You are on page 1of 65

Accepted Manuscript

Basalt-based fiber-reinforced materials and structural applications in civil en-


gineering

Elisabetta Monaldo, Francesca Nerilli, Giuseppe Vairo

PII: S0263-8223(18)34225-9
DOI: https://doi.org/10.1016/j.compstruct.2019.02.002
Reference: COST 10627

To appear in: Composite Structures

Please cite this article as: Monaldo, E., Nerilli, F., Vairo, G., Basalt-based fiber-reinforced materials and structural
applications in civil engineering, Composite Structures (2019), doi: https://doi.org/10.1016/j.compstruct.
2019.02.002

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Basalt-based fiber-reinforced materials and
structural applications in civil engineering
Elisabetta Monaldoa,∗, Francesca Nerillia , Giuseppe Vairob
a
Università degli Studi Niccolò Cusano, 00166 Rome, Italy
b
Università degli Studi di Roma “Tor Vergata”- Department of Civil Engineering and
Computer Science (DICII), 00133 Rome, Italy

Abstract
In the last decades, a growing interest in using basalt as reinforcement
for composite materials has emerged, since promising physico-chemical and
mechanical properties of basalt products, as well as good processability fea-
tures and cost-effectiveness of the corresponding production technologies. In
particular, basalt fibers allow to define composite materials really compet-
itive with those obtained by employing traditional glass or carbon fibers.
Recent experimental programs and analytical approaches reveal that basalt-
based fiber-reinforced materials may be effective for a number of structural
applications in civil engineering. Depending on the fiber treatment and ar-
rangement, as well as on the matrix type (polymeric or cementitious) different
composite materials have been conceived. For instance, strengthening and
retrofitting of existing structures (both concrete and masonry) may be per-
formed through basalt-based fiber-reinforced polymers (BFRP) and cementi-
tious matrices (BFRCM), as well as novel design concepts can be exploited by
referring to basalt-based rebars and fiber-reinforced concrete (BFRC). This
paper aims to furnish a systematic review of the state of the art on basalt
fibers, basalt-based composite materials and their applications in civil engi-
neering field, by tracing main available evidence and highlighting perspective
aspects and open problems.
Keywords: Basalt fibers, Basalt-based composites, BFRP, BFRCM, BFRP


Corresponding author
Email addresses: elisabetta.monaldo@unicusano.it (Elisabetta Monaldo),
francesca.nerilli@unicusano.it (Francesca Nerilli), vairo@ing.uniroma2.it
(Giuseppe Vairo)

Preprint submitted to Composite Structures January 17, 2019


bars, BFRC.

1. Introduction
The use of composite materials, due to their interesting physical and
mechanical properties, is nowadays widely spread in many engineering fields,
with advanced technical applications in civil, mechanical, aerospace, and
5 biomedical scenarios.
As a matter of fact, the development of composite materials has been a
topic of intensive interest in the last decades, but the concept of using two
or more elemental materials combined to form the constituent phases of a
composite solid has been employed ever since structural design processes were
10 consciously or unconsciously first used [1–3]. The ancient Jewish workers
during their tenure under the Pharaohs mixed chopped straws and clay for
producing bricks with enhanced mechanical strength. The Japanese Samurai
warriors forged their swords through laminated metals, by employing strips
of steel for strength and strips of iron for flexibility. The Mongols (1200 AD)
15 invented the first composite bow obtained by combining wood, bone, and
animal glue.
Within the framework of modern advanced composite materials, arised
with the plastic era, one of the most important type, from a structural point
of view, is the fiber-reinforced type. In this case, the use of fibers with spe-
20 cific physical properties and arranged following suitable patterns within an
embedding matrix, allows to design functional materials characterized by de-
sired levels of strength and stiffness, combined with effective chemical and
physical properties. The matrix is generally constituted by polymeric resins
(e.g., epoxy, vinylester or polyester thermosetting plastic), although metals
25 (e.g., aluminum) or mortar/concrete are also used in specific structural appli-
cations. Referring to widespread fiber-reinforced materials, fibers are mainly
made of glass, carbon, aramid compounds, although other fibers can be used
for structural applications. In this context, reference can be made to veg-
etable fibers (based, for instance, on cotton, hemp, jute, and flax), wood
30 fibers (generally distinguished from vegetable ones), mineral fibers (such
those based on asbestos) [4–6].
In addition, the use of fibers produced by molten basalt rock has gained in
the last years an increasing interest, as a result of the good physico-chemical
properties that can be achieved [7, 8]. In fact, basalt fibers exhibit good

2
35 strength and stiffness properties also at high temperatures, long-term dura-
bility, high acid and solvent resistance, low water absorbtion, remarkable
heat and sound insulation properties, good processability, as well as their
fabrication process is generally significantly cheaper than carbon and glass
fibers [8–13]. Moreover, basalt is widely available in nature with well-defined
40 chemical and mineralogical composition [14], it has non-toxic reactions with
air or water, is non-combustible, is explosion-proof, and it is generally charac-
terized by high levels of eco-compatibility and recyclability, resulting a high-
performance green inorganic material. Therefore, basalt fibers and composite
materials made from them can be retained to have a successful relationship
45 between quality and costs in comparison with other types of fibers such as
those based on glass and carbon. Accordingly, although basalt is not a new
material (since the Roman age basalt was employed as paving and build-
ing stone), its modern application can be surely retained as innovative and
convenient in many industrial and structural fields. In particular, different
50 types of rovings, chopped strands and twisted yarns can be produced by
using basalt filaments. Various kinds of woven and non-woven materials,
unidirectional and multiaxial fabrics can be also supplied. All these prod-
ucts are fully compatible with different types of resins (e.g., epoxy, phenolic,
polyester, vinylester systems), allowing to produce basalt-fiber-reinforced-
55 polymers (BFRP), and they do not suffer for possible releasing of breathable
particles dangerous for human health (as for asbestos).
The application fields of basalt-fiber products are extremely broad, in-
cluding implementations for automotive and aeronautic industry, naval and
civil constructions, sporting equipments, wind turbine systems [8, 15]. As
60 regards automotive applications, high-quality basalt rovings, fabrics and
chopped strands are often used, for instance, in the production of fuel stor-
age systems, brake pads, mufflers, and headliners. Moreover, basalt rovings
are highly suitable for realizing different sporting equipments, including skis,
snowboards, and bicycles. Since their good electrical insulating properties,
65 basalt fibers are also used for printed circuit boards, as well as for extra
fine resistant insulation of electrical cables and underground ducts. Due to
the high resistance to both corrosion and UV-radiation, combined with good
mechanical properties, BFRP composites and basalt-fiber-based woven and
fabrics are successfully employed as structural reinforcements for producing
70 wind turbine blades and in boat building.
A wide range of products are also available for civil applications [16].
Basalt fibers are used for realizing innovative building materials, defined as

3
the proper mixture of a concrete matrix with short (staple) or cut (chopped)
basalt fibers, so to improve the material strength-to-weight ratio [3]. Fiber-
75 basalt-based composites are employed in order to replace asbestos in almost
all its possible applications. Basalt filaments and fibers are adopted for pro-
ducing: high-strength rovings for pultruded load-bearing parts and concrete-
reinforcing bars; woven fabrics for heat/sound insulation and fire protection;
geogrids for road and land reinforcement; stucco nets for wall reinforcing and
80 renovation [16].
The increasing use of composite materials based on basalt fibers in the
field of civil construction, needs a suitable regulatory framework defining
technical specifications and design rules, by discriminating the cases of ap-
plication on new or existing structures. To date, European technical rules
85 and many national guidelines refer to composite materials in general, or at
most they refer to traditional composites (based on glass, carbon or aramid)
for which an extensive experimental investigation is available [17]. Aiming
to furnish similarly contributions and evidence for basalt-based composites
in civil applications, a great research effort has been recently emerged lead-
90 ing to a number of experimental and analytical studies. Among them firts
attempts towards the definition of specific technical guidelines can be found
in [18–32].
In order to gather in a consistent and comparative way the most recent
results proposed via both experimental and analytical approaches, present
95 paper proposes an updated review of the state of the art concerning the use of
basalt-based fiber-reinforced composite materials for structural applications
in civil engineering. The paper is organized as follows. The first part (Section
2) focuses on the basalt fibers, tracing - after an historical background -
the main aspects concerning production technologies and physico-chemical
100 properties. In the second part (Sections 3 to 6), the mechanical behaviour
and the physical properties of different composite materials realized through
basalt fibers are investigated, elucidating their main involvement in civil
applications and discussing the most relevant results recently proposed.

2. Basalt fibers
105 2.1. Background
The word basalt is derived from Late Latin basaltes, misspelling of the
latin word basanites (namely, very hard stone), which was imported from An-
cient Greek βασανι0 τ ηζ and originated in Egyptian bauhun (namely, slate).

4
The modern petrological term, basalt, describing a particular composition of
110 lava-derived rock, originates from its use by Georgius Agricola in 1556 in his
work of mining and mineralogy De re metallica, wherein Agricola denoted as
basalt the volcanic black and hard rock of the Schloßberg (local castle hill)
at Stolpen [33].
Basalt is a dark-coloured mafic extrusive rock and it is the most widespread
115 of all igneous rocks, resulting in more than 90% of all volcanic rocks. The
rate of cooling of molten lava strongly influences the microstructure of this
rock: when the rate is low, basalt structure exhibits an almost regular atomic
arrangement, whereas an amorphous (uncrystallized) structure arises when
the solidification rate is high.
120 From a mineralogical point of view, basalt is mainly composed of three
silicate minerals: plagioclases, pyroxenes and olivines. Its chemical compo-
sition depends on the nature of the magma it derives from; in each case,
silicon dioxide (SiO2 ) is the main component, followed by Al2 O3 , FeO +
Fe2 O3 , CaO, MgO, Na2 O, K2 O, TiO2 , and so on. The percentage of chem-
125 ical composition of basalt is strictly related to the geographic area. Basalt
composition is characterized by [8, 13]: 45-52% of SiO2 , 9-19% of Al2 O3 ,
6-15% of FeO + Fe2 O3 , 5-13% of CaO, 6-12% of MgO, 2-11% of Na2 O and
K2 O, and 0.5-2% of TiO2 . Basalt rocks, widely diffused all around the world,
are classified according to the SiO2 content as: alkaline (SiO2 up to 42%),
130 mildly acidic (43 to 46% SiO2 ) and acidic basalts (over 46% SiO2 ). Only
acidic basalts are suitable for fiber production, since a high content of SiO2
allows to obtain good levels of flexibility and chemical stability of the basalt
filaments. As regards the other components, Al2 O3 can increase chemical re-
sistance of the basalt filaments, as well as CaO, MgO and TiO2 can improve
135 the water resistance and corrosion resistance [13].
For many years, basalt has been used in casting processes to make tiles
and slabs for architectural applications. Due to its high abrasion resistance,
basalt is also adopted in industrial applications to realize cast basalt liners
for steel tubing. In crushed form, basalt finds use as aggregate in concrete.
140 Moreover, in the last decade basalt has emerged as a suitable and competitive
material to produce the reinforcement phase, mainly as fibers, in composite
materials [8, 13].
The first attempts to transform basalt rock into fibers by extrusion started
at the beginning of the 1920s and were attributed to the French Paul Dhè,
145 to which a U.S. Patent in 1922 [34] was granted. Around 1960, Soviet Union
(U.S.S.R.) began to investigate basalt fiber applications too, particularly for

5
military and aerospace purposes, succeeding in developing the first attempt
of production technology for continuous basalt fibers. Although in the 1970s
market and industrial strategies in United States changed mainly towards
150 the development of glass fibers rather than basalt ones, U.S. research on
basalt-fiber production continued. In 1979 a U.S. Patent was granted to
Austin and Subramanian [35], in which a process for increasing the tensile
strength properties of basalt fibers by adding both ferrous and ferric ox-
ides to the natural molten rock was proposed. Subramanian of Washington
155 State University and its co-workers correlated also the chemical composition
of basalt with the conditions for extrudability and physico-chemical charac-
teristics of the resulting fibers. In detail, they showed that silane coupling
agents can be effective for improving bond strength of basalt fibers in basalt
fiber-polymer systems [36–38], alumina addition can induce an improvement
160 in fiber strength [39], as well as hydrate zirconia films formed by sol-gel
process allow to enhance the alkali resistance fo basalt fibers [40]. In 1985
another granted U.S. Patent proposed a new chemical modification of basalt
fibers composition, modified with alkaline earth metal oxides for the use in
ceiling tile or boards [41]. Also U.S.S.R. scientists and technicians continued
165 their activities on basalt filaments and in 1985 the first industrial installation
for basalt continuous fibers production (the Teplozvukoizoliacia factory) was
realized near Kiev. Scientific traces of U.S.S.R. activity in those years can
be found, for instance, in [42–45]. In these pioneering studies, useful indi-
cations towards an effective fiber production process were given, in terms of
170 working and melting temperature ranges, viscosity levels and wetting ability
of molten basalt, crystallization features, as well as the influence of some
processing parameters on the mechanical and chemical properties of basalt
fibers began to be drawn. After the breakup of the Soviet Union in 1991, the
results of Soviet research were made available for civilian applications, and
175 scientific research institutes, as well as Russian factories (famous were those
in Berdyansk and Sudogda) worked for producing basalt continuous fibers
and basalt composites to export in Europe, U.S., Canada and Japan [46]. A
part U.S. and Russian countries (e.g., Ukraine, Georgia, Russia), after the
1990s the production of basalt fibers began also in other regions in the world.
180 Many different methods and technologies for producing basalt fibers have
been proposed in the last two decades. Just to report some relevant examples,
in the U.S. Patent by Aslanova [47] the proposed production strategy aimed
to shorten the industrial cycle and to increase the fiber resistance and thermal
endurance. It was mainly based on a preliminary heating of basalt rock,

6
185 before charging it into the melting furnace, and on a further stabilization
process to obtain a glass mass with a prescribed composition. In two U.S.
Patents granted to Brik [48, 49], a multifunctional apparatus and method
were proposed to produce high quality continuous amorphous basalt fibers
with flexible/ductile properties, from 7 to 100 µm in diameter, without traces
190 of crystalline phases and where platinum-rhodium metals (which limit the
efficiency of basalt fibers production and their commercial applications) were
replaced by corrosion resistant ceramic materials.
The research development of basalt fibers as structural elements in the
construction and transport field from 2002 to 2008 in China, is summarised
195 in [50]. The increase of the number of publications about this topic in these
years, grown of more than the half from the 2006 to the 2008 (from 20 to 70)
is there interestingly discussed. To confirm the interest in the manufacturing
technologies of basalt fibers, the trend of the patents published between 1997
to 2008 is also reported, with a great number in correspondence of the period
200 2006-2008, the production improvement reflecting into the increase of the
mechanical properties of the composites based on basalt fibers in terms of
tensile strength and elastic stiffness.
Finally, the interest of the research community is clearly highlighted by
the significant increase in the number of scientific publications in the last
205 three decades. The results, obtained through a search on Google Scholar of
the keywords basalt fiber, are more than 2230 in the nineties, more than 9160
in the first decade of the twenty first century and more than 16800 from 2010
until now.
The use of basalt fibers for civil applications, mainly addressed in this
210 paper, began around the 1990s, finding employment for strengthening and
retrofitting of civil structures, for thermal and acoustic insulation, as well as
for conceiving strategies of vibration damping. Many studies also confirmed
the feasibility of using composite bars made with basalt fibers to replace the
steel bars in reinforced-concrete (RC) structures [51].
215 As regards the market of basalt fibers, the economic pattern was unstable
and characterized by an oscillating marketplace until 1997 [51] because of the
fact that the production of continuous basalt roving and of the composites
made with basalt fibers were manufactured, on an industrial scale, only in
Ukraine.
220 Nowadays, the production of basalt fibers is largely widespread and en-
hanced, so the market is more stable and competitive. From data reported
in [52], it emerges that the largest application of continuous basalt fibers is

7
related to building and construction field for about 37% of the total demand
in 2012. Actually, the market is dominated by Russian and Ukraine play-
225 ers, due to high research activities and large basalt reserves, and the main
consumers are, in relevance order, North America, Asia Pacific and Europe.

2.2. Production technologies


Basalt fibers are obtained after melting raw material at very high temper-
ature (in the range of about 1200◦ C - 1500◦ C). However, unlike what happens
230 for glass fibers, production of basalt fibers doesn’t require particular prelim-
inary operations, such as constituents dosage. Basalt composition, indeed, is
intrinsically defined in the rock itself, and this simplification strongly reduces
the production costs [13].
Basalt fibers can be produced by means of two different technologies,
235 namely the Spinneret technology [2, 13] and the Junkers one [2, 53]. The
first is used to produce continuous fibers, whereas the latter allows to obtain
short fibers.
As regards Spinneret technology (see Figure 1(a)) basalt fibers can be
spun in a way which is substantially similar to that used for glass fibers.
240 Basalt rock is crushed and carefully washed before it is loaded into furnaces
heated using air gas mixture or electrically. Some electrodes are generally
immersed into the melting bath in order to promote an uniform heating
and to reduce the time for achieving thermal equilibrium. Once the melting
process is completed, molten basalt is poured onto platinum-rhodium heated
245 bushings, from which filaments are drawn under hydrostatic pressure. After
the cooling process, filaments are collected together to form a strand, whose
integrity and chemical stability are ensured through a preliminary lubrication
step. The filament size (in the order of some micrometers in diameter) is
controlled by varying the drawing speed and the melt temperature.
250 As confirmed in [54], main technological parameters of the continuous-
fiber production and main physical-chemical characteristics of continuous
basalt fibers are essentially independent from the place of provenance of the
starting raw material, provided that the content of SiO2 is in the range of
45% to 50%.
255 Junkers technology (see Figure 1(b)) consists on a melt-blowing process.
The manufacturing system is composed by three rotating cylinders with hor-
izontal axis and a nozzle plate perpendicular to the cylinders. Molten basalt
is poured on the upper rotating cylinder (namely, acceleration shaft). As an
effect of the tangential speed, molten basalt is subsequently conveyed on the

8
(a)

(b)

Figure 1: Simplified schemes of basalt fiber production methods: (a) Spinneret technology;
(b) Junkers technology.

9
260 underlying two cylinders (fibrillation shafts). Due to the centrifugal forces,
molten basalt meatus tends to break detaching in little drops. Compressed
air jets, coming from the nozzles behind the cylinders, turn the drops into
a thin and oblong shapes, that after the cooling process lead to the short
basalt fibers. It is worth observing that, when the air jets promote the drop
265 elongation, one or both the ends of each fiber may be characterized by a little
round head. As widely debated in the literature [13, 53, 55], this occurrence
may induce two counteracting effects. On one hand, fibers with heads are
more difficult to be pulled out from the matrix and therefore, the mechan-
ical properties of the resulting composites tends to be generally enhanced.
270 On the other hand, the occurrence of fiber heads may induce stress/strain
localizations leading to possible damage onset and propagation, and thereby
resulting in a possible degradation of the mechanical properties of the fiber-
reinforced composite material.

2.3. Mechanical properties


275 Basalt fibers exhibit an elastic behavior in tension, practically character-
ized by a linear stress-strain relationship until brittle failure [56, 57].
The influence of geometrical features and chemical composition, in terms
of SiO2 and Al2 O3 content, on the mechanical properties of basalt fibers has
been investigated in [57], by analysing fibers provided by different manu-
280 facturers and in comparison with the case of glass fibers. Furthermore, a
correlation between some chemical elements, such as SiO2 and the Al2 O3 ,
with the fiber tensile strength was also highlighted in [57–59]. The influence
of Al2 O3 on the fiber tensile strength was investigated by Gutnikov et al.
[60], showing that an increase in the content of Al2 O3 from 10% to 24% leads
285 to an improvement of the average tensile strength for basalt fibers from 1.7
GPa to 2.5 GPa.
The effect of the production technology and the influence of possible
volume and surface defects (such as microcracks, cavities, and indents) on
mechanical properties of basalt fibers was addressed in Gur’ev et al. [61].
290 They observed that the presence of defects in longitudinal and radial direc-
tions does not entail a significant variation in fiber stiffness, while seems to
produce a substantial decrease of tensile strength.
In order to partially limit this drawback, some authors proposed to coat
externally the fibers by means of a particular sizing during their production.
295 In particular, Greco et al. [56] compared basalt fibers modified with silane
sizing with those unsized, proving that sized fibers revealed smooth surfaces

10
and enhanced mechanical performance, as well as highlighting a certain en-
hancement in adhesion properties when fibers were embedded in a matrix.
In detail, the adhesion properties of basalt fibers were shown to be higher
300 than those of sized glass fibers, and comparable to those of carbon fibers. In
all cases, it was highlighted that the adhesion properties can be significantly
improved by increasing the polarity of the polymer matrix [56].
Aiming to provide some quantitative indications in terms of tensile strength
and elastic stiffness, Table 1 summarises mean values collected from some ex-
305 perimental data available in literature [57, 61–63], by comparing basalt fibers
with carbon and glass fibers. Although present data refer to different pro-
duction and experimental set-ups, thereby preventing rigorous comparisons,
they clearly show that basalt fibers exhibit mechanical properties fully com-
parable with those of glass fibers and almost competitive with those of the
310 most expensive carbon ones. As a matter of fact, the tensile strength of
basalt fibers is higher than glass fibers, and it results comparable or lower
than the carbon fibers [62]. Moreover, although carbon fibers are character-
ized by the highest values of the elastic modulus, basalt fibers experience a
stiffness fully comparable with glass fibers. This occurrence is in agreement
315 with many other studies and experimental evidence available in literature
[13, 56, 64–66], herein not discussed for the sake of brevity.

2.4. Chemical properties


The prolonged exposure of basalt fibers to alkalis or acid components
can influence their mechanical performance generally leading to mass loss
320 and to strength reduction. As a result, the mechanical properties of com-
posite materials reinforced with basalt fibers can undergo to a degradation
process. Altough many studies strictly address the influence of the envi-
romental agents on the mechanical behavior of the overall fiber-reinforced
composite material, a fundamental step in this framework is represented by
325 the characterization of the chemical properties of the fibers [12].
The influence of the fiber treatment in alkaline (NaOH) or acid (HCl)
solutions on the mechanical properties is addressed in some experimental
studies [12, 63, 67, 68]. Corresponding results revealed that the acid resis-
tance of basalt fibers is better than the alkaline one. In detail, Mingchao
330 et al. [68] observed that the fiber mass loss in HCl solution was about 8%,
resulting two times greater than the one experienced when basalt fibers were
immersed in a solution of NaOH with the same concentration (2 mol/L).

11
Table 1: Geometrical and mechanical features associated to some experimental data re-
garding basalt, glass and carbon fibers subjected to tensile tests. F : fiber type (B: basalt;
C: carbon; G: glass); d: diameter; L: length; Et : elastic tensile stiffness; fut : tensile
strength.

Ref. F d [µm] L [mm] Et [GPa] fut [MPa]


[62] C 8.2±0.4 15.0 183.0±29.0 2136.0±531.0
C 8.2±0.4 30.0 201.0±20.0 1526.0±643.0
C 8.2±0.4 40.0 195.0±12.0 1545.0±620.0
G 10.4±0.6 15.0 71.0±11.0 1641.0±673.0
G 10.4±0.6 30.0 69.0±14.0 1035.0±402.0
G 10.4±0.6 40.0 69.0±6.0 1202.0±571.0
B 12.3±1.1 15.0 66.0±5.0 1826.0±617.0
B 12.3±1.1 30.0 64.0±4.0 1586.0±433.0
B 12.3±1.1 40.0 67.0±4.0 2044.0±244.0
B 20.6±2.2 15.0 103.0±11.0 2277.0±392.0
B 20.6±2.2 30.0 82.0±19.0 1297.0±667.0
B 20.6±2.2 40.0 90.0±5.0 1295.0±510.0
[57] B 14.2±1.4 - 61.9±3.5 2016.0±434.0
B 12.7±1.5 - 62.0±3.6 1608.0±350.0
B 14.1±2.9 - 53.2±7.4 1811.0±331.0
G 16.8±1.6 - 57.0±3.0 1472.0±395.0
[61] B 10.1 - 91.9 2880.0
B 10.5 - 87.5 1760.0
B 9.5 - 86.1 3470.0
B 12.2 - 66.8 731.8
B 6.3 - 71.9 840.3
B 14.8 - 34.6 656.3
C 6.9 - 218.8 3779.8
C 6.9 - 227.6 3816.1
C 6.7 - 227.0 3929.9
[63] B 10.6 20.0 76.0 992.4

12
Moreover, a strength degradation of about 35% was registered in HCl solu-
tion and of about 20% in NaOH. The chemical stability of basalt fibers in
335 comparison to glass fibers was analysed by Wei et al. [12, 67] confirming that
for the basalt fibers the acid resistance is much better than the alkali resis-
tance, but for the glass fibers the acid resistance is similar to the alkali one.
In particular, the available evidence clearly proved that the acid resistance
of basalt fibers is greater than the one of glass fibers in terms of both mass
340 loss ratio and tensile strength maintenance ratio, respectively experienced as
about 10% and 40% for the basalt fibers and as about 30% and 20% for the
glass fibers. Conversely, the alkali resistance of basalt and glass fibers reve-
lead to be quite similar, as also confirmed by experimental results proposed
in [63]. All previously-referred studies highlighted the strong influence of the
345 exposure time on mechanical properties. For instance, referring to results
proposed in [63], when the exposure time in alkali solution passes from 7
days to 28 days, fiber volume reduction passes from about 20% to 70% and
strength reduction from about 50% to 80%, for both glass and basalt fibers.
Sim et al. [63] also compared the influence of alkali agents with respect to
350 carbon fibers. In detail, the volume loss and the tensile strength decrease
of basalt fibers and associated to different immersion periods (from 7 to 28
days) are strongly higher than those of carbon fibers (from about 10% to
20%, and from about 5% to 15%, respectively).
It is worth observing that many studies (e.g., [11, 40, 63, 69]) focus on
355 the chemical stability of basalt fibers to alkali agents, aiming to investigate
about feasibility of using basalt fibers as strengthening elements for struc-
tural concrete members. The immersion in alkaline solutions determines
reaction products due to the dissolving of the SiO2 -networks and a brittle
shell-like layer is generally found at the fiber surface [11, 12]. This layer is
360 usually characterized by a constant thickness around the fiber and, when
the immersion period increases, it tends to crumble away, resulting in visi-
ble volume reduction [11, 12, 63, 67]. As a matter of fact, similarly to the
case of glass fibers [11], the corrosion process in basalt fibers follows a layer-
by-layer mode, leading to a slow degradation of the original structure. In
365 order to counteract such an effect, Jung and Subramanian [40] suggested to
coat basalt fibers before immersing in an alkaline environment. In this way,
referring to zirconia-coated basalt fibers, the strength level was maintained
even after 90 days of alkali immersion, in contrast to uncoated fibers that
drastically lose their mechanical properties.
370 The degradation process due to immersion in cement solutions has been

13
specifically investigated in [11, 69]. The CaO absorption, the main alkaline
component of hydrating cements, leads to form a hardly soluble Ca-Si layer
at the fiber surface. Subsequently, voids with different sizes and depths tend
to appear on the surface but, in contrast to basalt fibers immersed in NaOH
375 solutions, the experienced reduction of the fiber diameter can be considered
as not significant [11, 69].
The influence of sunshine exposure on the tensile strength has been inves-
tigated in [63], highlighting that basalt fibers similarly to carbon and glass
ones exhibit a high stability with respect to this environmental condition. In
380 detail, strength values for basalt fibers reduce of about 13% after 20 years
of sunshine exposure (equivalent to 4000 hours of accelerated artificial expo-
sure).

2.5. Thermal properties


As regards thermal properties, basalt fibers show a very wide range of op-
385 erating temperatures and an excellent attitude to resist to high temperatures
(see Table 2), in comparison with the most common fiber types [13, 63, 70–
72]. The high thermal stability, in the range of from -200◦ C to 700◦ C, is
mainly related to the material characteristics of basalt rocks, which nucleate
at very high temperature [63, 72]. As a consequence, the softening tempera-
390 ture for basalt fibers is in the order of 960◦ C, resulting greater than that of
E-glass fibers of about 15% [10].

Table 2: Temperature withstand range (∆T ): comparison among basalt, E-glass, S-glass
and carbon fibers [13].

Fiber ∆T [◦ C]
basalt -260÷700
E-glass -50÷380
S-glass -50÷300
carbon -50÷700

This occurrence reflects in the preservation of good mechanical properties


also at high temperature. In detail, referring to experimental results proposed
in [70] (see Table 3), the average tensile strength of basalt filaments remains
395 almost constant up to 200◦ C, resulting practically unaffected by the exposure
time up to 1 hour. On the contrary, for temperature higher than 200◦ C,
the strength decrease becomes significant and improved by increasing the
exposure time. This evidence is fully in agreement with other experimental

14
studies available in literature [63, 70, 71]. In detail, results proposed by
400 Sim et al. [63] indicated that the strength reduction for basalt fibers when
temperature is greater than 200◦ C was less significant with respect to carbon
and glass fibers. Moreover, basalt fibers maintain an appreciable volumetric
integrity, also at high temperature (600◦ C - 1200◦ C), in comparison with
that of carbon and glass fibers [63]. In particular, thermal decomposition for
405 basalt fibers starts at about 200◦ C and, in the range 200◦ C - 800◦ C, it induces
a mass loss of about 0.74%, whereas for glass fibers thermal decomposition
begins at about 160◦ C and in the range 160◦ C - 850◦ C it induces a mass loss
of about 1.8% [71].

Table 3: Variation of the tensile strength fut of basalt-based tempered filament yarns for
different values of the exposure time t and temperature T [70].

t [min] 1 15 60
T [◦ C] fut [MPa] fut [MPa] fut [MPa]
20 1010 1010 1010
50 997 1050 1070
100 1030 9910 1010
200 986 1010 1090
300 893 743 424
400 743 701 112
500 254 348 94

2.6. Composite materials based on basalt fibers


410 Basalt fibers can be adopted as a reinforcement phase in composite ma-
terials, also employed for applications in civil engineering and mainly based
on the use of polymeric or cementitious matrices (see Fig. 2). Such mate-
rials can be realized with fibers characterized by different geometrical and
arrangement features. Continuous fibers, characterized by a nonwoven or wo-
415 ven (namely, fabrics) configurations, can be embedded into a polymeric resin
(e.g., epoxy, polyester or vinylester), leading to the so-called fiber-reinforced
polymers (FRP), to form fiber-reinforced plies. Single plies can be also com-
bined in different stacking sequences, obtaining FRP laminates. Further-
more, FRP concept can be employed to produce pultruted composite bars,
420 with continuous fibers mainly aligned along the bar axis, that can substitute
the steel bars in reinforced concrete structures. In other cases, continuous
fibers can be arranged in the form of open meshes, textiles, or fabrics and

15
embedded in cementitious matrices to obtain Fiber Reinforced Cementitious
Matrix (FRCM). Basalt-based composite materials can be also conceived by
425 considering short fibers. In particular, basalt short fibers can be randomly
dispersed in a cementitious matrix to form Fiber Reinforced Concrete (FRC)
or in a polymer matrix to form Short Fiber Reinforced Polymer (SFRP).
In the following, main available evidence and results addressing physico-
mechanical properties of the previously-introduced basalt-based composite
430 materials are summarized and discussed. Applications in civil engineering of
this kind of composite materials are reviewed, focusing on feasibility aspects
and perspective concepts.

3. Basalt-based FRP (BFRP)


Aiming to furnish useful comparative indications on the mechanical prop-
435 erties of basalt-based FRP with respect to glass-based (GFRP) and carbon-
based (CFRP) fiber reinforced polymers, main features related to fiber, poly-
meric resin, and fiber-matrix adhesion have to be accounted for.
Available experimental results highlight that strength and elastic modulus
of BFRP laminates based on continuous basalt fibers and realized through
440 different stacking sequences or weaving, are generally higher (or at least
comparable) than the GFRP’s, when the same fiber content and the same
polymeric resin are considered [62, 64, 65, 73, 74]. Reference is made to
some representative experimental studies [62, 64–66, 73, 74, 76], associated to
testing features summarized in Table 4, and corresponding to data obtained
445 from tensile, compressive and bending tests, synthetically collected in Tables
5 to 7.
Dorigato and Pegoretti [62] showed that laminates based on BFRP and
GFRP bidirectional fabrics immersed in epoxy resin have similar tensile prop-
erties when similar levels of fiber volume fraction were considered. Moreover,
450 an increase of about 7% in the basalt fiber volume fraction produced an in-
crease of the tensile strength in the order of 17% and of the elastic modulus
of about 20%. As a result, although they are characterized by smaller stiff-
ness, BFRP laminates with a fiber content in the order of 60% proved to
have a tensile strength lower than carbon-based laminates (with a similar
455 fiber volume fraction) of about 14% only (see Table 5).
Experimental evidence esthablished in [65, 74] and related to laminates
based on different fiber patterns, confirms the good BFRP mechanical prop-
erties in comparison with those of GFRP, both in tension and in compression

16
(a) (b)

(c) (d)

(e) (f)

(g) (h)

Figure 2: Basalt-based fiber-reinforced composites and some civil engineering applica-


tions. (a) BFRP sheet; (b) concrete elements reinforced via BFRP sheets; (c) BFRP bars
(indiamart.com); (d) BFRP stirrups (basalt-rebar.com); (e) Basalt-based open meshes
(build-on-prince.com); (f) Basalt-based Fiber Reinforced Cementitious Matrix (BFRCM)
(buildipedia.com); (g) short basalt fibers (basaltproductsgroup.com) and (h) Basalt Fiber
Reinforced Concrete (BFRC) [75].

17
(see Tables 5 and 6). In particular, tensile and compressive tests conducted by
460 Chairman and Kumaresh Babu [74] revealed that BFRP laminates are char-
acterized by tensile and compressive ultimate strength higher than GFRP
laminates of about 23% and 43%, respectively (see Tables 5 and 6). More-
over, as regards the ultimate strain obtained from monotonic tensile tests
[62, 64, 73], BFRPs revealed a larger elongation capacity with respect to
465 both carbon-based and glass-based laminates.
Results proposed in [65, 76] also confirm the better ultimate strength of
BFRP laminates than GFRP ones when bending tests are addressed (see
Table 7).
The influence of the matrix properties on the mechanical response of
470 BFRP laminates has been investigated in [66, 73]. Basalt reinforced epoxy
matrices (BE-FRP) exhibited more favourable mechanical properties, in terms
of strength and stiffness, compared to vinylester matrices (BV-FRP) [66]. It
is interesting to note that the experienced failure mode was different in such
two cases. The BE-FRP shows a linear stress-strain relationship until a sud-
475 den brittle tensile failure, whereas in BV-FRP a stress-strain plateau region
appeared before the failure, corresponding to a certain ductility degree. This
can be strictly related to the fact that BV-FRP specimens exhibited the ten-
dency of basalt fibers to separate from the vinylester resin, resulting in a
progressive adhesion reduction before the complete failure. Palmieri et al.
480 [73] compared the tensile performance of both basalt- and glass-FRP lam-
inates based on tixotropic epoxy and on a more viscous epoxy matrix (see
Tables 4 and 5). In both cases, the viscous epoxy matrix allowed to obtain
higher stiffness and a more larger elongation at failure, also thanks to its
wetting and impregnation properties. Epoxy, polyester and vinylester resins
485 belong to the class of thermosetting polymers but are not the only ones used
for composites reinforced with basalt fibers. It is worth mentioning also the
thermoplastic polymeric matrices (i.e., polypropylene and polyethylene) and
the metallic ones. Further details about their features and concerning me-
chanical properties of the corresponding basalt-based composite materials
490 can be found in [77].
The adhesion between fibers and matrix represents one of the most crit-
ical aspects that affects the mechanical response of the overall composite
materials. Several treatment strategies of the fiber surface, aiming to im-
prove the adhesion with the surrounding matrix and for ensuring an effective
495 load transfer mechanisms, can be found in literature specifically referring to
basalt fibers [78]. Some of the many strategies that have been investigated

18
are: enhancing surface roughness [79], increasing mechanical interlocking
with nanoparticles [80], colloidal silica and sol/gel techniques [78, 79], chem-
ical functionalization [38] and plasma treatments [81].
500 Bashtannik et al. [82] showed that the surface modification of basalt
fibers in acid and alkaline solutions tends to intensify the adhesion between
thermoplastic matrices and fibers, allowing to obtain a more developed fiber
surface. In particular, it was found that the acid treatment is more efficient
than the alkaline one, resulting in a significant improvement of the mechan-
505 ical properties of basalt-based reinforced composites. Such an occurrence
has been also confirmed by Manikandan et al. [83], highlighting that BFRP
laminates realized with superficially treated fibers revealed an appreciable in-
crease of tensile strength, inter-laminar shear strength and impact strength,
if compared with those realized via untreated fibers. In particular, since
510 the different mechanisms of acid and alkali attacks occurring on basalt and
glass fibers, treated BFRPs experienced a greater improvement of mechan-
ical properties than treated GFRPs, resulting in higher tensile and impact
strength when an acid treatment is considered, and in a higher shear strength
for alkali-treated laminates. Aiming to improve interlaminar shear strength
515 in BFRP laminates, one of the available strategies is based on the increase
of the fiber surface roughness by appliyng coatings made from pure epoxy
and SiO2 nanoparticles [80]. In this way, for a SiO2 weight concentration of
5%, Wei et al. [80] verified an increase of the interlaminar shear strength of
about 15%. Among the different surface modification strategies (e.g., reactive
520 sol-gel epoxy-silane methods, colloidal silica methods and combined forms)
aiming to exploit the chemical interactions between fiber and matrix with
the increase of the mechanical interlocking, the sol-gel epoxy-silane method
seems to furnish the most effective influence on the tensile strength of BFRP
composites [78].

525 3.1. BFRP based on short fibers


Although BFRP composite materials are generally conceived by consid-
ering long fibers, some attempts to use short basalt fibers for reinforce a
polymeric matrix can be found in literature. For instance, Zhang et al. [84]
studied wear and friction properties of short-basalt-fiber-reinforced polyimide
530 composites, revealing that a certain content of basalt fibers can improve the
tribological behavior of the composite, inducing a reduction in both fric-
tion coefficient and wear rate. Botev et al. [85] investigated the mechanical
properties of basalt fiber-reinforced polypropylene with different contents of

19
Table 4: Geometrical features associated to the available experimental data regarding FRP
laminates subjected to tensile, compressive or flexural tests. N: identification number for
each reference; F : type of fiber (B: basalt; C: carbon; G: glass; E-G: alkali-free glass; PBO:
polyparaphenylenl benzobisoxazole); t: laminate thickness; fw : fiber weight fraction; fv :
fiber volume fraction.
N-Ref. stacking sequence F resin t [mm] fw [%] fv [%]
1-[74] bidirectional fabric B epoxy 3.00 53.38 -
2-[74] bidirectional fabric E-G epoxy 3.00 57.76 -
1-[64] unidirectional fabric C epoxy - - 50.0
2-[64] unidirectional fabric PBO epoxy - - 50.0
3-[64] unidirectional fabric G epoxy - - 50.0
4-[64] bidirectional fabric B epoxy - - 50.0
1-[65] [0◦ /90◦ ]16 B epoxy 2.50 - 51.0
2-[65] [0◦ /90◦ ]8 G epoxy 2.50 - 46.0
3-[65] [0◦ /90◦ ]25 B epoxy 4.00 - 47.0
4-[65] [0◦ /90◦ ]10 G epoxy 4.00 - 40.0
5-[65] [0◦ /90◦ ]16 B epoxy 10.00 - 51.0
6-[65] [0◦ /90◦ ]10 G epoxy 10.00 - 46.0
1-[62] bidirectional fabric C epoxy 0.89 - 63.5
2-[62] bidirectional fabric G epoxy 0.72 - 56.3
3-[62] bidirectional fabric B epoxy 0.68 - 61.3
4-[62] bidirectional fabric B epoxy 1.01 - 54.1
1-[73] - B tixotropic epoxy 0.89 - 12.6
2-[73] - E-G tixotropic epoxy 0.89 - 12.6
3-[73] - B viscous epoxy 0.89 - 12.6
4-[73] - E-G viscous epoxy 0.89 - 12.6
1-[66] [0◦ /90◦ /+45◦ /-45◦ ]2s B vinylester 3.50 - 50.0
2-[66] [0◦ /90◦ /+45◦ /-45◦ ]2s B epoxy 3.50 - 50.0
1-[76] bidirectional fabric B vinylester 3.20 48 28.0
2-[76] bidirectional fabric E-G vinylester 3.20 48 28.0

20
Table 5: Mechanical features associated to the available experimental data regarding FRP
laminates subjected to tensile tests and corresponding to geometrical features in Table 4.
fut : ultimate tensile strength; Et : elastic tensile modulus; εu : ultimate strain.

N-Ref. fut [MPa] Et [GPa] εu [-]


1-[74] 330± 17 - -
2-[74] 250± 14 - -
1-[64] 258±6.12 24.2 1.74
2-[64] 250±5.86 26.6 1.6
3-[64] 178±8.39 87 2.45
4-[64] 58±2.49 91 2.56
1-[65] 506±20 24.4±1.0 -
2-[65] 580±11 15.3±0.5 -
1-[62] 728±67 65.1±0.53 1.1±0.2
2-[62] 461±25 25.2±0.21 2.3±0.2
3-[62] 600±36 29.6±0.27 2.6±0.2
4-[62] 502±26 26.5±0.19 2.4±0.2
1-[73] 138.51 19.48 1.186
2-[73] 124.91 14.06 1.225
3-[73] 274.10 25.80 2.080
4-[73] 218.35 18.95 1.820
1-[66] 317.3±8.5 17.9±0.78 -
2-[66] 409.2±9.2 20.34±0.74 -

Table 6: Mechanical features associated to the available experimental data regarding FRP
laminates subjected to compressive tests and corresponding to geometrical features in
Table 4. fuc : ultimate compressive strength; Ec : elastic compressive modulus.

N-Ref. fuc [MPa] Ec [GPa]


1-[74] 300.0±17.0 -
2-[74] 170.0±8.0 -
3-[65] 280.0±20.0 22.2±1.4
4-[65] 174.0±15.0 13.5±2.3
1-[66] 169.3±14.1 -
2-[66] 311.8±16.7 -

21
Table 7: Mechanical features associated to the available experimental data regarding FRP
laminates subjected to bending tests and corresponding to geometrical features in Table
4. fub : ultimate bending strength; Eb : elastic bending modulus.

N-Ref. fub [MPa] Eb [GPa]


5-[65] 504.0±23.0 22.7±1.1
6-[65] 224.0±11.0 14.6±1.4
1-[76] 236.0±7.0 13.5±1.4
2-[76] 195.0±8.0 18.7±1.2

short basalt fibers, highlighting the influence of a fiber treatment based on


535 PP-g-MA copolymer as a coupling agent. In detail, the incorporation of un-
treated short basalt fibers in a polypropylene matrix revealed to lead to the
deterioration of its tensile properties due to the lack of strong fiber/matrix
interfacial bonds. Whereas, by using the coupling agent, the matrix tend to
benefit of the fiber reinforcement in terms of both yielding tensile strength
540 and impact strength. This occurrence was also confirmed by the evidence
proposed in [86] where the fibers were treated with a reaction mixture of
maleic acid anhydride and sunflower oil. Generally, such a kind of treatments
did not significantly affect the elastic modulus, which tends to increase by
increasing the fiber content [85]. The correlation between the tensile strength
545 and both the fiber length and the fiber content has been estimated in [87].
These two latter parameters were proved to significantly affect the effective-
ness of the fiber/matrix interface bonding via the number of the fiber ends,
where stress concentrations may occur leading to possible onset of local fail-
ure mechanisms. As a matter of fact, results proposed by Amuthakkannan
550 et al. [87] allow to identify an optimal fiber weight content of about 68% and
an optimal fiber length of 10 mm, that maximizing strength features of the
short-basalt-fiber-reinforced polymer.

3.2. Hybrid FRP based on basalt fibers


With the aim to enhance possible non-optimal mechanical features of FRP
555 composites based on the use of a single fiber type, composite materials defined
by employing two or more different type of fibers embedded in a common
matrix can be generally conceived, namely referred to as hybrid FRP. In
this way, possible counteracting mechanical aspects (related for instance to
strength, stiffness and inelastic mechanisms) can be mediated by mixing
560 features associated to specific fiber types.

22
For instance, the flexural properties of carbon-basalt (C/B) hybrid com-
posites realized by considering different stacking sequences of CFRP and
BFRP laminae are discussed in [88, 89], resulting in a strong influence of fiber
arrangement on flexural strength, fatigue behavior and impact performance.
565 In all the considered cases, the hybridization results in an improvement of the
strength with respect to BFRP of about 73%. In particular, when the car-
bon layers are disposed on the compressive side of the laminate, the highest
flexural strength and modulus are reached. Moreover, the best performance
in terms of ductility was obtained when basalt layers were placed in both
570 tensile and compressive zones. As a matter of fact, although a reduction of
strength with respect to the case of FRP based on carbon fibers only was
observed (about 14%), a greater compliance was experienced than the CFRP
case, due to the controlled reduction in the overall stiffness related to the use
of basalt fibers.
575 The combined use of carbon and basalt fibers in such a material design
concept allows to obtain an enhancement also in the fatigue response in com-
parison to CFRP. This has been proved by Wu et al. [64] by showing that
the presence of both carbon fibers and basalt fibers, since the correspond-
ing stiffness mismatch, allows to counteract the propagation of progressive
580 fracture and delamination mechanisms. Moreover, unlike what commonly
happens for carbon-glass (C/G) hybrid composites, for C/B ones no delami-
nation occurs, probably due to the better adhesion between the rough basalt
fibers and the resin than the glass case [64]. Dorigato and Pegoretti [90] also
confirmed that the use of basalt fibers in C/B composites tends to hinder
585 the damage propagation and allows an improved impact energy absorption,
generally avoiding a catastrophic failure. This role is better performed by
inserting basalt fibers instead of glass fibers, due to the higher ductility index
values of C/B composites with respect to C/G ones [90].
Hybridizations defined by combining short basalt fibers with carbon, glass
590 or hemp fibers are compared in [86]. The experimental results there proposed
revealed that, with respect to basalt-based monocomposites, basalt-carbon
and primarily basalt-glass hybrid materials show a significant increase in
mechanical properties (in terms of both tensile and bending stiffness and
strength), whereas the addition of hemp fibers to basalt-based composites
595 does not allow an effective improvement [86]. The good performance of
basalt-glass (B/G) hybrid laminates with different stacking sequences is also
assessed in [91]. Several experimental evidences concerning hybrid compos-
ites made with basalt and natural fibers (such as flax, jute and hemp), show

23
lower (resp., higher) mechanical performance with respect to BFRP (resp.,
600 natural-fiber-based-FRP) [92, 93].
The feasibility of basalt-aramid (B/A) hybrid laminates realized via two
different configurations has been investigated by Sarasini et al. [94], consider-
ing different low-velocity impact conditions. Corresponding results indicated
that the hybridization of basalt-based laminates with aramid layers highly
605 improves the damage resistance capability with respect to BFRP. In particu-
lar, B/A laminates with an alternating sequence of basalt and aramid fabrics
were compared with laminates defined by basalt fabric layers at the core of
the composite and aramid fabrics as skin. The first ones exhibited better
energy absorption capability (also in comparison with aramid laminates), as
610 well as the most favourable degradation pattern (with smaller delaminations
between dissimilar layers) and the best combination of post-impact flexural
strength and damage tolerance. Accordingly, by properly designing the lay-
up sequence, such results indicated that B/A hybrids may exhibit similar
features in terms of impact absorption and damage tolerance with respect to
615 monocomposite aramid laminates.
Finally, hybrid composites realized by adding two fiber types among flax,
hemp, and glass (F, H, and G respectively) to epoxy resins reinforced with
basalt fibers, are considered in [92]. The G/F/B hybrid configuration exhib-
ited better mechanical properties, with an increase of about 19% and 32% in
620 tensile strength and of about 22% and 5% in tensile modulus, with respect
to G/H/B and F/H/B ones, respectively.

3.3. Chemical properties


As already outlined, the service life of FRP composites can be reduced
by the exposure to corrosive agents. By referring to civil engineering ap-
625 plications, the exposure of FRP composites to a concrete-related alkaline
environment or to corrosive natural environment (this latter associated for
instance to rain, acid rain or sea water) can induce a significant degradation
of the corresponding mechanical performance. This occurrence is strictly
related to the corresponding degradation of fibers, matrix and/or interface
630 adhesion.
The degradation of epoxy-based matrix reinforced with basalt fibers, af-
ter various periods of exposure in corrosive solutions, has been investigated
in [95]. Results show that BFRP exposed to water, salt, and alkaline so-
lutions, has a corrosion resistance similar to that of GFRP, whereas BFRP
635 resistance in acid solution is less than the GFRP’s. Mingchao et al. [68]

24
analysed the degradation of BFRP induced by eight different kinds of chem-
ical solutions (30% vitriol, 5% hydrochloric acid, 5% nitric acid, 10% sodium
hydroxide, saturated sodium carbonate solution, 10% ammonia, acetone and
distilled water) confirming the better chemical durability and corrosion be-
640 havior in the case of alkaline solutions. Moreover, in agreement also with
results proposed by Wu et al. [95], experimental tests discussed in [68] in-
dicated that corrosion effects in BFRP generally tend to affect the ultimate
strength rather than the elastic modulus. Nevertheless, basalt-FRP com-
posites exhibit greater improvement in their tensile strength retention un-
645 der different types of corrosion than isolated basalt fibers, mainly since the
protection role of the matrix [95]. The latter tends to endure or to delay
the penetration of corrosive agent into the composite, preventing the severe
damage of basalt fibers. Such an effect is limited by the degradation of the
fiber-matrix interface that generally, by referring to surface-treated basalt
650 fibers, is reduced in comparison to the case of glass fibers [83].
The degradation behavior of an epoxy resin reinforced with basalt fibers
previously treated with artificial seawater at a salt weight concentration of
6% has been investigated in [96], and compared with the analogous case con-
cerning GFRP. Such a treatment induces specific surface chemical reactions
655 leading to the presence of a thin layer of reaction products at the fiber-matrix
interface. This layer was responsible for a marked adherence loss, resulting
for both basalt- and glass-FRP in a significant and progressive decrease of the
tensile strength when the treating time was increased. Nevertheless, BFRP
exhibits a better behavior than GFRP, resulting in the fact that the mechan-
660 ical degradation level observed after a 30-day immersion time for BFRP was
experienced for GFRP after 10 days only. This effect can be ascribed to the
BFRP’s better interface adhesion properties, thanks to the lower presence of
Fe2+ elements that lead to a higher stability in seawater [96].

3.4. Thermal properties


665 Starting from the good thermal resistance of isolated basalt fibers (see
Section 2.5), it is easy to become convinced that the BFRP composites show
high fire resistance and good performance under strong environmental ther-
mal variations.
The heat resistance of BFRPs, made with unidirectional basalt fibers
670 and a polysiloxane matrix, subjected to partial pyrolysis of their precursors
(namely, exposed to a temperature of 650◦ C or 750◦ C in nitrogen), has been
investigated in [97] by addressing the corresponding degradation of flexural

25
strength. The cases of composites based on basalt fibers and obtained with
and without lubrication have been considered. Both removal of lubrication
675 and increasing of the treatment temperature were found to diminish the
flexural strength. Comparing results obtained for specimens treated at 650◦ C
and 750◦ C, the pyrolysis mainly affects matrix-governed elastic properties
(e.g., shear modulus) rather than those dominated by the fibers (e.g., along-
the-fiber elastic modulus). Moreover, a transition of the flexural failure mode
680 towards a brittle collapse was experienced when specimens were exposed to
hot air after the pyrolysis process, due to the occurrence of microcracks in
the matrix [97].
Landucci et al. [98] investigated the suitability of BFRP as passive fire
protection, simulating jet fires at laboratory scale acting on panels comprising
685 basalt fabrics embedded in a epoxy resin. Results show that the BFRP panels
exhibit very localized high temperature regions concentrated only where the
flame impacts the panel. The extension of high temperature zones is reduced
in comparison to the GFRP case, as a straight result of the lower thermal
conductivity (about 0.035 W/mK for basalt fibers, about 0.05 W/mK for
690 glass fibers). This aspect, in combination with a higher softening temperature
of basalt than glass, leads to a reduction also of the region affected by weight
loss, confirming the feasibility of fire-protection devices based on BFRP.
The thermal stability of BFRP and basalt-carbon hybrid FRP under the
effect of freeze-thaw cycles has been verified in [99]. Also in this case, BFRPs
695 exhibit better performance compared with GFRPs and CFRPs, being char-
acterized by higher retention rates in mechanical properties. In comparison
with CFRP and GFRP composites, BFRP reveal a better freeze-thaw re-
sistance with a loss of only 5% of the strain capacity after 300 cycles [100].
Moreover, B/C hybrid FRP sheets show a better freeze-thaw cycling resistant
700 capacity than monocomposite FRP, resulting in a no significant reduction of
ultimate strength and elastic modulus. This evidence is fully confirmed by
recent results proposed in [101]

3.5. Applications
In civil engineering field, BFRP in the form of laminates or sheets can be
705 mainly adopted for strengthening and retrofitting of existing structures based
on reinforced concrete (RC) and masonry. Figure 3 reports some exemplary
applications of FRP in civil engineering.

26
(a) (b) (c)

(d) (e) (f)

Figure 3: Exemplary applications of FRP laminates and sheets in civil engineering.


(a) Bridge strengthening (equipmentworld.com), (b) column confinement (theconstruc-
tor.org), (c) upgrading of existing structures (homesecurity.press), (d) flexural strength-
ening of masonry wall (seblog.strongtie.com), (e) strengthening of masonry arches [117],
(f) pile protection (pilebuck.com).

BFRP in reinforced concrete structures


Following well-established strategies introduced for strenghtening and
710 retrofitting of reinforced-concrete structures via fiber reinforced laminates
[102–112], BFRP can be successfully employed for improving the flexural
behavior of beam elements, for enhancing the confinement performance in
column elements, as well as for the reinforcement of the column-beam joints.
The effectiveness of BFRP reinforcements applied in the tension zone
715 of concrete elements undergoing flexural behavior has been investigated in
[63, 113, 114]. The experimental program carried out by Sim et al. [63]
focuses on the influence of the BFRP dimensions on the strengthening effect,
revealing that yielding, ultimate strength, and post-yielding stiffness increase
as the number of the composite layers increase, in agreement also with results
720 proposed in [114]. In particular, passing from BFRP reinforcements based on
one sheet (0.5 mm thick and disposed on the 80% of the total beam length) to

27
three sheets, the increment in the yielding load (respectively, in the ultimate
strength) passes from 15% to 27% (respectively, from 0 to 29%), with a
transition of the failure mode from the flexural-RC collapse to a BFRP-
725 RC debonding mode [63]. The variation of BFRP-RC bonded length, when
it results greater than 80% of the total beam lenght, generally does not
significantly affect the load-deflection behavior [63].
Gribniak et al. [113] consider the combined use of externally-bonded
BFRP sheets and BFRP bars as internal reinforcement in the concrete ten-
730 sion zone, resulting in an average value of the stiffness increment of about
30% (with a minimum value of about 13%). In all the considered specimens,
failure mechanisms associated to bending tests do not involve BFRP-RC
debonding for cases based on both two and three BFRP sheets [113].
The mechanical performance of BFRP flexural reinforcements generally
735 can be considered slightly lower in comparison with those achieved via GFRP
and CFRP in terms of percentage increments in yielding and ultimate strengths.
For instance, an increment in the ultimate flexural load in the order of 17%
to 50% (depending on the steel amount) was experienced in [102] for RC
beams reinforced with CFRP. Moreover, the strengthening effect of glass
740 fiber sheets was indicated in [63] to be better by 19% (one sheet) and 33%
(two sheets) than the case associated to basalt fibers. Nevertheless, for flex-
ural reinforcements based on both CFRP and GFRP, catastrophic strength
reduction occurred at relatively small deflection, resulting in lower ductility
levels than control beams [63, 102, 103]. On the contrary, better ductility
745 levels can be achieved in the case of BFRP flexural reinforcements, since the
reduced stiffness of basalt fibers.
Such an evidence has been also outlined in [22], where a non-linear ana-
lytical procedure has been proposed and compared with results reported in
[63]. In particular, referring to flexural reinforcements almost comparable in
750 thickness, discussed results highlight that beams reinforced with BFRP have
a bending ultimate load higher of about 20% than the case of GFRP, and
lower of about 40% than the case of CFRP, with an increase of ductility levels
in the order of 18% with respect to GFRP (respectively, 28% for CFRP).
As it is well known, the effectiveness of FRP-based flexural reinforce-
755 ments is strongly affected by the FRP-concrete bond performance. In this
framework, and specifically referring to the case of BFRP-concrete interface,
debonding tests have been carried out in [21, 25]. Results and comparisons
provided by Diab and Farghal [25] confirmed that the use of a flexible ad-
hesive permits to achieve higher effective bond length and bond capacity of

28
760 the FRP-concrete interface. Moreover, although CFRP-concrete debonding
occurs at loading levels greater (about double) than the case of BFRP, the
fracture energy remains almost comparable in both cases. Double-shear tests
proposed in [21] highlighted a mode-II debonding failure mechanism of the
BFRP sheet from the concrete support, with the debonding load increasing
765 with both the reinforcement width ratio and thickness. Moreover, the active
transfer length at the debonding failure state tends to reduce by increasing
the width ratio, resulting practically independent on the BFRP thickness and
at the most equal to 200 mm. Aiming to furnish useful design indications, a
great effort has been recently paid for describing in an analytical way bond
770 performance, in terms of bond strength and effective bond length [21, 19, 25],
highlighting the need of different design approaches when different FRP types
are considered.
The BFRP strengthening effect for RC elements undergoing a bending
load can be reduced in freeze-thaw environment, since the occurrence of
775 different thermal expansion-contraction of components. As a result, cyclic
stresses arises at the interface, changing the failure mode from debonding in
the concrete layer to debonding in the adhesive layer [100]. This phenomenon
is emphasised by coupling effects induced from sustained load that tends to
promote the degradation of bond-slip relationship within the bond length.
780 Freeze-thaw cycling not only reduced the ultimate load, but also influenced
the ultimate slip and the stiffness before initial debonding. However, the
improvement of the adhesive durability tends to produce a better durability
of the BFRP–concrete bond capacity in a freeze-thaw environment [100].
As regards the FRP-based confinement, the compressive behavior of con-
785 crete cylinders fully and partially wrapped with BFRP sheets has been stud-
ied in [115], considering both monotonic and cyclic tests, and in comparison
with the use of CFRP. The effectiveness of BFRP-based confinement resulted
surely reduced in comparison with CFRP. Nevertheless, since the lower en-
vironmental cost (in the ratio 9:1 compared to carbon fibers) and very high
790 temperature resistance, confinement based on BFRP can be considered as a
really promising alternative. Confinement performance strongly depends on
the number of BFRP layers, resulting in a significant increase of the ultimate
strain (double for one sheet and three times for three sheets than unconfined
cylinders), of bearing compressive load (about 10%-12%), and ductility. Such
795 an evidence has been also confirmed by experimental tests provided in [120].
Campione et al. [115] highlighted also that a confinement based on full wrap-
ping configuration revealed better than a discrete wrapping scheme, allowing

29
for an effective lateral pressure about three times greater. It is worth observ-
ing that, since the different stiffness, specimens confined with BFRP wraps
800 revealed a softening behavior, whereas the use of CFRP induced a mono-
tonic stress-strain response for the confined cylinders [115]. As regards cyclic
tests, the specimens wrapped with BFRP exhibited a very similar envelope
curve with respect to the stress–strain curve obtained from the monotonic re-
sponse. The authors have been also compared the experimental results with
805 those provided by analytical models available in literature and performed for
carbon, aramid or glass fibers, confirming a good correspondence between
experimental and analytical data [115].
Ouyang et al. [116] compared the performance under seismic cyclic load-
ing of unretrofitted square RC columns with those confined with BFRP and
810 CFRP jackets, also considering the hybrid B/C case. Failure of unretrofitted
columns was experienced as brittle since the lack of appropriate transverse
reinforcement. On the contrary, the confinement by FRP sheets significantly
reduces the spalling of the cover concrete in retrofitted columns. In this case,
a significant increase in ductility (about 72%-112%), in energy dissipation
815 and damping behavior, as well as in stiffness retention have been yielded.
The effectiveness of BFRP wrapping combined with BFRP bars as inter-
nal reinforcement of square columns has been investigated with reference to
steel-reinforced concrete columns, achieving also in this case a successfully
ductile-recoverable performance [118].
820 Further experimental studies on confinement with hybrid FRP composites
are reported in scientific literature [119, 120]. De Luca at al. [119] consid-
ered full-scale square and rectangular RC columns confined with GFRP and
hybrid glass/basalt FRP under axial loads. The hybrid laminate revealed a
comparable performance than the case of GFRP monocomposite. Further-
825 more, hybrid carbon/basalt FRP wrapping for concrete cylinder columns has
been evaluated by Long and Zhu [120]. They proved that, due to a conve-
nient stiffness distribution, the hybrid configuration based on CFRP as inner
layers and BFRP as outward layers leads to better strength and ductility
performance than the opposite case.
830 The flexural strengthening of beams and the confinement of columns could
not be sufficient for achieving an effective strengthtening and retrofitting of
concrete structures, especially in terms of seismic performance. In this con-
text, the column-beam joints highly affect earthquake-induced failure mech-
anisms in frame structure, as outlined in [121]. In the previously-cited refer-
835 ence, the use of BFRP reinforcements in column-beam joints has been proved

30
to induce an increase in the ultimate load (of more than 11%), allowing also
to obtain an increase in ductility levels (about 70%). As a result, the use of
BFRP reinforcements for joints damaged by cyclic loads leads to a signifi-
cant improvement of the corresponding seismic performance. In particular,
840 an efficient control of the bending damage at the column tip, shearing failure
of the core area and anchor damage of the joint can be achieved [121]. This
evidence is fully in agreement also with results discussed in [122, 123].
In addition to considerations discussed in Sections 3.3 and 3.4, long-term
durability issues of BFRP laminates and sheets employed for strengthening
845 concrete structures have to be addressed by strictly accounting for the bond
capacity at the BFRP-concrete interface. In this context, experimental re-
sults provided in [100] suggest that the application of a toughening-modified
resin may induce an enhanced durability of BFRP–concrete interfaces in
harsh environments (e.g., hydrothermal, freeze-thaw cycling, and salt-water
850 immersion). Moreover, test results showed that the bond capacity of the
BFRP-concrete interface decreases with increasing freeze-thaw cycles. As an
effect of an extra degradation of the bond-slip response, possibly associated
to the coupling betweeen freeze-thaw cycling and sustained load, authors also
observed a change in the failure mode from debonding in the concrete layer
855 into debonding in the adhesive layer. Further durability indications, useful
in the framework of BFRP adopted for concrete structures, can be found in
[124]. In detail, Alaimo et al. [124] addressed durability in BFRP laminates
undergoing cycles of artificial ageing in climatic chamber and outdoor expo-
sure. The results showed a good effectiveness over time of BFRP laminates,
860 associated with an initial increase of the mechanical performances after the
first step of the accelerated ageing, and also related to a slight influence of
UV radiation on resin behavior.

BFRP in masonry structures


Although the FRP-based strengthening of masonry structures, such as
865 arches, walls, vaulted systems, columns, have been widely focused in recent
literature [117, 125–127], very few examples can be found as specifically re-
ferred to the case of basalt fiber-reinforced composites.
Di Ludovico et al. [128] assessed the confinement performance of FRP
wrappings of tuff and clay brick square columns with different fiber types,
870 namely carbon, glass and basalt. Corresponding results show that the ef-
fectiveness of FRP confining systems is more significant on clay bricks, in-
dependently on the fiber type. This may be due to the limited transverse

31
dilatation at failure of tuff masonry, that tends to reduce the confinement
effect. Moreover, BFRP confining system was proved to lead to a gain in the
875 compressive strength in the order of about 57% and fully comparable with
that achieved via GFRP. Nevertheless, BFRP wrapping was more effective in
terms of the ultimate axial strain gain (413% for BFRP confinement against
259% for GFRP one) even if its mechanical external reinforcement ratio was
lower than GFRP [128]. In agreement with results previously discussed and
880 relevant to the confinement of concrete columns [115], also in the case of ma-
sonry columns, confinement effects (in terms of strength and ductility gains)
were maximized when a full wrapping configuration was considered. This
occurrence has been highlighted in [129] by considering different wrapping
configurations based on GFRP and BFRP for cylindrical columns made with
885 calcareous stones.
As regards masonry retrofitting, the in-plane mechanical behaviour of un-
reinforced masonry walls, strengthened or repaired after damage with BFRP
sheets, has been studied in [130]. In particular, referring to strengthen-
ing/repaired configurations based on horizontal and/or diagonal strips, re-
890 inforced walls exhibited a dominant flexural failure behavior. With respect
to the reference specimen, an effective improvement of the shear strength
of about 44% – 61% (resp., 40% - 50%), proportional to the BFRP rein-
forcement amount, and of the average cracking load of about 11% (resp.,
9% - 10%), have been experienced for strengthened specimens (resp., for
895 repaired walls). As regards the initial stiffness, no significant difference oc-
curred in strengthened specimens, whereas a certain increase was experienced
for repaired walls. An interesting result is that, when a mixed strengthen-
ing scheme with horizontal and diagonal strips was considered, even as the
severe debonding and buckling of diagonal BFRP strips occurred, the strips
900 still contributed in carrying the loads along with the masonry wall, until the
anchorage system continued to work [130].
As regards durability issues in BFRP retrofitted masonry elements, ex-
perimental and numerical studies have been proposed in [131, 132]. Ghiassi
et al. [131] analyzed the degradation mechanisms due to environmental ex-
905 posure of different FRP-masonry systems. In general, they showed that both
global and local responses of BFRP-masonry systems may be affected by
a coupling effect induced by temperature cycles and relative humidity. As
matter of fact, the combined influence of temperature and humidity tends to
induce a global reduction in bearing capacity and stiffness, and locally mod-
910 ifies the stress-strain patterns, leading to an increase in the effective bond

32
length.
In the context of the use of BFRP for masonry structures, a possible
application field is represented by the historic masonries. Nevertheless, even
if FRP composites can be generally considered for strengthening and con-
915 servation of historic masonry [133, 134], some limitations due to the use of
organic resins have been recently outlined [129, 135]. In particular, when
a poor support is found (i.e., as in the case of heritage buildings), early
FRP debonding may dramatically reduce the effectiveness of the strength-
ening system. Moreover, the use of polymeric resins obstacles the moisture
920 evaporation cycle through the masonry core, inducing a further issue for the
masonry conservation related to low levels of FRP transpirability [129]. In
this framework, basalt-based FRPs do not surely represent to date an ex-
ception. Accordingly, since the lack of a full mechanical and physical FRP
compatibility with the masonry support, the use of Fiber Reinforced Cemen-
925 titious Matrices is generally suggested [129, 135].

4. BFRP bars
As previously outlined, BFRPs can be also produced in the form of bars,
that are mainly used as internal reinforcement of structural elements. In
this case, unidirectional continuous fibers are employed with a fiber content
930 characterized by a volume fraction in the order of 80%, thereby resulting
much higher than that usually adopted for FRP laminates. The external
surface of BFRP bars is generally shaped in an helical braid and in some
cases it is covered with a sand coat (see Figure 4) in order to improve the
adhesion between the structural element that has to be reinforced and the
935 bars themselves [136, 137].
For civil engineering applications, basalt bars can be considered as an
effective alternative to the ordinary steel bars and to the most common FRP
bars, such as carbon-based and glass-based ones [138]. As a matter of fact,
basalt rebars exhibit an appreciable resistance in aggressive environments, a
940 density in the order of about only one-third that of steel, a tensile strength
about two to three times that of steel, and a thermal expansion coefficient
very close to that of concrete [139].
In order to furnish a quantitative outline of mechanical performance of
BFRP bars, and referring to the available literature [136, 140–143], Table
945 8 summarizes average experimental measures obtained from tensile tests in
comparison with those experienced on steel and other FRP bars.

33
(a) (b)

Figure 4: BFRP bars: (a) helical braid configuration (marketglobalnews.com); (b) sand
coating (sefindia.org).

Some experimental characterizations of BFRP rods have been carried


out in [26, 27, 136, 144]. As a result, BFRP rods reveal good values of
tensile strength and deformation levels, better than those experienced in
950 similar testing conditions for GFRP rods. Moreover, similarly to the case of
GFRP [27, 136], the stress-strain relatioship measured for BFRP bars with
diameter in the order of 4-12 mm generally exhibited a linear trend until a
sudden brittle failure mechanism. This occurrence has been also confirmed
by Serbescu et al. [26], highlighting that brittle failure mechanism may be
955 associated to a significant splintering in the central region of the bar when
diameter in the order of 6-8 mm is considered, such an effect tending to
disappear for diameter values of about 3 mm.
As regards rebars for concrete structural elements, the fundamental is-
sue of the bond behavior was focused for the BFRP case by Elrefai et al.
960 [146]. In detail, they performed pullout tests on concrete cylinders rein-
forced with FRP rebars, addressing both BFRP and GFRP for comparison
purposes. Corresponding results indicated that the failure mechanism was
mainly driven by the shear strength along the interface between the grained
layer and the subsequent core layers of the bar, revealing that BFRP bars
965 may be characterized by an average bond strength of about 75% of that of
the GFRP ones [146]. A further contribution on this topic was provided by
High et al. [27], that investigated the bond strength of BFRP bars embedded
in concrete elements when different surface deformations (ribbed and dented)
are considered. They concluded that bar surface deformation does not sig-
970 nificantly affect the bond strength, experiencing also that the full strength
of the BFRP bars is developed with a bonded length 32 times of the bar

34
Table 8: Mechanical features associated to some experimental data regarding steel, CFRP,
GFRP, AFRP and BFRP bars subjected to tensile test. fut : tensile strength; Et elastic
tensile modulus; εu : ultimate strain.

Ref. fut [MPa] Et [GPa] εu [%]


steel [140] 482-689 200 6-12
GFRP [140] 482-1585 35-51 1.2-3.1
GFRP [141] 770 37 2.1
GFRP [143] 824.5 50 1.78
GFRP [143] 760 40 1.85
GFRP [142] 918 35.2 2.69
CFRP [140] 600-3688 103-579 0.5-1.9
CFRP [141] 2250 147 1.5
AFRP [141] 1724-2537 41-125 1.9-4.4
BFRP [142] 707 24.8 3.03
BFRP [136] 1282 27.91 4.6

diameter.

4.1. Applications
Main applications in civil engineering involving BFRP bars are briefly
975 described in the following, highlighting some relevant experimental findings.
BFRP bars can be used as reinforcement in the tension zone of concrete
beams, as an alternative to steel rebars. The flexural behavior of concrete
beams reinforced with BFRP bars (BFRP-RC beams) has been assessed in
[144] via four point bending tests, in comparison with conventional steel-
980 based RC beams. In the case of BFRP reinforcement, concrete failure did
not occur in a sudden way and it was not accompanied by bar slipping, as in-
stead it may generally occur in steel-based RC beams. Due to the low tensile
stiffness of BFRP bars in comparison to steel ones (see Table 8), BFRP-RC
beams exhibited greater deflections and wider concrete cracks (about four
985 times) than steel-RC beams [144]. Such an evidence is fully in agreement
with results proposed in [147], where quantitative indications on the rela-
tionship between the load carrying capacity and the reinforcement area were
also provided. In detail, tested specimens exhibited a ductile behavior char-
acterized by a failure load increasing with the reinforcement area. Tomlinson
990 and Fam [148] considered BFRP bars not only as longitudinal reinforcement

35
but also as shear reinforcement in the form of stirrups. They observed differ-
ent failure modes: beams without stirrups failed in shear, beams with BFRP
stirrups failed in shear exhibiting the stirrup rupture, whereas beams with
steel stirrups failed in flexure. In all the considered cases both the ultimate
995 and the service loads increased by increasing the flexural reinforcement ratio.
BFRP bars can be also employed in combination with high-performance
cementitious composites (i.e., engineered cementitious composites - ECC),
generally characterized by pseudo-strain-hardening behavior and excellent
crack control [149, 150], for improving the mechanical performance of con-
1000 ventional concrete beams. Yuan et al. [137] considered the flexural behavior
of ECC and concrete/ECC beams reinforced with BFRP bars, by referring to
ECC comprising fly ash and polyvinyl alcohol fibers. As a result, the addition
of ECC in BFRP-reinforced concrete beams allowed to achieve an effective
enhancement in load-carrying and deformation capacities, shear resistance,
1005 and ductility (in terms of energy dissipation ratio).
BFRP bars can be also employed as alternative connecting rods in timber
structures by allowing an improvement in terms of load transfer and stiffness-
weight ratio, in comparison with conventional materials (e.g., steel) [151,
152]. In this context, Yeboah et al. [151] considered the pull-out behavior of
1010 bonded-in BFRP rods loaded perpendicular to the grain of glue-laminated
timber (generally referred as glulam) members. These latter are engineered
wood elements that can be successfully employed for trusses, purlins, floor
beams, cantilevers and other structural elements. Tested BFRP-glulam spec-
imens failed in shear and in timber splitting, with a failure mechanism gener-
1015 ally starting at the interface between the timber and the adhesive layer used
for bonding the BFRP rods. It was found that the samples exhibited double
branch behaviour consisting of ascending (brittle) and softening (pseudo-
ductile) behaviour for the pre- and post-peak bond, respectively. Moreover,
the pull-out capacity increased almost linearly with the bond length up to
1020 the maximum level which occurred at a bond length of 15 times the hole
diameter, and it did not increase beyond this bond length value [151].
With the aim to combine the advantages of elasto-plastic steel and lin-
early elastic BFRP, steel-fiber-reinforced polymer composite bars (SFCB)
have been conceived. They consist in a ribbed steel bar as inner core on
1025 which bundles of continuous fibers are wrapped, allowing to achieve, under
a monotonic load, a stable stiffness even after the yielding of the inner steel
bar [153]. Sun et al. [154] considered SFCBs as the reinforcement of concrete
columns, and they conducted lateral cyclic loading tests to compare steel-

36
BFRP bars with steel-CFRP ones. Results show that columns reinforced
1030 with basalt-based SFCBs have an ultimate elongation two times higher than
those reinforced with steel-CFRP bars, and thereby a better seismic perfor-
mance.
As regard durability issues, the performance of BFRP bars in aggressive
environments has been investigated in [26, 145, 155, 156].
1035 Serbescu et al. [26] carried out a number of tensile tests after condi-
tioning treatments, aiming to reproduce different and coupled environmental
effects. In detail, BFRP bars were treated by means of different degradation
processes, combining three different pH values (7, 9, and 13), temperature
levels (20, 40, and 60◦ C) and different exposure time periods (100, 200, 1000,
1040 and 5000 hours). Proposed results showed that, among these parameters,
the temperature induces the most significant effects, resulting in the domi-
nant cause of the degradation process acceleration. In any case, despite the
strength deterioration, an average increase of the elastic modulus of about
6.5%, probably due to the post-curing of the resin matrix, was found. More-
1045 over, Serbescu et al. [26], starting from such an experimental evidence, pro-
posed a comprehensive strength degradation model to predict the long-term
durability of BFRP bars in real-life environments.
The resistance to alkaline corrosion of BFRP bars was addressed by Wu
et al. [145]. In particular, they showed that, after 9 weeks in 55◦ C alkaline
1050 solution, BFRP bars maintain higher tensile strength (more than 60% of the
initial value) and strength retention over time with respect to GFRP bars, the
highest residual strength being experienced for the BFRP bars characterized
by the smallest tested diameter (6 mm).
Quagliarini et al. [155] underlined the importance of a proper resin choice
1055 to improve the BFRP bars’ durability in aggressive environments. In par-
ticular, experimental results showed that vinylester resin does not seem to
provide adequate protection to basalt fibers in alkaline environments. Con-
versely, BFRP bars with vinylester resin seem to work well if they are subject
to chloride attacks, (i.e., in marine environment or in presence of de-icing
1060 salts). Similar studies have been reported in [156], where BFRP bars with
epoxy resin exhibited lower degradation of their physical and mechanical
properties after conditioning in alkaline solutions, in comparison to BFRP
bars with vinylester resin.
Bond durability of sand-coated BFRP bars was investigated in [157]. Pull-
1065 out specimens were tested under direct tensile load after being exposed to
accelerated conditioning processes, by considering acid, saline, and alkaline

37
environments, as well as different exposure times (30, 60, and 90 days). The
BFRP bars showed higher adhesion and bond strengths to concrete than the
ribbed GFRP bars, irrespective on the fiber type and the exposure condi-
1070 tion. All specimens failed in pullout mode by inter-laminar shear between
the layers of the bar, exhibiting a conditioning-induced reduction of the bond
strength for the BFRP bars by 14-25%.
The continuous basalt fibers can be also processed via a classic textile
transformation to obtain ropes, that can be though as a special kind of BFRP
1075 bars without introducing any matrix. To the best of the authors knowledge,
basalt ropes are not applied in structural field yet but there are some studies
that suggest them as possible strengthening elements for historic masonry
structures [136, 158]. A tensile characterization of basalt ropes is proposed
in [136], where the obtained results are compared with those relevent to
1080 other fiber ropes commercially available. In particular, basalt ropes revealed
better levels of ultimate load than canvas and polyester, the latter generally
used in marine field, and lower performance with respect to steel cables.
Moreover, Monni and Quagliarini [158] carried out an experimental program
using basalt ropes 4 mm in diameter, addressing their use as continuous
1085 flexible elements for masonry walls stitching. Reticular and lozenge stitching
layouts have been applied on brick masonry samples, revealing that such a
technique can be able to improve masonry wall performance against out-of-
plane actions, as well as for enhancing the safety of bearing masonry structure
when subjected to vertical bending mechanism activation.
1090 Notwithstanding a lack of a systematic evidence about the durability of
basalt ropes, a first experimental study can be found in [155]. Results showed
that basalt ropes exhibit good levels of durability in terms of tensile strength
after water treatment. Under acid treatment, the fiber microstructure expe-
rienced a slight degradation, resulting in a certain reduction of the tensile
1095 strength (about 23%). When alkali environment was considered, basalt ropes
underwent a strongly damage, associated to a severe dissolution of the ma-
terial structure and inducing a significant reduction of the tensile strength
(about 90%).

5. Basalt Fiber Reinforced Cementitious Matrix (BFRCM)


1100 The BFRCMs are realized with open fabric meshes in which the basalt
rovings are assembled in two directions, generally orthogonal, by means of
several techniques such as weaving, knitting, tufting, or braiding. The ce-

38
mentitious material, besides being the matrix phase, serves as adhesive on
the surface on which the FRCM is applied [29, 30].
1105 The constitutive behavior of the BFRCMs, resulting from a uniaxial ten-
sile test, is generally characterized by three stages [29, 159, 160]. The stress-
strain relationship is linear until the first cracking of the cementitious matrix.
In the second stage, multiple crack patterns in the matrix occur, possibly ac-
companied by a textile-matrix slippage and without significant increase in
1110 stress. When the cracking pattern is stable, the BFRCM sample exhibits in
the third stage again an almost linear behavior until the specimen failure.
Before the rupture of basalt filaments, a certain degree of softening can occur
due to the debonding at the textile-matrix interface [159]. In some cases, the
constitutive behavior may be not trilinear and the second stage continues un-
1115 til the failure [31]. This may be due to a small fraction of the fibres breaking
during crack formation, as well as to the tensile performance of costituents
and to the matrix-textile bond features that can promote a boosting of the
tension stiffening effect [31].
Such an experimental evidence has been recently exploited to validate an-
1120 alytical models and computational approaches that can be useful to provide
criteria for reliably design and retrofit applications based on BFRCM [29–
32, 159, 161]. Nevertheless, the available experimental results do not claim
to provide effective design parameters, as they still refer to specific test-
ing conditions on a reduced number of specimens, enabling to derive robust
1125 statistics. Accordingly, in order to identify optimal patterns and features of
BFRCM constituents, both a larger number of experimental investigations
and probabilistic approaches would be needed [30].

5.1. Applications
The BFRCMs can be employed for strengthening masonries [162–164]
1130 and reinforced concrete structural elements such as walls, arches, columns
and beams [32, 161, 165, 166, 171]. In some cases, BFRCM is preferable
to BFRP because it offers increased resistance to high temperatures and
increased compatibility with the substrate on which it is applied [31, 166].
The effectiveness of the basalt textile-reinforced concrete as flexural strength-
1135 ening of RC beams has been investigated in [165, 166]. The monotonic
four-point bending tests, carried out by Gopinath et al. [165], revealed an
improvement of 84% in ductility for the reinforced beam, while the load
carrying capacity was not significantly affected in comparison with the not-
strengthened sample. Under low cycle fatigue load, a decrease of 20% in

39
1140 the failure load and of 27% in ductility with respect to the monotonic case
have been also reported. Experimental evidence proposed by Elsanadedy et
al. [166] indicated that, with respect to CFRP-based strengthening systems,
the BFRCM ones are generally more effective in terms of deflection ductility
than of flexural strength. Moreover, they highlighted that BFRCM realized
1145 with polymer-modified cementitious matrix ensures a better bond action be-
tween basalt textile and concrete substrate, improving ductility features and
flexural capacity in comparison with BFRCM realized with a conventional
cementitious matrix. In this framework, the positive influence on the shear
strength when polymer-modified mortar is employed has been also docu-
1150 mented in [161].
The shear strengthening of RC beams via basalt textile-reinforced con-
crete has been considered in [32, 161]. Escrig et al. [32] highlighted that
the effectiveness of the strengthening technique is strongly influenced by the
bonding behavior among the FRCM’s matrix, the reinforcing textile and/or
1155 the concrete substrate. In particular, the bonding performance of basalt- and
glass-based FRCMs allows to obtain a higher flexural toughness than that of
carbon-based ones, notwithstanding the lower elastic stiffness. Al-Salloum
et al. [161] considered the case of multiple BFRCM layers, by addressing
different textile orientation within the shear span. As a result, the shear
1160 resistance improves as the number of basalt-textile layers increased from two
to four per side and when the 45◦ /-45◦ basalt-textile configuration is applied
instead of the one characterized by the 0◦ /90◦ orientation.
Di Ludovico et al. [171] studied the advantages to employ basalt fibers
embedded in a cement-based matrix as a strengthening material for confine-
1165 ment of reinforced concrete cylinders. They considered basalt fibers preim-
pregnated with epoxy resin or latex, in order to improve the load transfer
through fibers, the epoxy resin resulting more effective especially in terms of
the ultimate compressive strength. As a matter of fact, confinement based
on BFRCM revealed promising, with average gains of 49% in compressive
1170 strength and of 91% in axial deformability when preimpregnation with epoxy
resin is considered. Moreover, such an approach induced failure modes char-
acterized by reduced brittle levels if compared with GFRP-wrapped members
[171].
Application of BFRCM for improving the out-of-plane behaviour of hol-
1175 low block masonry panels has been addressed in [162], by means of four-point
bending tests. Authors suggested that the use of inorganic mortars allows to
obtain good performance in terms of both sustainable load and displacement

40
levels. In comparison with the cases based on steel and glass reinforcements,
the basalt textile results very promising in terms of ultimate load, experienc-
1180 ing also the lowest displacements at the peak load. The feasibility of basalt-
based externally bonded grids for increasing the load-carrying and deforma-
tion capacity of stone-block masonry walls subjected to both out-of-plane
and in-plane cyclic loading conditions has been investigated in Papanicolaou
et al. [163]. In particular, proposed results clearly show that the use of the
1185 FRCM technique is more effective than the FRP one in terms of strength
and deformation capacity. The stress transfer mechanism between FRCM
and a masonry substrate has been also investigated in [164], by comparing
the effectiveness of steel and basalt fibers. Proposed results highlighted that,
depending on the fiber type, different failure modes of the FRCM-masonry
1190 joints can be obtained. In particular, FRCM-masonry joints were prone to
a delamination mechanism involving the external layer of the cementitious
matrix when high density basalt fibers were employed, whereas for steel and
low density basalt fibers the failure mechanism was mainly driven by the
fiber rupture. Moreover, although steel fibers allowed higher peak loads, the
1195 use of BFRCM enabled to obtain greater ductility levels.
Since better compatibility with the masonry substrate of BFRCM than
BFRP, as well as better reversibility and sustainability properties, basalt
fiber reinforced cementitious matrices could be conveniently applied in the
field of strengthening and preservation of historic masonries [31, 167–169].
1200 Nevertheless, to the best of the authors’ knowledge, specific studies are not
available in literature, as well as specific guidelines for application to existing
structures are still lacking [168, 169].
Due to the presence of a cementitious matrix, specific durability issues
need to be investigated for BFRCM, as strictly related to the occurrence of
1205 the fibrous reinforcement immersed in an alkaline environment. Micelli and
Aiello [135], aiming to contribute towards the definition of technical stan-
dards for FRCM, studied durability problems that may arise under service
conditions, specifically referring to the case of basalt and E-glass fibers. They
found that the degradation of basalt fibers immersed in a cement solution,
1210 with an exposure time of 180 days, induces a tensile strength drop of about
35%, significantly lower than the corresponding case of E-glass fibers (ten-
sile strength drop of about 76%). Reduced levels of tensile strength drop
were experienced when a lime-mortar bath was considered (31% for basalt
fibers and 47% for E-glass fibers). In order to improve durability perfor-
1215 mance of BFRCM a proper fiber coating can be employed, as implemented

41
in the post-earthquake reconstruction interventions on L’Aquila masonries
[170]. Nevertheless, discussed evidence can not be considered as conclusive
for the systematic assessment of durability issues on BFRCM. Accordingly,
further experimental programs should be specifically considered, by account-
1220 ing for different cementitious matrices, fiber treatments, geometrical features
and physico-chemical properties of the substrate on which the BFRCM is
applied.

6. Basalt Fiber Reinforced Concrete (BFRC)


The fiber reinforced concrete (FRC) consists in mixtures of cement, mor-
1225 tar or concrete containing short discrete fibers, uniformly dispersed and ran-
domly oriented. The fiber addition improves the structural integrity and the
fracture toughness, limits the loading-induced opening of cracks and their
propagation through bridging action in the brittle matrix, and contributes
to preserve concrete from corrosive agents [28, 172]. In this context, a key
1230 aspect is the definition of an optimal fiber volume fraction since, by employ-
ing a large amount of fibers, balled patterns can arise, also leading to voids
that can induce weakening effects [173].
Different fiber types have been employed for defining FRC, such as steel,
carbon, and glass. Nevertheless each of them shows some drawbacks. Steel
1235 fibers, despite the high elastic stiffness, increase considerably the structure
weight, as well as reduce processability and corrosion resistance; glass fibers
are very sensitive to chemical attacks in alkaline environments; carbon fibers
are very expensive; polymeric fibers exhibit low elastic stiffness and lead to
an ineffective adherence with the cementitious matrix. In this framework,
1240 the use of basalt fibers can be surely considered as promising, in agreement
with the feasibility indications provided in [172–175].
Jiang et al. [173] realized an extended mechanical characterization of
BFRC, in comparison with polypropylene (PP) FRC, through compressive,
flexural and splitting tensile tests. Corresponding results show that the addi-
1245 tion of chopped basalt fibers produces a significant increase in both flexural
and tensile strength, better than PP-FRC, although a slight reduction of
the compressive strength can occur. By increasing basalt fiber amount and
length, strength and toughness index of concrete increase, respectively, re-
sulting in an optimal condition identified by a fiber volume fraction equal to
1250 about 0.3% and by a fiber length of 22 mm [173]. A similar evidence was
experienced by Abdulhadi [174], specifically referring to Portland cement re-

42
inforced with basalt and polypropylene fibers. In particular, corresponding
results documented that, by considering 0.6% fiber volume fraction, the split-
ting tensile strength shows a gain of about 22.9% for BFRC and of about
1255 7.8% for PP-FRC. Moreover, short basalt fibers added to Portland cement
tend to decrease the drying shrinkage effect, exhibiting good bond level with
the hydrated cement matrix, especially at early ages (i.e., in the first 28 days)
[176].
Kabay [172] compared the mechanical performance of BFRC obtained
1260 from cement with two different water/cement (w/c) ratios (namely 0.45 and
0.60). In the case of higher (respectively, lower) w/c ratio, the flexural
strength increase of about 13% (resp., 9%), the fracture energy increase of
about 140% (resp., 126%) and the abrasive wear decreases in the order of 14-
18% (resp., 2-4%) with respect the unreinforced concrete. Neverthless, such
1265 an influence of the w/c ratio can be not considered completely understood
since, for instance, results proposed by High et al. [27] seem to indicate an
opposite trend. In particular, measuraments discussed in [27] highlight that
the reduction of the w/c ratio and the addition of fly ash and admixtures
produce an enhancement of the flexural strength in BFRC specimens. Kabay
1270 [172] also identifies a strict a correlation among flexural strength, abrasive
wear and void content in BFRC, suggesting an analytical description based
on the fitting of the corresponding experimental results.
The comparison of cementitious matrices reinforced with basalt, zirconia
doped basalt (ZrSi10 ), and AR-glass fibers is presented in [175], resulting in
1275 an increase of the flexural strength of about 8%, 18%, and 23%, respectively.
Moreover, the tensile strength of the reinforced concrete increases of about
43% considering both AR-glass and zirconia doped basalt fibers suggesting
this reinforcement type as a promising alternative for defining effective FRC.
The behaviour of different mixes of high performance concrete reinforced
1280 with basalt fibers (HP-BFRC) has been investigated in Ayub et al. [75],
highlighting that the splitting tensile strength and the strain capacity tend
to increase with the fiber volume content. On the contrary, the compressive
strength does not seem to be significantly affected by the addition of short
basalt fibers. However, the highest compressive strength - of about 24% with
1285 respect to plain concrete - was found for the HP-BFRC characterized by
10% of met kaolin and 3% of basalt fiber volume content, probably due to a
consequent improvement of the pore structure.
The glass aggregate concrete reinforced with chopped basalt fibers has
been studied in [177], referring to fiber volume content less than 0.5% and

43
1290 fine glass amount less than 60%. The analysis of the corresponding results
suggests that the optimum content of glass sand is 20%, resulting in an
enhancement of about 4% and 15% in compressive strength and splitting
tensile strength, respectively, after a drying stage of 28 days. Generally, the
use of basalt fibers leads to an enhancement in concrete strength with an
1295 optimal fiber content strictly related to the fine glass amount.
The passive fire protection of modified Portland concrete reinforced with
chopped strand basalt fibers has been compared in [98] to that of a BFRP
defined by a basalt woven fabric in epoxy resin. The analyzed BFRC showed
a better thermal behavior, being quite inert to flame impingement, and ex-
1300 hibiting a maximum wall temperature 20% lower than that obtained for
BFRP.
Recently, short basalt fibers have been also proposed as possible reinforce-
ment for hydraulic lime-based mortars, these latter being largely applied in
building and masonry restoration fields. In detail, Santarelli et al. [178]
1305 proved that the addition of basalt fibers, independently from the fiber type
(i.e., chopped or milled in the form of a loose powder), decreases the capillary
water absorption coefficient. On the other hand, the mechanical properties
seem to be degradated by the occurrence of basalt fibers and they strongly
depend on the fiber type, with higher values of compressive and flexural
1310 strength experienced in the case of milled fibers [178]. In this context, As-
prone et al. [28] confirmed the reduction of the compressive strength due to
the fiber presence. Fiber addition affects negatively also mortar workability
and it induces an increase in the porosity level that hinders the access to lime
hydration products. At the same time, the addition of basalt fibers produces
1315 a transition of the post-cracking flexural behavior from brittle-like to ductile-
like response, allowing to achieve higher ultimate displacements. Asprone et
al. [28] also addressed the dynamic behavior of a basalt-fiber-reinforced nat-
ural hydraulic mortar. It revealed a very strain-rate sensitive material, with
the tensile strength and the fracture energy increasing with the strain-rate
1320 more than other types of reinforcing fibers, such as steel and inorganic fibers.
Comparisons with theoretical estimates provided by CEB-FIP Model Code
[179] highlighted a theoretical overestimation in terms of tensile strength DIF
(dynamic increase factor) with respect to the experimental evidence [28].
Besides the most used Portland concrete, nowadays geopolymeric con-
1325 cretes are employed for construction of pavements, retaining walls, water
tanks, precast bridge decks [180]. Such a kind of material is an inorganic
polymeric concrete realized with waste materials, such as fly ash and ground

44
granulated blast furnace slag, and the corresponding manufacturing pro-
cess is characterized by lower CO2 emission than the ordinary Portland
1330 concrete (OPC). Dias et al. [181] compared, in terms of mechanical per-
formance, a basalt-fiber-reinforced geopolymeric concrete (BFRGC) and a
basalt-fiber-reinforced OPC. The BFRGCs provide higher compressive and
splitting tensile strength than OPCs. A linear load-deflection relationship
was obtained in both cases, but BFRGC experienced higher ultimate loads
1335 and displacements. The authors concluded that basalt fibers are more ef-
ficient in strengthening geopolymer concrete with respect to Portland one,
when a volume fiber content of 1% was adopted [181]. The dynamic compres-
sive strength, deformation and energy absorption capacity of BFRGC have
been also investigated in [182]. Corresponding results revealed a great strain-
1340 rate dependency, with an almost linear stress-strain rate relationship. As a
main evidence, the basalt fiber addition did not provide a significant increase
in dynamic compressive strength, but it produced consistent improvement in
deformation and energy absorption capacities [182].
As already highlighted for BFRCM, the exposure of basalt fibers to a
1345 concrete-related alkaline environment can adversely affect the overall me-
chanical properties of BFRC, thereby inducing a critical durability issue. In
order to overcome this drawback, Lipatov et al. [175] considered concrete re-
inforced with zirconia doped basalt fibers. In detail, the best alkali resistance
properties were found for basalt fibers doped with a ZrO2 weight concentra-
1350 tion of 5.7%. In this case, alkali treatment resulted in the formation of a
protective surface layer on fibers, characterized by the presence of insolu-
ble compounds of Zr4+ , Fe3+ and Mg2+ . Basalt fibers doped with zirconia
also revealed high performance in concrete, fully comparable to the case of
glass-based FRC.
1355 In this context, Borhan et al. [177] investigated the durability of glass
aggregate concrete reinforced with short basalt fibers. Combining high per-
centages of glass sand and short basalt fibers the resistance against the alkali
silica reaction increased, leading to an enhancement in the overall compres-
sive strength of the composite [177].

1360 7. Conclusions
A review of the state of the art concerning basalt fibers and their use as
reinforcement phase in composite materials for applications in civil engineer-
ing field has been provided.

45
Basalt fibers, obtained from melted natural basalt rocks and characterized
1365 by significantly cheaper production processes in comparison with other fibers
(e.g., carbon and glass), exhibit very appealing physico-chemical properties:
good strength and stiffness, high temperature resistance, long-term durabil-
ity, acid and alkali resistance, heat and sound insulation, as well as good
processability. These aspects, in addition to the high eco-sustainability lev-
1370 els of basalt fiber production and using, have induced a widespread diffusion
of basalt-based products.
In particular, basalt fibers characterized by different geometrical features
and configurations can be added to polymeric or cement-based matrices to de-
fine composite materials employed in novel structures as well as for strength-
1375 ening and retrofitting of existing structures. Basalt-based fiber-reinforced
composites can be roughly classified by considering four main groups: basalt-
based FRP (fiber reinforced polymer) in the form of laminates and sheets,
basalt-based FRP bars, basalt-based FRCM (fiber reinforced cementitious
matrix) and basalt-based FRC (fiber reinforced concrete). The promising
1380 potentialities of such kind of novel composite materials have recently at-
tracted the attention of the scientific community addressing cutting-edge
topics and applications in the field of civil engineering. In agreement with
previous classification, the most relevant results recently proposed and their
straight impact for structural purposes have been systematically traced and
1385 discussed, also referring to the critical issue of the durability performance.
The proposed picture contributes to show the effectiveness of composite
materials based on basalt fibers for civil engineering structures, and thereby
it brings light to novel and perspective applications. Nevertheless, the com-
petitive relationship, in terms of mechanical performance and cost effective,
1390 between basalt-based composites and other fiber reinforced materials (espe-
cially with respect to the glass-based ones) may lead to effective employing
in real-life cases only if further investigations will be provided. As a mat-
ter of fact, available design guidelines and technical codes revealed generally
not fully suitable and accurate when basalt-based materials are considered.
1395 In this context, the definition of a suitable and effective regulatory frame-
work, identifying technical indications and design rules when new or existing
structures are addressed, can be surely retained as an open issue.

46
Acknowledgements
This work was partially supported by the Italian Civil Protection Depart-
1400 ment [ReLUIS-DPC 2014-18, CUP: E82F17000870001].

References
[1] Vinson JR, Sierakowski RL. The behavior of structures composed of
composite materials. Dordrecht: Martinus Nijhoff 1986.

[2] Jamshaid H, Mishra R. A green material from rock: basalt fiber – a


1405 review. J Text I 2016;107(7):923–937.

[3] Brandt AM. Fibre reinforced cement-based (FRC) composites after


over 40 years of development in building and civil engineering. Compos
Struct 2008;86(1-3):3–9.

[4] Codispoti R, Oliveira DV, Olivito RS, Lourenco PB, Fangueiro R.


1410 Mechanical performance of natural fiber-reinforced composites for the
strengthening of masonry. Compos Part B: Eng 2015;77:74–83.

[5] Olivito RS, Codispoti R, Cevallos OA. Bond behavior of Flax-FRCM


and PBO-FRCM composites applied on clay bricks: experimental and
theoretical study. Compos Struct 2016;146:221–231.

1415 [6] Rao J, Bao L, Wang B, Fan M, Feo L. Plasma surface modification
and bonding enhancement for bamboo composites. Compos Part B:
Eng 2018;138:157–167.

[7] Ross A. Basalt fibers: alternative to glass? Compos Technol


2006;12(4).

1420 [8] Dhand V, Mittal G, Rhee KY, Park SJ, Hui D. A short review on basalt
fiber reinforced polymer composites. Compos Part B Eng 2015;73:166–
180.

[9] Czigàny T. Basalt fiber reinforced hybrid polymer composites. Mater


Sci Forum 2005;473:59–66.

1425 [10] Militký J, Kovacic V, Bajzı́k V. Mechanical properties of basalt fila-


ments. Fibres Text East Eur 2007;15(5-6):64–65.

47
[11] Scheffler C, Förster T, Mäder E, Heinrich G, Hempel S, Mechtcherine
V. Aging of alkali-resistant glass and basalt fibers in alkaline solutions:
Evaluation of the failure stress by Weibull distribution function. J Non-
1430 Cryst Solids 2009;355(52-54):2588–2595.

[12] Wei B, Cao H, Song S. Environmental resistance and mechanical per-


formance of basalt and glass fibers. Mat Sci Eng: A-Struct 2010;527(18-
19):4708–4715.

[13] Singha K. A short review on basalt fiber. Int J Text Sci 2012;1(4):19–
1435 28.

[14] Perevozchikova BV, Pisciotta A, Osovetsky BM, Menshikov EA, Kazy-


mov KP. Quality Evaluation of the Kuluevskaya Basalt Outcrop for the
Production of Mineral Fiber, Southern Urals, Russia. Energy Procedia
2014;59:309–314.

1440 [15] Fiore V, Di Bella G, Valenza A. Glass–basalt/epoxy hybrid composites


for marine applications. Mater Design 2011;32(4):2091–2099.

[16] Fangueiro R. Fibrous and composite materials for civil engineering ap-
plications. Elsevier 2011.

[17] Minghini F, Tullini N, Ascione F. Updating italian design guide CNR


1445 DT-205/2007 in view of recent research findings: requirements for pul-
truded FRP profiles. Am J Eng Appl Sci 2016, 9(3), 702-712.

[18] Valluzzi MR, Oliveira DV, Caratelli A, Castori G, Corradi M, De Fe-


lice G, Garbin E, Garcia D, Garmendia L, Grande E, Ianniruberto U,
Kwiecien A, Leone M, Lignola GP, Lourenço PB, Malena M, Micelli
1450 F, Panizza M, Papanicolaou CG, Prota A, Sacco E, Triantafillou TC,
Viskovic A, Zajac B, Zuccarino G. Round robin test for composite-to-
brick shear bond characterization. Mater Struct 2012;45:1761–1791.

[19] Monaldo E, Nerilli F, Vairo G. Effectiveness of some technical standards


for debonding analysis in FRP-concrete systems. Compos Part B: Eng
1455 2019;160:254–267.

[20] Monaldo E, Nerilli F, Vairo G. Technical standards for debonding in


FRP-concrete systems: an experimental contribution for basalt-FRP,
CICE 2018, Paris, France.

48
[21] Nerilli F, Vairo G. Experimental investigation on the debonding failure
1460 mode of basalt-based FRP sheets from concrete. Compos Part B: Eng
2018;153:205–216.

[22] Nerilli F, Vairo G. Strengthening of reinforced concrete beams with


basalt-based FRP sheets: An analytical assessment. In AIP Conference
Proceedings 2016 (Vol.1738, No.1, p.270016). AIP Publishing.

1465 [23] Nerilli F, Vairo G. Progressive damage in composite bolted joints


via a computational micromechanical approach. Compos Part B: Eng
2017;111:357–371.

[24] Nerilli F, Marino M, Vairo G. A numerical failure analysis of multi-


bolted joints in FRP laminates based on basalt fibers. Procedia Eng
1470 2015;109:492–506.

[25] Diab HM, Farghal OA. Bond strength and effective bond length of
FRP sheets/plates bonded to concrete considering the type of adhesive
layer. Compos Part B: Eng 2014;58:618–624.

[26] Serbescu A, Guadagnini M, Pilakoutas K. Mechanical characterization


1475 of basalt FRP rebars and long-term strength predictive model. J Com-
pos Constr 2014;19(2):04014037.

[27] High C, Seliem HM, El-Safty A, Rizkalla SH. Use of basalt fibers for
concrete structures. Constr Build Mater 2015;96:37–46.

[28] Asprone D, Cadoni E, Iucolano F, Prota A. Analysis of the strain-rate


1480 behavior of a basalt fiber reinforced natural hydraulic mortar. Cement
Concrete Comp 2014;53:52–58.

[29] Larrinaga P, Chastre C, Biscaia HC, San-José JT. Experimental and


numerical modeling of basalt textile reinforced mortar behavior under
uniaxial tensile stress. Mater Design 2014;55:66–74.

1485 [30] Ascione L, de Felice G, De Santis S. A qualification method for exter-


nally bonded Fibre Reinforced Cementitious Matrix (FRCM) strength-
ening systems. Compos Part B: Eng 2015;78:497–506.

[31] Lignola GP, Caggegi C, Ceroni F, De Santis S, Krajewski P, Lourenco


PB, Morganti M, Papanicolaou C, Pellegrino C, Prota A, Zuccarino

49
1490 L. Performance assessment of basalt FRCM for retrofit applications on
masonry. Compos Part B: Eng 2017;128:1–18.

[32] Escrig C, Gil L, Bernat-Maso E, Puigvert F. Experimental and analyti-


cal study of reinforced concrete beams shear strengthened with different
types of textile-reinforced mortar. Constr Build Mater 2015;83:248–260.

1495 [33] Rapp, G. Archaeomineralogy. Springer Science & Business Media 2009.

[34] Dhè P. Filament composed of basalt. U.S. Patent 1438428, 1922.

[35] Austin HF, Subramanian RV. Extrusion in the presence of a reducing


agent. U.S. Patent 4149866, 1979.

[36] Subramanian RV, Shu KHH. Silane coupling agents for basalt fiber
1500 reinforced polymer composites. In Ishida K, Kumar G, editors. Molec-
ular characterization of composite interfaces. New York: Springer 1985.
p.205–236.

[37] Park JM, Subramanian RV. Interfacial shear strength and dura-
bility improvement by monomeric and polymeric silanes in basalt
1505 fiber/epoxy single-filament composite specimens. J Adhes Sci Technol
1991;5(6):459–477.

[38] Lee JJ, Nam I, Kim H. Thermal stability and physical properties
of epoxy composite reinforced with silane treated basalt fiber. Fiber
Polym 2017;18(1):140–147.

1510 [39] Jung T, Subramanian RV. Strengthening of basalt fiber by alumina


addition. Scripta Metall et Mater 1993;28(4):527–532.

[40] Jung TH, Subramanian RV. Alkali resistance enhancement of basalt


fibers by hydrated zirconia films formed by the sol-gel process. J Mater
Res 1994;9(4):1006–1013.

1515 [41] Rapp CF, Fausey WH, Gonterman JR. Basalt compositions and their
fibers. U.S. Patent 4560606, 1985.

[42] Dzhigiris DD, Volynskii AK, Kozlovskii PP. Principles of production


technology for and properties of basalt fibers. Basalt fiber composite
materials and structures 1980, 54–81.

50
1520 [43] Dzhigiris DD, Makhova MF,Gorobinskaya VD. Continuous basalt fiber.
Glass Ceram 1983;40(9):467–470.

[44] Wojnarovits I. Mechanical change of basalt and glass wool fibres


during aqueous corrosion at different temperatures. TIZ-Fachberichte
1987;111(8):554–558.

1525 [45] Sokolinskaya MA, Zabava LK, Tsybulya TM, Medvedev AA. Strength
properties of basalt fibers. Glass Ceram 1991;48(10):435–437.

[46] Osnons S. Past, present and future of continuous basalt fibre. JEC
Composites 2007;35:24–25.

[47] Aslanova LG. Method and apparatus for producing basaltic fibers. U.S.
1530 Patent 20020069678, 2002.

[48] Brik VB. Multifunctional apparatus for manufacturing mineral basalt


fibers. U.S. Patent 6647747, 2003.

[49] Brik VB. Apparatus integrated with ceramic bushing for manufacturing
mineral/basalt fibers. U.S. Patent 20060218972, 2006.

1535 [50] Zhishen W, Gang W, Xin W, Xianqi H, Jianbiao J. New Progress


in R& D of Basalt Fibres and BFRP in Infrastructure Engineering.
chinagbf.com

[51] Brik VB. Performance evaluation of basalt fibers and composite rebars
as concrete reinforcement. Tech Res Report submitted to NCHRP-
1540 IDEA, Project 1999.

[52] Transparency Market Research. Continuous Basalt Fiber Market -


Global Industry Analysis, Size, Share, Growth, Trends and Forecast,
2013-2019. www.transparencymarketresearch.com 2014.

[53] Czigány T, Vad J, Pölöskei K. Basalt fiber as a reinforcement of poly-


1545 mer composites. Period Polytech Mech Eng 2005;49(1):3–14.

[54] Novitskii AG, Efremov MV. Technological aspects of the suitability of


rocks from different deposits for the production of continuous basalt
fiber. Glass Ceram 2013;69(11-12):409–412.

51
[55] Vas LM, Pölöskei K, Felhōs D, Deák T, Czigány T. Theoretical and
1550 experimental study of the effect of fiber heads on the mechanical prop-
erties of non-continuous basalt fiber reinforced composites. Express
Polym Lett 2007;1(2):109–121.

[56] Greco A, Maffezzoli A, Casciaro G, Caretto F. Mechanical properties


of basalt fibers and their adhesion to polypropylene matrices. Compos
1555 Part B: Eng 2014;67:233–238.

[57] Deák T, Czigány T. Chemical composition and mechanical properties


of basalt and glass fibers: A comparison. Text Res J 2009;79(7):645–
651.

[58] Liu J, Yang J, Chen M, Lei L, Wu Z. Effect of SiO2, Al2O3 on heat


1560 resistance of basalt fiber. Thermochimica Acta 2018;660:56–60.

[59] Raj S, Kumar VR, Kumar BB, Iyer NR. Basalt: structural insight as
a construction material. Sādhanā 2017;42(1):75–84.

[60] Gutnikov SI, Malakho AP, Lazoryak BI, Loginov VS. Influence of alu-
mina on the properties of continuous basalt fibers. Russ J Inorg Chem
1565 2009;54(2):191–196.

[61] Gur’ev VV, Neproshin EI, Mostovoi GE. The effect of basalt fiber
production technology on mechanical properties of fiber. Glass Ceram
2001;58:62–65.

[62] Dorigato A, Pegoretti A. Fatigue resistance of basalt fibers-reinforced


1570 laminates. J Compos Mater 2012;46(15):1773–1785.

[63] Sim J, Park C, Moon DY. Characteristics of basalt fiber as a strength-


ening material for concrete structures. Compos Part B: Eng 2005;36(6-
7):504–512.

[64] Wu Z, Wang X, Iwashita K, and Sasaki T, Hamaguchi Y. Tensile fa-


1575 tigue behaviour of FRP and hybrid FRP sheets. Compos Part B: Eng
2010;41(5):396–402.

[65] Lopresto V, Leone C, De Iorio I. Mechanical characterisation of basalt


fibre reinforced plastic. Compos Part B: Eng 2011;42(4):717–723.

52
[66] Colombo C, Vergani L, Burman M. Static and fatigue character-
1580 isation of new basalt fibre reinforced composites. Compos Struct
2012;94(3):1165–1174.
[67] Wei B, Cao H, Song S. Tensile behavior contrast of basalt and glass
fibers after chemical treatment. Mater Design 2010;31(9):4244–4250.
[68] Mingchao W, Zuoguang Z, Yubin L, Min L, Zhijie S. Chemical dura-
1585 bility and mechanical properties of alkali-proof basalt fiber and its re-
inforced epoxy composites. J Reinf Plast Comp 2008;27(4):393–407.
[69] Rabinovich FN, Zueva VN, Makeeva LV. Stability of basalt fibers in a
medium of hydrating cement. Glass Ceram 2001;58(11-12):431–434.
[70] Militký J, Kovačič V, Rubnerová J. Influence of thermal treatment on
1590 tensile failure of basalt fibers. Eng Fract Mech 2002;69(9):1025–1033.
[71] Hao L, Yu W. Evaluation of thermal protective performance of basalt
fiber nonwoven fabrics. J Therm Anal Calorim 2009;100(2):551–555.
[72] Ying S, Zhou X. Chemical and thermal resistance of basalt fiber in
inclement environments. Journal of Wuhan University of Technology-
1595 Mater. Sci. Ed. 2013;28(3):560–565.
[73] Palmieri A, Matthys S, Tierens M. Basalt fibres: Mechanical properties
and applications for concrete structures. Int Conf on Concrete Solutions
2009;CRC Press/Balkema:165–169.
[74] Chairman CA, Kumaresh Babu SP. Mechanical and abrasive wear be-
1600 havior of glass and basalt fabric-reinforced epoxy composites. J Appl
Polym Sci 2013;130(1):120–130.
[75] Ayub T, Shafiq N, Nuruddin MF. Mechanical properties of high-
performance concrete reinforced with basalt fibers. Procedia Engineer-
ing 2014;77:131–139.
1605 [76] Carmisciano S, Rosa IM, Sarasini F, Tamburrano A, Valente M. Basalt
woven fiber reinforced vinylester composites: Flexural and electrical
properties. Mater Design 2011;32(1):337–342.
[77] Fiore V, Scalici T, Di Bella G, Valenza A. A review on basalt fibre and
its composites. Compos Part B: Eng 2015;74:74–94.

53
1610 [78] Varley RJ, Tian W, Leong KH, Leong AY, Fredo F, Quaresimin M.
The effect of surface treatments on the mechanical properties of basalt-
reinforced epoxy composites. Polym Compos 2013;34(3):320–329.

[79] Wei B, Cao H, Song S. Surface modification and characterization of


basalt fibers with hybrid sizings. Compos Part A: Appl S 2011;42(1):22–
1615 29.

[80] Wei B, Song S, Cao H. Strengthening of basalt fibers with nano-


SiO2 –epoxy composite coating. Mater Design 2011,32(8-9):4180–4186.

[81] Wang GJ, Liu YW, Guo YJ, Zhang ZX, Xu MX, Yang ZX. Surface
modification and characterizations of basalt fibers with non-thermal
1620 plasma. Surf Coat Tech 2007;201(15):6565–6568.

[82] Bashtannik PI, Kabak AI, Yakovchuk YY. The effect of adhesion in-
teraction on the mechanical properties of thermoplastic basalt plastics.
Mech Compos Mater 2003;39(1):85–88.

[83] Manikandan V, Winowlin Jappes JT, Suresh Kumar SM, Amuthakkan-


1625 nan P. Investigation of the effect of surface modifications on the me-
chanical properties of basalt fibre reinforced polymer composites. Com-
pos Part B: Eng 2012;43(2):812–818.

[84] Zhang X, Pei X, Wang Q. Friction and wear properties of polyimide


matrix composites reinforced with short basalt fibers. Journal of Ap-
1630 plied Polymer Science 2009, 111(6), 2980–2985.

[85] Botev M, Betchev H, Bikiaris D, Panayiotou C. Mechanical properties


and viscoelastic behavior of basalt fiber-reinforced polypropylene. J
Appl Polym Sci 1999;74(3):523–531.

[86] Czigány T. Special manufacturing and characteristics of basalt fiber


1635 reinforced hybrid polypropylene composites: mechanical properties and
acoustic emission study. Compos Sci Tech 2006;66(16):3210–3220.

[87] Amuthakkannan P, Manikandan V, Jappes JW, Uthayakumar M. Ef-


fect of fibre length and fibre content on mechanical properties of short
basalt fibre reinforced polymer matrix composites. Mater Phys Mech
1640 2013;16:107–117.

54
[88] Subagia IA, Kim Y. A study on flexural properties of carbon-
basalt/epoxy hybrid composites. J Mech Sci Tech 2003;27(4):987–992.

[89] Subagia IA, Kim Y, Tijing LD, Kim CS, Shon HK. Effect of stacking
sequence on the flexural properties of hybrid composites reinforced with
1645 carbon and basalt fibers. Compos Part B: Eng 2014;58:251–258.

[90] Dorigato A, Pegoretti A. Flexural and impact behaviour of car-


bon/basalt fibers hybrid laminates. J Compos Mater 2014;48(9):1121–
1130.

[91] Amuthakkannan P, Manikandan V, Uthayakumar M. Mechanical prop-


1650 erties of basalt and glass fiber reinforced polymer hybrid composites. J
Adv Microsc Res 2014;9(1):44–49.

[92] Petrucci R, Santulli C, Puglia D, Sarasini F, Torre L, Kenny JM. Me-


chanical characterisation of hybrid composite laminates based on basalt
fibres in combination with flax, hemp and glass fibres manufactured by
1655 vacuum infusion. Mater Design 2013;49:728–735.

[93] Amuthakkannan P, Manikandan V, Jappes JTW, Uthayakumar M.


Influence of stacking sequence on mechanical properties of basalt-jute
fiber-reinforced polymer hybrid composites. J Polym Eng 2012;32(8-
9):547–554.

1660 [94] Sarasini F, Tirillò J, Valente M, Ferrante L, Cioffi S, Iannace S, Sor-


rentino L. Hybrid composites based on aramid and basalt woven fabrics:
Impact damage modes and residual flexural properties. Mater Design
2013;49:290–302.

[95] Wu G, Wang X, Wu Z, Dong Z, Zhang, G. Durability of basalt


1665 fibers and composites in corrosive environments. J Compos Mater
2015;49(7):873–887.

[96] Wei B, Cao H, Song S. Degradation of basalt fibre and glass fibre/epoxy
resin composites in seawater. Corros Sci 2011;53(1):426–431.

[97] C̆erný M, Glogar P, Sucharda Z. Mechanical properties of basalt fiber


1670 reinforced composites prepared by partial pyrolysis of a polymer pre-
cursor. J Compos Mater 2009;43(9):1109–1120.

55
[98] Landucci G, Rossi F, Nicolella C, Zanelli S. Design and testing
of innovative materials for passive fire protection. Fire Safety J
2009;44(8):1103–1109.

1675 [99] Shi JW, Zhu H, Wu ZS, Wu G. Durability of BFRP and hybrid FRP
sheets under freeze-thaw cycling. In Advanced Materials Research 2011
(Vol.163, pp.3297–3300).

[100] Shi J, Zhu H, Wu Z, Seracino R, Wu G. Bond behavior between basalt


fiber–reinforced polymer sheet and concrete substrate under the cou-
1680 pled effects of freeze-thaw cycling and sustained load. J Compos Constr
2012;17(4):530–542.

[101] Wang J, GangaRao H, Liang R, Liu W. Durability and prediction


models of fiber-reinforced polymer composites under various environ-
mental conditions: A critical review. J Reinforced Plastics Compos
1685 2016:35(3);179–211.

[102] Ceroni F. Experimental performances of RC beams strengthened with


FRP materials. Constr Build Mater 2010;24(9):1547–1559.

[103] Spadea G, Bencardino F, Sorrenti F, Swamy RN. Structural effec-


tiveness of FRP materials in strengthening RC beams. Eng Struct
1690 2015;99:631–641.

[104] Bilotta A, Ceroni F, Nigro E, Pecce M. Efficiency of CFRP NSM strips


and EBR plates for flexural strengthening of RC beams and loading
pattern influence. Compos Struct 2015;124:163–175.

[105] Berardi, VP, Feo L, Mancusi G, De Piano M. Influence of reinforcement


1695 viscous properties on reliability of existing structures strengthened with
externally bonded composites. Compos Struct 2018.

[106] Colangelo F, Russo P, Cimino F, Cioffi R, Farina I, Fraternali F, Feo


L. Epoxy/glass fibres composites for civil applications: Comparison
between thermal and microwave crosslinking routes. Compos Part B:
1700 Eng 2017;126:100–107.

[107] Zhou A, Qin R, Feo L, Penna R, Lau D. Investigation on interfacial


defect criticality of FRP-bonded concrete beams. Compos Part B: Eng
2017;113:80–90.

56
[108] Ascione F, Lamberti M, Razaqpur AG, Spadea S, Malagic M. Pseudo-
1705 ductile failure of adhesively joined GFRP beam-column connec-
tions: An experimental and numerical investigation. Compos Struct
2018;200:864–873.

[109] Ascione F, Lamberti M, Napoli A, Razaqpur G, Realfonzo R. An exper-


imental investigation on the bond behavior of steel reinforced polymers
1710 on concrete substrate. Composite Structures 2017;181:58–72.

[110] Carloni C, Ascione F, Camata G, de Felice G, De Santis S, Lamberti


M, ... Cescatti E. An Overview of the Design Approach to Strengthen
Existing Reinforced Concrete Structures with SRG. ACI Special Pub-
lication 2018;326:101-1.

1715 [111] de Felice G, De Santis S, Realfonzo R, Napoli A, Ascione F, Stievanin E,


... Camata G. State of the Art of Steel Reinforced Grout Applications
to Strengthen Masonry Structures. ACI Special Publication 2018;326:
102-1.

[112] Ascione F, Lamberti M, Napoli A, Realfonzo R. SRP/SRG Strips


1720 Bonded to Concrete Substrate: Experimental Characterization. ACI
Special Publication 2018;326:110-1.

[113] Gribniak V, Arnautov AK, Kaklauskas G, Tamulenas V, Timinskas E,


Sokolov A. Investigation on application of basalt materials as reinforce-
ment for flexural elements of concrete bridges. Balt J Road Bridge Eng
1725 2015;10(3):201–206.

[114] Ouyang LJ, Lu ZD, Chen WZ. Flexural experimental study on con-
tinuous reinforced concrete beams strengthened with basalt fiber rein-
forced polymer/plastic. Journal of Shanghai Jiaotong University (Sci-
ence) 2012;17(5):613–618.

1730 [115] Campione G, La Mendola L, Monaco A, Valenza A, Fiore V. Behavior


in compression of concrete cylinders externally wrapped with basalt
fibers. Compos Part B: Eng 2015;69:576–586.

[116] Ouyang LJ, Gao WY, Zhen B, Lu ZD. Seismic retrofit of square rein-
forced concrete columns using basalt and carbon fiber-reinforced poly-
1735 mer sheets: A comparative study. Compos Struct 2017;162:294–307.

57
[117] Oliveira DV, Basilio I, Lourenco PB. Experimental behavior of FRP
strengthened masonry arches. J Compos Constr 2010;14(3):312–322.

[118] Ibrahim AM, Wu Z, Fahmy MF, Kamal D. Experimental study on


cyclic response of concrete bridge columns reinforced by steel and basalt
1740 FRP reinforcements. J Compos Constr 2015;20(3):04015062.

[119] De Luca A, Nardone F, Matta F, Nanni A, Lignola GP, Prota A. Struc-


tural evaluation of full-scale FRP-confined reinforced concrete columns.
J Compos Constr 2010;15(1):112–123.

[120] Long, YL, Zhu J. Experimental study on concrete columns with various
1745 sizes confined by BFRP and hybrid FRP under axial compression. In
Advanced materials research 2014 (Vol.838, pp.407–411). Trans Tech
Publications.

[121] Lu ZD, Su L, Yu JT. Experimental study on the seismic behaviour


of strengthened concrete column-beam joints by simulated earthquake.
1750 Procedia Engineering 2011;14:1871–1878.

[122] Maruthachalam D, Muthukrishnan V. Behaviour of reinforced concrete


exterior beam column joint: A general review. Int J Adv Scient and
Tech Res 2012;6(2).

[123] Yu J, Shang X, Lu Z. Efficiency of externally bonded L-shaped FRP


1755 laminates in strengthening reinforced-concrete interior beam-column
joints. J Compos Constr 2015;20(3):04015064.

[124] Alaimo G, Valenza A, Enea D, Fiore V. The durability of basalt fi-


bres reinforced polymer (BFRP) panels for cladding. Mater Struct
2016;49(6):2053–2064.

1760 [125] Fedele R, Milani G. Three-dimensional effects induced by FRP-from-


masonry delamination. Compos Struct 2011;93(7):1819–1831.

[126] D’Altri AM, Carloni C, de Miranda S, Castellazzi G. Numerical mod-


eling of FRP strips bonded to a masonry substrate. Compos Struct
2018.

1765 [127] Focacci F, Carloni C. Periodic variation of the transferable load at the
FRP-masonry interface. Compos Struct 2015;129:90–100.

58
[128] Di Ludovico M, D’Ambra C, Prota A, Manfredi G. FRP confinement
of tuff and clay brick columns: Experimental study and assessment of
analytical models. J Compos Constr 2010;14(5):583–596.

1770 [129] Micelli F, Angiuli R, Corvaglia P, Aiello, MA. Passive and SMA-
activated confinement of circular masonry columns with basalt and
glass fibers composites. Compos Part B: Eng 2014;67:348–362.

[130] Zhou D, Lei Z, Wang J. In-plane behavior of seismically damaged ma-


sonry walls repaired with external BFRP. Compos Struct 2013;102:9–
1775 19.

[131] Ghiassi B, Oliveira DV, Lourenco PB. Recent developments in durabil-


ity of FRP-masonry systems. In International Conference on Rehabil-
itation and Restoration of Structures 2013 (pp. 107–116). Iit Madras.

[132] Ghiassi B, Marcari G, Oliveira DV. Lourenco PB. Simplified numerical


1780 analysis of bond degradation of FRP-masonry systems for durability
purposes. In Sixth International Conference of Seismology and Earth-
quake Engineering 2011 (pp. 1–8). Iiees.

[133] Shrive NG, Reda Taha MM, Masia MJ. Restoration and strengthening
with fibre reinforced polymers: Issues to consider. Proceedings of Struc-
1785 tural Analysis of Historical Constructions. Taylor and Francis Group
2004; 829–835.

[134] Shrive NG. The use of fibre reinforced polymers to improve seismic
resistance of masonry. Constr Build Mater 2006;20(4):269–277.

[135] Micelli F, Aiello MA. Residual tensile strength of dry and impregnated
1790 reinforcement fibres after exposure to alkaline environments. Compos
Part B Eng 2017.

[136] Quagliarini E, Monni F, Lenci S, Bondioli F. Tensile characterization


of basalt fiber rods and ropes: A first contribution. Constr Build Mater
2012;34:372–380.

1795 [137] Yuan F, Pan J, Leung CKY. Flexural behaviors of ECC and con-
crete/ECC composite beams reinforced with basalt fiber-reinforced
polymer. J Compos Constr 2013;17(5):591–602.

59
[138] Cosenza E, Manfredi G, Realfonzo R. Behavior and modeling of bond
of FRP rebars to concrete. J Compos Constr 1997;1(2):40–51.

1800 [139] Liu Q, Shaw MT, Parnas RS, McDonnell AM. Investigation of basalt
fiber composite mechanical properties for applications in transporta-
tion. Polym Composite 2006;27(1):41–48.

[140] Kocaoz S, Samaranayake VA, Nanni A. Tensile characterization of glass


FRP bars. Compos Part B: Eng 2005;36(2):127–134.

1805 [141] Gravina RJ, Smith ST. Flexural behaviour of indeterminate concrete
beams reinforced with FRP bars. Eng Struct 2008;30(9):2370–2380.

[142] Reyes-Araiza JL, Manzano-Ramı́rez A, Rubio-Avalos JC, González-


Sosa E, Pérez-Robles JF, Arroyo-Contreras M, ... Vorobiev YV. Com-
parative study on tensile behavior of inorganic fibers embedded in
1810 unsaturated polyester bisphenol A-styrene copolymer. Inorg Mater
2008;44(5):549–554.

[143] Galati N, Tumialan G, Nanni A. Strengthening with FRP bars of URM


walls subject to out-of-plane loads. Constr Build Mater 2006;20(1-
2):101–110.

1815 [144] Urbanski M, Lapko A, Garbacz A. Investigation on concrete beams


reinforced with basalt rebars as an effective alternative of conventional
R/C structures. Procedia Engineering 2013;57:1183–1191.

[145] Wu G, Wang X, Wu Z, Dong Z, Xie Q. Degradation of basalt FRP bars


in alkaline environment. Sci Eng Compos Mater 2015;22(6):649–657.

1820 [146] El Refai A, Ammar MA, Masmoudi R. Bond performance of


basalt fiber-reinforced polymer bars to concrete. J Compos Constr
2014;19(3):04014050.

[147] Patnaik A. Applications of basalt fiber reinforced polymer (BFRP)


reinforcement for transportation infrastructure. Developing a Research
1825 Agenda for Transportation Infrastructure 2009;175–184.

[148] Tomlinson D, Fam A. Performance of concrete beams rein-


forced with basalt FRP for flexure and shear. J Compos Constr
2014;19(2):04014036.

60
[149] Zhang J, Leung CK, Cheung YN. Flexural performance of lay-
1830 ered ECC-concrete composite beam. Compos Sci Technol 2006;66(11-
12):1501–1512.

[150] Ge W, Ashour AF, Cao D, Lu W, Gao P, Yu J, .. Cai C Experimental


study on flexural behavior of ECC-concrete composite beams reinforced
with FRP bars. Compos Struct 2018.

1835 [151] Yeboah D, Taylor S, McPolin D, Gilfillan R. Pull-out behaviour of


axially loaded Basalt Fibre Reinforced Polymer (BFRP) rods bonded
perpendicular to the grain of glulam elements. Constr Build Mater
2013;38:962–969.

[152] O’Neill C, McPolin D, Taylor SE, Martin T, Harte AM. Glued-in basalt
1840 FRP rods under combined axial force and bending moment: An exper-
imental study. Compos Struct 2018;186:267–273.

[153] Wu G, Wu ZS, Luo YB, Sun ZY, Hu X. Q. Mechanical properties


of steel-FRP composite bar under uniaxial and cyclic tensile loads. J
Mater Civil Eng 2010;22(10):1056–1066.

1845 [154] Sun ZY, Wu G, Wu ZS, Zhang M. Seismic behavior of concrete


columns reinforced by steel-FRP composite bars. J Compos Constr
2011;15(5):696–706.

[155] Quagliarini E, Monni F, Bondioli F, Lenci S. Basalt fiber ropes and


rods: Durability tests for their use in building engineering. J Build
1850 Eng 2016;5:142–150.

[156] Benmokrane B, Elgabbas F, Ahmed EA, Cousin P. Characterization


and comparative durability study of glass/vinylester, basalt/vinylester,
and basalt/epoxy FRP bars. J Compos Constr 2015;19(6):04015008.

[157] Altalmas A, El Refai A, Abed F. Bond degradation of basalt fiber-


1855 reinforced polymer (BFRP) bars exposed to accelerated aging condi-
tions. Constr Build Mater 2015;81:162–171.

[158] Monni F, Quagliarini E. Dry masonry strenghtening through basalt


fibre ropes: experimental results against out-of-plane actions. In Key
Engineering Materials 2015;624.

61
1860 [159] Larrinaga P, Chastre C, San-José JT, Garmendia L. Non-linear ana-
lytical model of composites based on basalt textile reinforced mortar
under uniaxial tension. Compos Part B: Eng 2013;55:518–527.

[160] Rambo DAS, De Andrade Silva F, Toledo Filho RD, Gomes ODFM.
Effect of elevated temperatures on the mechanical behavior of basalt
1865 textile reinforced refractory concrete. Mater Design 2015;65:24–33.

[161] Al-Salloum YA, Elsanadedy HM, Alsayed SH, Iqbal RA. Experimental
and numerical study for the shear strengthening of reinforced concrete
beams using textile-reinforced mortar. J Compos Constr 2011;16(1):74–
90.

1870 [162] Valluzzi MR, Da Porto F, Garbin E, Panizza M. Out-of-plane be-


haviour of infill masonry panels strengthened with composite materials.
Mater Struct 2014;47(12):2131–2145.

[163] Papanicolaou C, Triantafillou T, Lekka M. Externally bonded grids


as strengthening and seismic retrofitting materials of masonry panels.
1875 Constr Build Mater 2011; 25(2):504–514.

[164] Santandrea M, Imohamed IAO, Carloni C, Mazzotti C, de Miranda S,


Ubertini F. A study of the debonding mechanism in steel and basalt
FRCM-masonry joints. In Int. Brick Block Mason. Conf. 2016, Padua,
Italy (pp. 433–440).

1880 [165] Gopinath S, Murthy AR, Iyer NR, Prabha M. Behaviour of reinforced
concrete beams strengthened with basalt textile reinforced concrete. J
Ind Text 2015;44(6):924–933.

[166] Elsanadedy HM, Almusallam TH, Alsayed SH, Al-Salloum YA. Flex-
ural strengthening of RC beams using textile reinforced mortar – Ex-
1885 perimental and numerical study. Compos Struct 2013;97:40–55.

[167] Maddaloni G, Cascardi A, Balsamo A, Di Ludovico M, Micelli F, Aiello


MA, Prota A. Confinement of full-scale masonry columns with FRCM
systems. In Key Engineering Materials 2017 (Vol. 747, pp. 374–381).
Trans Tech Publications.

62
1890 [168] Valluzzi MR. Challenges and perspectives for the protection of ma-
sonry structures in historic centers: the role of innovative materials
and techniques. RILEM Technical Letters, 1, 45–49.
[169] Abdulsalam B, Ali AH, The preservation of historical masonry her-
itage structures using advanced composite materials. Int J Eng Sci
1895 2015;4(8):14-24.
[170] Giacomin G. Innovative strengthening materials for the post-
earthquake reconstruction of L’Aquila masonries. In Structural Anal-
ysis of Historical Constructions: Anamnesis, Diagnosis, Therapy, Con-
trols: Proceedings of the 10th International Conference on Structural
1900 Analysis of Historical Constructions 2016. CRC Press.
[171] Di Ludovico M, Prota A, Manfredi G. Structural upgrade using basalt
fibers for concrete confinement. J Compos Constr 2010;14(5):541–552.
[172] Kabay N. Abrasion resistance and fracture energy of concretes with
basalt fiber. Constr Building Mater 2014;50:95–101.
1905 [173] Jiang C, Fan K, Wu F, Chen D. Experimental study on the mechan-
ical properties and microstructure of chopped basalt fibre reinforced
concrete. Mater Design 2014;58:187–193.
[174] Abdulhadi M. A comparative study of basalt and polypropylene fibers
reinforced concrete on compressive and tensile behavior. Int J Eng
1910 Trends Tech 2014;9:295–300.
[175] Lipatov YV, Gutnikov SI, Manylov MS, Zhukovskaya ES, Lazoryak BI.
High alkali-resistant basalt fiber for reinforcing concrete. Mater Design
2015;73:60–66.
[176] Jiang CH, McCarthy TJ, Chen D, Dong QQ. Influence of basalt fiber
1915 on performance of cement mortar. In Key Engineering Materials 2010
(Vol.426, pp.93–96). Trans Tech Publications.
[177] Borhan TM. Properties of glass concrete reinforced with short basalt
fibre. Mater Design 2012;42:265–271.
[178] Santarelli ML, Sbardella F, Zuena M, Tirillò J, Sarasini F. Basalt fiber
1920 reinforced natural hydraulic lime mortars: A potential bio-based ma-
terial for restoration. Mater Design 2014;63:398–406.

63
[179] Comité Euro-International du Béton. CEB-FIP Model Code 1990 Red-
wood Books. Trowbridge. Wiltshire. UK; 1993.

[180] Shaikh FUA. Review of mechanical properties of short fibre reinforced


1925 geopolymer composites. Constr Build Mater 2013;43:37–49.

[181] Dias DP, Thaumaturgo C. Fracture toughness of geopolymeric


concretes reinforced with basalt fibers. Cement Concrete Comp
2005;27(1):49–54.

[182] Li W, Xu J. Mechanical properties of basalt fiber reinforced geopoly-


1930 meric concrete under impact loading. Mat Sci Eng: A-Struct
2009;505(1-2):178–186.

64

You might also like