koh, Tensile behavior of ultra high
penetration in a double-layered target of ultra-high-performance fiber performance hybrid fiber reinforced concrete, Cem. Concr. Compos. 34 (2)
reinforced concrete and armour steel: Experimental and numerical (2012) 172–184.
analyses, Mater. Des. 102 (2016) 131–141. [164] A. Kamen, E. Denarié, H. Sadouki, E. Brühwiler, UHPFRC tensile creep at early
[156] D.-Y. Yoo, K.-Y. Kwon, J.-J. Park, Y.-S. Yoon, Local bond-slip response of GFRP age, Mater. Struct. 42 (1) (2009) 113.
rebar in ultra-high-performance fiber-reinforced concrete, Compos. Struct. [165] M.L. Nehdi, S. Abbas, A.M. Soliman, Exploratory study of ultra-high
120 (2015) 53–64. performance fiber reinforced concrete tunnel lining segments with varying
[157] F. Farhat, D. Nicolaides, A. Kanellopoulos, B.L. Karihaloo, High performance steel fiber lengths and dosages, Eng. Struct. 101 (2015) 733–742.
fibre-reinforced cementitious composite (CARDIFRC)–Performance and

Potrebbero piacerti anche