You are on page 1of 10

Composite Structures 245 (2020) 112317

Contents lists available at ScienceDirect

Composite Structures
journal homepage: www.elsevier.com/locate/compstruct

Analysis of failure modes for a non-crimp basalt fiber reinforced epoxy T


composite under flexural and interlaminar shear loading
Indraneel R. Chowdhury, Niamh H. Nash, Alexandre Portela, Noel P. O'Dowd, Anthony J. Comer

School of Engineering, Irish Centre for Composite Materials (IComp), Bernal Institute, University of Limerick, Limerick V94 T9PX, Ireland

ARTICLE INFO ABSTRACT

Keywords: This study investigates the mechanical properties (interlaminar shear and flexural strength) and failure modes of
Basalt fiber a basalt/epoxy composite, manufactured using a non-crimp-fabric (NCF) with a vacuum assisted resin infusion
Non-crimp-fabric process. Under flexural bending, damage initiated on the compression side between 20 and 50% of peak load and
Resin infusion progressed from ply to ply with increasing load. Failure at the tension surface of the flexural bending specimen
Failure modes
was confined to the bottom ply and was evident only close to final failure. Fiber kinking was the dominant
Optical microscopy
failure mechanism on the compression side whereas fiber breakage was the dominant mechanism on the tension
side. Regarding interlaminar shear, interlaminar shear cracks initiated once samples were subjected to stress
levels above 50% of peak stress and grew until failure with the crack following the fibre matrix interface of 90°
tows. Overall, comparing with values available in the open literature, the NCF basalt/epoxy composite out­
performed plain-woven basalt/epoxy and plain-woven E-glass/epoxy composites in terms of both flexural and
interlaminar shear strength but demonstrated lower strength than NCF E-glass/epoxy composite.

1. Introduction packed fiber tows with low through thickness permeability. However,
VaRTM is cost-effective and eliminates the part size constraint asso­
Renewable energy generation technologies are key solutions to re­ ciated with autoclaves.
duce carbon emissions ameliorating the impact of anthropogenic cli­ The properties of different fibers used in composite materials are
mate change [1]. In 2017, wind energy had a worldwide installed ca­ listed in Table 1. Typically, E-glass fibers (borosilicate glass is called ‘E-
pacity of 539 GW; this is expected to reach 840 GW by the end of 2020, glass’ or ‘electric glass’ because of its high electrical resistance) are used
accounting for nearly 20% of the global energy requirement [2]. Tidal as reinforcement in turbine rotor blades, for example, in the blade skins
and wave energy installed capacity is expected to reach 337 GW by and shear webs of the main spar [12]. Carbon fibers are often used
2050 [3]. For both the wind and tidal energy sector, the dominant alongside E-glass fibers in the spar cap section in order to increase the
generation method is by horizontal axis turbines [4,5]. Fiber compo­ bending stiffness of the blade [7]. Alternatives to E-glass include glass
sites are commonly used in turbine rotor blade sections because of their fibers with modified compositions, such as S-glass (high strength glass),
high specific stiffness, strength and fatigue resistance [6]. The main as well as carbon, aramid and basalt fibers [12]. S-glass fibers exhibit
spar of the turbine blade section, which runs from the blade root to the approximately 40% higher tensile and flexural strength as well as
tip [7], is the main load bearing component. Unidirectional (UD) non- 10–20% higher compressive and flexural modulus, compared to E-glass
crimp fabric (NCF) based composites are generally used for the spar fibers [12,13]. However, S-glass is more expensive (approx. 20%) than
[8,9,10]. In such composites, the unidirectional fiber bundles are E-glass [12]. Carbon fibers have much higher stiffness and lower den­
aligned along the blade axis and are often stitched to fiber bundles sity than E-glass [11] but again are more expensive [12,14,15,16].
aligned in an off-axis direction [6]. Aramid fibers (sold commercially as Kevlar® and Twaron®) have higher
Vacuum-assisted resin transfer molding (VaRTM) also designated as tensile strength and superior damage tolerance to E-glass. However,
resin infusion under flexible tooling (RIFT-2) process [11], is an out-of- they have low compressive strength, absorb more moisture and degrade
autoclave process commonly used for the manufacture of turbine under ultraviolet radiation [17]. Another potential alternative to E-
blades. It can be challenging to achieve optimal fiber volume fraction glass is basalt [12]. Basalt fibers are approx. 30% stronger, 15–20%
when using VaRTM with NCF, as such fabrics tends to have densely stiffer and 8–10% lighter than E-glass fibres [12,18,19] and are cheaper


Corresponding author.
E-mail address: anthony.comer@ul.ie (A.J. Comer).

https://doi.org/10.1016/j.compstruct.2020.112317
Received 29 November 2019; Accepted 30 March 2020
Available online 03 April 2020
0263-8223/ © 2020 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/BY/4.0/).
I.R. Chowdhury, et al. Composite Structures 245 (2020) 112317

Table 1 basalt/epoxy composites were manufactured using vacuum assisted


Comparison of properties of fibers commonly used in fiber reinforced polymer resin infusion using a vacuum bag in numerous studies [27,29,30,32].
composites [12,18,52–60,62]. In other studies, hand lamination using an impregnation roller tech­
A. Mechanical properties nique followed by a hot mould press [28] and Resin transfer-moulding
using a fixed cavity mould [31] was utilised for manufacturing plain-
Fiber material Elastic modulus Tensile strength Density (g/ woven basalt/epoxy composites. Flexural strength values reported
(GPa) (MPa) cm3)
varied from approx. 229 MPa [27] to 505 MPa [32] with vacuum as­
Basalt 89 4840 2.8 sisted resin infusion used in both cases. Flexural modulus varied from
E-glass 72.4 3450 2.6 approx. 4.8 GPa [28] to 23 GPa [32] for hand lamination and vacuum
S-glass 85.5 4580 2.5 assisted resin infusion respectively. The most commonly observed
Carbon (high strength) 240 3500 1.75
failure modes for plain woven basalt/epoxy under flexural loading were
Carbon (high modulus) 350–650 2500–4000 1.95
Aramid 70–112 2900–3000 1.44 fiber pull-out at tension side, fiber–matrix debonding [27,28], fiber
buckling and kink bands on the compression side [30,31] and fiber
B. Fiber diameter breakage on the tension side [30,32].
Interlaminar shear strength (ILSS) was reported in [29,31–34]. ILSS
Fiber material Fiber diameter (µm) values varied from approx. 18.9 MPa for vacuum assisted resin infusion
Basalt 9–23
using fixed cavity mould [33] to 41 MPa [32]. The interlaminar shear
Glass 9–13 strength of bi-directional woven basalt/epoxy composite was reported
Carbon 4–7.5 to be 60 MPa [35]. Laminates were manufactured using positive pres­
Aramid 5–18 sure infusion of 1 MPa in a fixed cavity mould and cured under pressure
and vacuum for 2 h at 50 °C and 2 h at 80 °C [35].
Table 2 Plain-woven basalt/epoxy composites have been compared to plain-
Chemical composition of basalt and E-glass fibers [18,61]. woven E-glass using the same manufacturing technique. The flexural
strength and interlaminar shear strength were found to be up to 55%,
Oxides content (wt. %) Basalt E-glass
and 50% higher, respectively, for the basalt/epoxy composite [32,34].
SiO2 47.5–53.0 53.4 In contrast, the interlaminar shear strength of plain woven basalt
Al2O3 13.3–18.0 14.3 fabric/epoxy based vinylester resin was approx. 20% higher than an E-
Fe2O3 7.0–14.0 0.28 glass based laminate [36], but flexural strength was lower by approx.
CaO 8.0–11.0 19.0
MgO 3.5–5.0 3.3
15% [36].
B2O3 0.8 10.3 In another study, ILSS and flexural properties of NCF basalt/epoxy
TiO2 0.2–3.5 0.14 composites were evaluated [37]. Laminates were manufactured uti­
Na2O + K2O 2.5–6.0 0.29 lising vacuum assisted resin infusion using the vacuum bag technique.
ZrO2 0.0 0.8
ILSS of NCF basalt/epoxy composites was reported to be 44 MPa
MnO 0.17–0.22 N/A
whereas the flexural strength and modulus were reported to be
698 MPa and 38.4 GPa respectively [37]. NCF basalt/epoxy composites
than carbon fiber [12,17,18]. The chemical mechanical and geome­ outperformed NCF E-glass/epoxy by approx. 15% in terms of flexural
trical properties of basalt fibres were evaluated and compared with E- strength whereas ILSS of NCF E-glass/epoxy was approx. 8% higher
glass fibres [19]. The results revealed that basalt and E-glass fibre have than that of NCF basalt/epoxy thus demonstrating a better interfacial
similar elemental composition with basalt fibres demonstrating higher strength [37]. In a different study relating to NCF E-glass based lami­
tensile strength than E-glass [19]. Basalt fibers exhibit good resistance nates, the interlaminar shear and flexural strength were reported to be
to chemical degradation and moisture absorption, are non-combustible 58 MPa [38] and 917 MPa [39], respectively. Laminates were manu­
and possess good resistance to acidic or alkaline media [17,20]. When factured by vacuum assisted resin infusion with a vacuum bag and
exposed to corrosive media basalt fibers exhibit better mechanical epoxy Prime™27 resin. Laminates were post-cured at 60 °C for 7 h. Fiber
properties than glass fibers [21]. Indeed, the use of basalt fibers buckling/kinking at the compression side was the failure mechanism
alongside carbon fibers as hybrid reinforcement for small wind turbine observed for the interlaminar shear test [38].
blades demonstrated encouraging results [12,22,23]. Among different The majority of the results in the literature are for plain-woven
‘natural’ fibers, basalt fiber is regarded as a mineral fiber [18], which is basalt fabric-based composites and little attention has been afforded to
produced directly from basalt rock with no additional ingredients non-crimp-fabric (NCF), which is widely used in industry in conjunction
[18,24]. The manufacturing process for basalt fiber is similar to glass with resin infusion manufacturing processes for structural applications
fiber with less energy consumption [18]. A comparison between the such as wind and tidal turbine blades. The overarching aim of the
different chemical constituents of basalt and E-glass fibers is provided current study is to understand the different failure modes occurring in
in Table 2. An important difference between basalt and E-glass is the NCF basalt fiber/epoxy composites. This will be achieved by evaluating
relatively large proportion of Fe3O4 (ferric oxide) in basalt [18,25], the mechanical properties of a basalt/epoxy composite under flexural
which, as it is a natural nucleating agent, imparts high thermal stability and interlaminar shear loading and by analyzing the failure modes at
to basalt [25]. Epoxy resin is currently the most common matrix ma­ different stress levels (20%, 50%, 80% of failure and full failure) using
terial used in conjunction with basalt fibers [18]. It has been demon­ optical microscopy.
strated that fiber sizing has a significant effect on the mechanical and
chemical bonding of basalt fibers with a polymer matrix (poly­ 2. Materials and methods
propylene) [26].
The majority of studies on basalt/epoxy composites in the literature Basalt fiber (BAS UNI 350, non-crimp fabric (NCF), 267 kg/dm3)
are for plain-woven fabric based composites [27–36]. The values re­ was sourced from Basaltex (Belgium) [40]. An image of the as-received
ported for flexural strength, flexural modulus and interlaminar shear fabric is shown in Fig. 1. The orientation of the 0° longitudinal basalt
strength depend not only on the material system but also strongly on fiber tows, 90° transverse basalt fiber tows, and the polyester stitches
the manufacturing method used which include vacuum assisted resin are indicated in Fig. 1(a) and (b). The total areal weight of the re­
infusion, hand lamination and resin transfer moulding. lain-woven inforcement fabric is 416 g/m2 [40] with 85% of the basalt fibers
aligned in the 0° direction and 13% of the basalt fibers aligned in the

2
I.R. Chowdhury, et al. Composite Structures 245 (2020) 112317

Fig. 1. Non-crimp basalt fabric structure: (a) front, (b) back. Different components of the fabric are identified: (i) 0° unidirectional basalt fiber tows, (ii) 90°
stabilizing basalt fiber tows, (iii) polyester stitches.

Fig. 2. Vacuum assisted resin infusion using vacuum bagging manufacturing arrangement: (i) resin container (ii) layup arrangement, (iii) cross-section of layup
arrangement, (iv) resin trap, (v) vacuum pump.

90° direction. The remainder (2%) is polyester stitching which, pro­ 4:1 part by weight ratio. The mounted specimens were ground and
vides stability to the fabric, preventing misalignment of the 0° fibers polished using an EcoMet™ semi-automatic Buehler grinder polisher
[41]. The average fiber diameter is 18 ± 4 μm. Epoxy resin Prime™27 machine. Images were captured using a Zeiss Axio Imager 2 Polarized
and slow hardener (Prime 20) were sourced from Gurit [42]. Prime™27 light microscope. The mounting epoxy filled gaps created by crack
is a low viscosity resin (480 – 510 mPa∙s at room temperature) with a propagation and delamination during mechanical testing. ImageJ soft­
gel time of 2 h 40 min at 25 °C and is suitable for the vacuum assisted ware has been used to mask regions containing mounting epoxy for the
resin infusion process [42]. A schematic of the manufacturing process is purpose of improving the contrast between pristine matrix material and
shown in Fig. 2. Before laying up the fibers, a semi-permanent solvent cracks or delaminations.
based epoxy release reagent (Loctite 770-NC Frekote) was applied to Fiber volume fraction (Vf) was determined using the burn-off
the glass tool surface. A dry- fiber stacking sequence of [0°]16 was used method [43] with specimen dimensions of 2 mm × 2 mm × 6 mm.
with the 90° tows facing towards the surface of the dry fiber stack. The Three samples were tested and Vf was calculated using Eq. (1):
fiber preform was covered with a layer of peel ply and a layer of dis­
tribution media. This arrangement was contained within a vacuum c mf
membrane secured by sealant tape to the glass tool. The epoxy resin and Vf = 100
f mi (1)
hardener were mixed at 100:28 parts by weight ratio and a vacuum
pump was used to draw resin into the dry fiber preform at ambient
where mi = initial mass of the specimen, mf = mass of fibers after
temperature and consolidate the laminate. The laminate was cured at
combustion (g), ρc = density of composite, ρf = density of fiber.
60 °C for 7 h, in accordance with the supplier guidelines [42].
Dynamic mechanical thermal analysis (DMTA) was performed to
Specimens were extracted from the manufactured laminates using a
measure the glass transition temperature (Tg) and to ensure that spe­
water-cooled semi-automatic panel cutter with a diamond-coated cir­
cimens were fully cured [44]. DMTA was performed using a DMA Q800
cular disc blade. Samples for mechanical testing were extracted parallel
instrument from TA instruments. Specimen dimensions were
to the 0° fiber direction i.e. the orientation of the 0° reinforcing fibers is
60 mm × 12 mm × 6 mm. A 3-point bending configuration was used to
parallel to the specimen’s largest dimension. Samples for imaging were
determine Tg using the tan delta peak method [44]. Six samples were
mounted using Epoxy Cure™ resin and epoxy slow hardener, mixed at
tested. The displacement amplitude was 10 μm, frequency was 1 Hz and

3
I.R. Chowdhury, et al. Composite Structures 245 (2020) 112317

heating rate was 5 °C/min.


Flexural strength and ILSS tests were conducted in accordance with
the relevant ASTM standards [45,46]. Nominal specimen dimensions
were 150 mm × 13 mm × 6 mm (L × W × t) for the flexural samples
and 36 mm × 12 mm × 6 mm (L × W × t) for the ILSS samples. The
span to thickness ratio was 20:1 (120 mm span length) and 4:1 (24 mm
span length) for the flexural and ILSS specimens, respectively. For both
flexural and ILSS tests, samples were located in the test fixture such that
the bottom ply containing the 90° fibers was under tension and the top
ply containing 0° fibers was under compression. The tests were per­
formed using a Tinius Olsen universal testing machine with a 25 kN
load cell. A crosshead speed of 1 mm/min was used for all tests. Eight
specimens of each type were tested: five specimens were tested to
failure and three further tests of each type were interrupted at 20%,
50% and 80% of the average strength determined from the failed spe­ Fig. 3. Flexural stress vs. strain curves: samples tested to failure (5 samples).
cimens. Following testing, the specimens were sectioned and prepared
for optical microscopy, as described earlier.
The flexural stress, σ, flexural strain, ε and interlaminar shear
strength (ILSS) were determined using Eqs. (2)–(4) respectively:
3PL
=
2bh2 (2)

6 h
=
L2 (3)

P
ILSS = 0.75
bh (4)

where P is the applied load (N), L is span (mm), b is specimen width


(mm), h is the specimen thickness (mm) and δ is the mid-span dis­
placement (mm).
Fig. 4. Flexural stress vs. strain curves: interrupted tests (3 samples).
3. Results and discussions

This section summarises the results from the various physical and
mechanical test methods described in section 2. The average fiber vo­
lume fraction was determined to be 55.3 ± 0.5%; the average glass
transition temperature (Tg) was measured to be 89.3 ± 3.8 °C.
The results of the mechanical testing are summarised in Table 3;
stress–strain/displacement curves are presented in Figs. 3–6. The
average peak flexural strength is 630 ± 9 MPa and the average peak
interlaminar shear strength is 47 ± 3 MPa. The flexural modulus was
determined to be 22.7 ± 0.1 GPa.
The flexural stress–strain and interlaminar shear-displacement
curves for the basalt/epoxy specimens are shown in Figs. 3 and 5. The
flexural stress–strain curves were near linear up to peak load, with
failure occurring subsequently. The interlaminar shear-displacement
Fig. 5. Interlaminar shear stress vs. displacement curves: samples tested to
curves went nonlinear from around 80% of failure. In Figs. 4 and 6, the
failure (5 samples).
results for the interrupted tests are shown. The line labelled ‘peak stress’
in each figure is the average of the 5 tests taken to failure (from Figs. 3
and 5, respectively). Single specimen tests were interrupted at 20%, surrounded by resin-rich regions in the interlaminar region. The NCF in
50% and 80% of the average peak stress (630 MPa and 47 MPa for the current study has a [0°]16 lay-up sequence resulting in laminates
flexural strength and ILSS, respectively) corresponding to 126 MPa, with a thickness of just under 6 mm.
315 MPa and 504 MPa for the flexural tests and at 9.3 MPa, 23.2 MPa The results for flexural strength and interlaminar shear strength of
and 37.2 MPa for the ILSS test. the NCF basalt/epoxy composites obtained from the present study have
A cross-sectional view of the as-manufactured basalt fiber reinforced been benchmarked against selected studies from the literature which
polymer composite specimen is shown in Fig. 7. No significant voids focus on plain-woven basalt/epoxy composites [27–34,36], plain-
can be seen at this magnification indicating good resin impregnation of woven E-glass/epoxy composites [32,34,36], NCF Basalt/epoxy [37]
the fiber bundle. The 90° fiber tows (labelled (ii) in Fig. 7) are and NCF E-glass/epoxy composite [37,38,39] (See Figs. 8 and 9). The

Table 3
Dimensions and results for specimens tested for flexural and interlaminar shear strength (ILSS) tests.
Test Thickness (mm) Width (mm) Length (mm) Span to thickness ratio Flexural strength (MPa) Flexural modulus (GPa) ILSS (MPa)

Flexure 5.89 ± 0.02 13.5 ± 0.2 150.0 ± 0.1 20:1 630 ± 9 22.7 ± 0.1 –
ILSS 5.76 ± 0.07 11.7 ± 0.1 34.5 ± 0.8 4:1 – – 47 ± 3

4
I.R. Chowdhury, et al. Composite Structures 245 (2020) 112317

However, the NCF basalt/epoxy composites from this study demon­


strated a reduction in strength of approximately 31% compared to NCF
E-glass/epoxy laminates [39]. The E-glass composites in [39] were
manufactured using the same manufacturing technique (vacuum as­
sisted resin infusion using a vacuum bag) and the same epoxy matrix
(Prime™27) used in the current study. However, the NCF E-glass fabric
contained higher percentage of 0° fibres (approx. 90%) compared to
approx. 85% in the NCF basalt fabric, which may have contributed to
the enhanced strength.
The interlaminar shear strength for the NCF basalt/epoxy con­
sidered in this study is compared to data available for plain-woven
basalt/epoxy [29,31–34] and plain-woven E-glass/epoxy [32,34,36] in
Fig. 9. ILSS of NCF basalt/epoxy was up to 55% higher than that of
plain-woven basalt/epoxy in [29,32,33], which used a similar manu­
Fig. 6. Interlaminar shear stress vs. displacement curves: interrupted tests (3 facturing process (vacuum assisted resin infusion with a vacuum bag).
samples). Similarly, the interlaminar shear strength of NCF basalt/epoxy com­
posites from this study was up to 40% higher than that for plain-woven
basalt/epoxy in [31,36], where resin transfer moulding process using a
fixed cavity mould was utilised as the manufacturing process. ILSS of
NCF basalt/epoxy was approx. 60% higher than that of plain-woven
basalt/epoxy in [33,34] where vacuum assisted resin infusion process
using fixed cavity mould and hand lamination using impregnated roller
technique were utilised for manufacturing. Apart from demonstrating
superior interlaminar shear strength compared to plain-woven basalt/
epoxy, NCF basalt/epoxy also demonstrated higher interlaminar shear
strength (approx. 23%, 80% and 46%) than plain-woven E-glass/epoxy
from the literature [32,34,36]. However, ILSS of the NCF basalt/epoxy
was approx. 6% higher than NCF basalt/epoxy [37] but approx. 2%
lower than NCF E-glass/epoxy [37]. Similarly, ILSS of NCF basalt/
epoxy was approx. 20% lower than NCF E-glass/epoxy with the same
matrix and manufacturing method [38].

Fig. 7. Cross-section of the untested basalt based composite: (i) 0° unidirec­ 3.1. Failure modes in flexural bending tests
tional basalt fiber tows, (ii) 90° stabilizing basalt fiber tows, (iii) polyester
stitches. 5 of 16 plies are included in the figure. Ply boundaries marked with
Composite failure is a combination of different failure modes, which
white dashed lines. Inset to the figure indicates the direction of view.
can occur in any of the constituents, their interfaces and by the inter­
action between them [47]. Under 3 point bending, flexural loading with
NCF basalt/epoxy composite from the present study demonstrated a relatively high span-to-thickness ratio, the bending stress is sig­
higher flexural strength (ranging between 20 and 70% higher) com­ nificantly larger than the maximum shear stress and the damage is a
pared to those obtained from the literature for plain-woven basalt/ combination of compressive and tensile failure modes at the top and
epoxy composites. [27–32,36]. In contrast, the flexural strength of NCF bottom of the test specimen, respectively. Generally, for a unidirec­
Basalt/epoxy laminate considered in this study was approx. 10% lower tional specimen with 0° fibers in the span direction, compressive failure
than the NCF Basalt/epoxy composites studied in the literature [37]. is expected to dominate as the compressive strength is typically lower
The same NCF basalt fabric (BAS UNI 350) was used in both studies. than the tensile strength.

Fig. 8. Comparison of flexural strength values from present study to those obtained from the literature.

5
I.R. Chowdhury, et al. Composite Structures 245 (2020) 112317

Fig. 9. Comparison of interlaminar shear strength values from present study to those obtained from the literature.

Fig. 10 illustrates the evolution of damage in the flexural specimen highly loaded fiber just prior to failure. Failure of the 0° degree fibers
from the interrupted tests (each image is from a different specimen may cause a sudden release of energy leading to the interply delami­
loaded to a different stress level). Typical failure modes observed on the nation and the cracks observed in Fig. 12 in the 90° tows. It may be
compression side (top ply indicated with a white star in Fig. 10b, d, f, noted that neither the transverse cracks or interply cracking seen in
and h) are fiber kinking (a buckling mechanism [48]), intra-ply dela­ Fig. 12 are evident at 80% of failure load(Fig. 10e). Although the
mination (delamination within a ply) and inter-ply delamination (de­ presence of 90° fiber tows provides stability to the fabric structure, they
lamination at the interface of two plies). On the tension side (bottom are aligned transverse to the loading direction and exhibit a relatively
ply indicated with a white star in Fig. 10 a, c, e, and g), fiber breakage low interfacial fracture toughness. Therefore, transverse shear crack
(0° fibers), intra-ply delamination (delamination inside one ply) and development in the 90° tows may also initiate delamination between
transverse shear cracking in the 90° tows are observed. Damage pro­ the 0° and 90° tows. In summary, for the flexural tests, fiber buckling/
pagation on the compression side is found to progress from ply to ply as kinking at the compression side and fiber breakage at the tension side
the load increases. For the specimen tested to 20% of peak stress were the dominating failure modes observed for NCF basalt/epoxy
(Fig. 10b), no evidence of fiber kinking is observed on the compression which are similar to the failure modes observed for plain-woven basalt/
side. At 50% of peak stress (Fig. 10d), evidence of fiber kinking is ob­ epoxy from the literature [27,28,30–32].
served in the top ply. Significant evidence of fiber kinking is observed
for the test specimen at 80% of peak stress, where kinking has pro­
3.2. Failure modes in interlaminar shear strength (ILSS) tests
gressed to the second ply (Fig. 10f). Further damage progression affects
the third ply of the specimen tested to failure (Fig. 10h). Fiber kinking is
Failure in an ILSS test is generally by crack formation between two
referred to as a phenomenon where layers of fibers under compression
neighbouring plies at mid thickness [52]. A short span to thickness ratio
undergo local buckling until a fiber breaks suddenly [49]. Although
is used for the ILSS test so that the shear stresses are significantly higher
several failure modes are apparent at low-stress levels (50% of peak
than the bending stresses. The shear stress varies parabolically through
stress) on the compression side, damage is only observed on the tension
the thickness, initiating failure at the mid-thickness of the specimen
side for the specimen tested to failure (Fig. 10g). No evidence of da­
[33].
mage is observed on the tension side for the specimens tested to 20%,
Fig. 13 illustrates the evolution of damage at the mid-thickness re­
50% and 80% of peak stress (Fig. 10a, c, and e). The zero degree fibers
gion of an ILSS specimen from the interrupted tests (each image is from
close to the surface are the most highly loaded fibers in compression
a different specimen loaded to a different stress level). The white star
during the flexure test. Kink bands formed at relatively low load levels
indicates the 8th and 9th plies (of 16 in total). Fig. 13(b) shows that
(50 and 80% of failure) and ultimately resulted in the fracture of the 0°
damage has occurred at the mid thickness region between the 8th and
fibers. Fig. 11 is a higher magnification image of Fig. 10(h) illustrating
9th ply somewhere between 50% and 80% of the failure load. This
the failure modes on the compression side of a specimen tested to
interlaminar cracking causes the ILSS stress-displacement curve to be­
failure. Evidence of intra-ply delamination can also be seen in Fig. 11.
come non-linear above 80% of failure load (see Fig. 6). It is evident
This is due to the kinking causing separation of the 0° and 90° tows.
from Fig. 13(b) that the crack propagates along the fibre–matrix in­
Delamination possibly occurs at the same time as kink band for­
terface of the 90° fiber tows.
mation as evident at both 50% (Fig. 10d) and 80% (Fig. 10f) of failure.
Fig. 14 illustrates the evolution of damage at the tension and
The separation between fiber and matrix is commonly referred to as
compression sides of the ILSS specimen. It is evident that no damage is
debonding [49], which can occur following fiber fracture [50,51]. Here,
observed even up to 80% of the peak load either on the compression
it could be caused by the sudden release of energy, which occurs during
side (Fig. 14b, d) or the tension side (Fig. 14a, c), indicating that the
kink band formation or during delamination.
failure mode for the specimen is shear dominated. The damage occur­
Fig. 12 is a high magnification image of Fig. 10(g), which is an
ring at peak stress on the tension side and on the compression, side is
image of the tension side at failure. 0° fiber breakage, interply dela­
shown in Fig. 14e and f. This damage occurs close to final failure after
mination and transverse cracks in 90° tows are evident. On the tension
damage has occurred at mid-thickness. This failure mode in NCF basalt/
side, damage is often initiated by fiber breakage that can lead to cat­
epoxy is different from that observed for NCF E-glass/epoxy in [38]
astrophic failure [49]. This failure mode is evident only in the bottom
where the specimen failed due to micro buckling/kinking at the com­
two plies on the tension side. The 0° fibers are expected to be the most
pression side designating a strong interfacial strength [38].

6
I.R. Chowdhury, et al. Composite Structures 245 (2020) 112317

Fig. 10. Cross-section of composite tested under flexural loading showing tension side (left) and compression side (right) at mid-span of the specimen: (a, b) 20% of
peak stress, (c, d) 50% of peak stress, (e, f) 80% of peak stress and (g, h) loaded to failure. Ply boundaries are indicated with white dashed lines. The top ply on the
compression side and bottom ply on the tension side indicated with a white star.

4. Conclusions • The NCF basalt/epoxy composite considered in this study has de­
monstrated superior flexural and interlaminar shear strength com­
Interlaminar shear and flexural properties have been evaluated in pared to results reported for plain-woven basalt/epoxy and plain-
the current study and failure modes have been analyzed using optical woven E-glass/epoxy reported in the literature.
microscopy. • Under flexural loading, failure was dominated by damage occurring

7
I.R. Chowdhury, et al. Composite Structures 245 (2020) 112317

Fig. 11. High magnification image of Fig. 10(h) indicating failure modes in
compression under flexural loading: (i) kink band formation in 0° fibers, (ii)
intra-ply delamination, (iii) inter-ply delamination. Ply boundaries are in­
dicated with white dashed lines.

Fig. 12. High-magnification image of Fig. 10(g) indicating failure modes in


tension under flexural loading: (i) transverse crack along 90° fibers, (ii) intra-
ply delamination, (iii) fiber breakage. Ply boundaries are indicated with white
dashed lines.

on the compression side of the sample. The dominant failure mode


was fiber kinking in the 0° fibers on the compression side and fiber
breakage in the 0° fibers on the tension side. Compression damage
was observed to initiate between 20 and 50% of peak load, mainly
in the form of kink band formation and delamination. Damage
progressed from ply-to-ply as the load increased. The tension side of
the flexural specimen exhibited no significant damage up to 80% of
the peak load.
• Interlaminar shear cracking was observed at mid thickness some­
where between 50 and 80% of the peak load. The 80% load level
also corresponded with the onset of non-linearity in the stress-dis­ Fig. 13. Cross-section of composite tested under interlaminar shear at the mid-
thickness region of the specimen: (a) 50% of peak stress, (b) 80% of peak stress,
placement curve indicating significant damage and redistribution of
(c) loaded to failure. Ply boundaries indicated with white dashed lines. 8th and
stresses had occurred by 80% of failure load. Damage did not occur
9th plies (of 16 in total) indicated with a white star. The zoomed-in figures
on either the tension or compression side until after 80% of peak gives more insight to the interlaminar shear crack development.
load indicating that interlaminar shear stresses dominated the
failure process. It was observed that the crack followed the fi­
ber–matrix interface of the 90° fiber tows. glass fabric may contribute to the lower strength exhibited by the

• NCF basalt/epoxy composites demonstrated inferior flexural and NCF basalt/epoxy composite.
interlaminar shear strength compared to NCF E-glass/epoxy com­
posites, reported in the literature [38]. It can also be noted that the Declaration of Competing Interest
failure of NCF basalt/epoxy composite under ILSS was by inter­
laminar shear cracking at the mid-thickness region. Whereas, NCF E- The authors declare that there is no conflict of interest.
glass/epoxy demonstrated buckling failure on the compression side
indicative of a strong interfacial strength [38]. Also, the lower
Acknowledgments
percentage of 0° fibers (approx. 85%) in the NCF basalt fabric
compared to the percentage of 0° fibers (approx. 90%) in NCF E-
The authors would like to acknowledge the support provided by the

8
I.R. Chowdhury, et al. Composite Structures 245 (2020) 112317

Fig. 14. Cross-section of composite tested under interlaminar shear showing tension side (left) and compression side (right) at mid-span of the specimen: (a, b) 50%
of peak stress, (c, d) 80% of peak stress and (e, f) loaded to failure. Ply boundaries are indicated with white dashed lines. The topmost ply on the compression side and
bottom ply on the tension side are indicated with a white star.

ESB (Electricity Supply Board) and the Faculty of Science and [3] Ocean Energy Forum, Ocean energy strategic roadmap 2016, building ocean energy
Engineering at the University of Limerick for this project. The authors for Europe; 2016.
[4] Magagna D, Shortall R, Telsnig T, Uihlein A, Vasquez Hernandez C. Supply chain of
would also like to thank Anthony O’Carroll and Adrian McEvoy for renewable energy technologies in Europe - An analysis for wind, geothermal and
facilitating training and access to the manufacturing and testing facil­ ocean energy, EUR 28831 EN, Publications Office of the European Union,
ities utilized for this study. Luxembourg, 2017, ISBN 978-92-79-74281-1, DOI:10.2760/271949, JRC108106.
[5] Fagan E. Design of fiber-reinforced polymer composite blades for wind and tidal
turbines Ph.D. thesis Galway: National University of Ireland; 2017.
Appendix A. Supplementary data [6] Jespersen KM, Glud JA, Zangenberg J, Hosoi A, Kawada H, Mikkelsen LP.
Uncovering the fatigue damage initiation and progression in uni-directional non-
crimp fabric reinforced polyester composite. Compos A 2018;109:481–97.
Supplementary data to this article can be found online at https://
[7] Thomsen OT. Sandwich materials for wind turbine blades – Present and future,
doi.org/10.1016/j.compstruct.2020.112317. journal of sandwich structures and materials, vol. 11, ; 2009. pp. 7–26.
[8] Nijssen RPL, Brøndsted P. Fatigue as a design driver for composite wind turbine
blades in Adv Wind Turb Blade. Des Mater 2013. https://doi.org/10.1533/
References
9780857097286.2.175.
[9] Røndsted P, Lilholt H, Lystrup HA. Composite materials for wind power turbine
[1] Chong WT, Muzammil WK, Wong KH, Wang CT, Gwani M, Chu YJ, et al. Cross axis blades. Annu Rev Mater Res 2006;35:505–38. https://doi.org/10.1146/annurev.
wind turbine: pushing the limit of wind turbine technology with complementary matsci.35.100303.110641.
design. Appl Energy 2017;207:78–95. [10] Nijssen RPL. Fatigue life prediction and strength degradation of wind turbine rotor
[2] GWEC, Global Wind Report. Annual Market Update 2017, Glob. Wind Energy blade composites PhD. Thesis Delft University of Technology; 2006.
Counc., ; 2017. p. 72. [11] Summerscales J, Searle TJ. Low-pressure (vacuum infusion) techniques for molding

9
I.R. Chowdhury, et al. Composite Structures 245 (2020) 112317

large composite structures. Proc Inst Mech Eng Part L: J Mater: Des Appl [37] Davies P, Verbouwe W. Evaluation of basalt fiber composites for material appli­
2005;219:45–58. cations. Appl Compos Mater 2018(25):299–308.
[12] Overview An, Mishnaevsky Jr. L, Branner K, Petersen HN, Beauson J, McGugan M, [38] Nash NH, Portela A, Bachour C, Manolakis I, Comer AJ. Effect of environmental
Sorensen BF. Materials for wind turbine blades. Materials 2017;10:1–24. conditioning on the properties of thermosetting- and thermoplastic-matrix com­
[13] Fecko D. High strength glass reinforcements still being discovered. Reinf Plast posite materials by resin infusion for marine applications. Compos Part B 2019.
2006;50:40–4. https://doi.org/10.1016/j.compositesb.2019.107271.
[14] Carbon Fiber vs. Fiberglass: A Comparison between the Two Materials Which [39] Nash NH, Portela A, Bachour C, Manolakis I, Comer AJ. Effect of Environmental
Material Is Superior? Available online: https://infogr.am/carbon-fiber-vs-fiberglass Conditioning on the Flexural Properties of Thermosetting- and Thermoplastic-
[accessed on 11 August 2019]. Matrix Composite Materials by Resin Infusion for Marine Applications, In pre­
[15] Grande JA. Wind Power Blades Energize Composites Manufacturing, Plastics paration, 2019.
Technology, 2008. [Online]. Available: https://www.ptonline.com/articles/wind- [40] Basaltex, “Technical Data Sheet: Basalt Multiaxial Fabric,” vol. 32, p. 8560, 2017.
power-blades-energize-composites-manufacturing. [Accessed; 17-Mar-2019]. [41] Tanaka K, Tokura D, Katayama T. Effect of stitch tension of non-crimp fabric on the
[16] Haberkern H. Tailor-made reinforcements. Reinf Plast 2006;50(4):28–33. https:// mechanical properties of 331 CFRTP, in Recent Advances in Structural Integrity
doi.org/10.1016/S0034-3617(06)70974-2. Analysis: Proceedings of the International Congress (APCF/SIF-2014), 2014, pp.
[17] Monaldo E, Nerilli F, Vairo G. Basalt-based fiber-reinforced materials and structural 331–335 (615).
applications in civil engineering. Compos Struct 2019;214:246–63. [42] Gurit, Full General Datasheet: PrimeTM27 epoxy infusion system; 2018. pp. 1–8.
[18] Fiore V, Scalici T, Di Bella G, Valenza A. A review on basalt fiber and its composites. [43] ASTM D3171-15, Standard test methods for constituent content of composite ma­
Compos Part B Eng 2015;74:74–94. terials, West Conshohocken, PA: the United States, ASTM International; 2015. pp.
[19] Ralph C, Lemoine P, Summerscales J, Archer E, Mcllhagger A. Relationship among 1–11.
the chemical, mechanical and geometrical properties of basalt fibres. Text Res J [44] Turi E. Thermal characterization of polymeric materials. PA: New York: Academic
2019;89(15):3056–66. Press; 1981.
[20] Wei B, Cao H, Song S. Tensile behavior contrast of basalt and glass fibers after [45] ASTM D7264/D7264M – 07, Standard test method for flexural properties of
chemical treatment. Mater Des 2010;31(9):4244–50. polymer matrix composite materials, West Conshohocken, PA: the United States,
[21] Nasir V, Karimipour H, Taheri-Behrooz F, Shokrieh MM. Corrosion behavior and ASTM International; 2007. pp. 1–11.
crack formation mechanism of basalt fiber in sulphuric acid. Corros Sci [46] ASTM D2344/D2344M, Standard test method for short-beam strength of polymer
2012;64:1–7. matrix composite materials and their laminates, West Conshohocken, PA: the
[22] Mengal AN, Karuppanan S, Wahab AA. Basalt carbon hybrid composite for wind United States, ASTM International; 2016. pp. 1–8.
turbine rotor blades: a short review. Adv Mater Res 2014;970:67–73. [47] Thomsen OT, Kratmann KK. Experimental characterisation of parameters control­
[23] Abashidze S, Marquis FD, Abashidze GS. Hybrid fiber, and nanopowder reinforced ling the compressive failure of pultruded unidirectional carbon fibre composites.
composite for wind turbine blades. J Mater Res Technol 2015;4:60–7. Appl. Mech. Mater. 2010;24–25:15–22.
[24] Soukhanov AV, Dalinkevich AA, Gumargalieva KZ, Marakhovsky SS. Modern basalt [48] Budiansky B, Fleck NA. Compressive kinking of fiber composites : A topical review.
fibers, and epoxy basalt to plastics: properties and applications. Key Engineering Appl. Mech. 2016;47(6):246–50.
Materials. Volume II: Interdisciplinary concepts and research. New York: CRC Press, [49] Costa S. Physically based constitutive models for carsh of composites, Ph.D Thesis
Taylor & Francis Group; 2014. p. 69–94. Chalmers University of Technology; 2019.
[25] Sim J, Park C, Moon DY. Characteristics of basalt fiber as a strengthening material [50] Huchette C, Vandellos T, Laurin F. Influence of intralaminar damage on the dela­
for concrete structures. Compos. Part B Eng 2005;36(6–7):504–12. mination crack evolution. Damage growth in aerospace composites. Switzerland:
[26] Ralph C, Lemoine P, Boyd A, Archer E, Mcllhagger A. The effect of fibre sizing on Springer International Publishing; 2015. p. 1–34.
the modification of basalt fibre surface in preparation for bonding to polypropylene. [51] Adams D. Can flexure testing provide estimates of composite strength properties?”
Appl Surf Sci 2019;475:435–45. Composites World, 2017. [Online]. Available: https://www.compositesworld.com/
[27] Ary Subagia IDG, Tijing LD, Kim Y, Kim CS, Vista IV FP, Shon HK. Mechanical articles/can-flexure-testing-provide -estimates-of-composite-strength-properties.
performance of multiscale basalt fiber–epoxy laminates containing tourmaline [Accessed: 15-Jan-2019].
micro/nanoparticles. Compos Part B 2014;58:611–7. [52] Colombo C, Vergani L, Burman M. Static and fatigue characterization of new basalt
[28] Bulut M. Mechanical characterization of Basalt/epoxy composite laminates con­ fiber reinforced composites. Compos Struct 2012;94(3):1165–74.
taining graphene nanopellets. Compos Part B 2017;122:71–8. [53] Landucci G, Rossi F, Nicolella C, Zanelli S. Design and testing of innovative mate­
[29] Petrucci R, Santulli C, Puglia D, Sarasini F, Torre L, Kenny JM. Mechanical char­ rials for passive fire protection. Fire Saf J 2009;44:1103–9.
acterization of hybrid composite laminates based on basalt fibers in combination [54] Jamshaid H, Mishra R. A green material from rock: basalt fiber – a review. J Textile
with flax, hemp and glass fibers manufactured by vacuum infusion. Mater Des Inst 2015:1–15.
2013;49:728–35. [55] Liu Q, Shaw MT, Parnas RS. Investigation of basalt fiber composite mechanical
[30] Sun G, Tong S, Chen D, Gong Z, Li Q. Mechanical properties of hybrid composites properties for applications in transportation. Polym Compos 2006:41–8.
reinforced by carbon and basalt fibers. Int J Mech Sci 2018;148:636–51. [56] Loos M. Composites in Carbon nanotube reinforced composites (CNR Polymer
[31] Sarasini F, Tirillo J, Ferrante L, Valente M, Valente T, Lampani L, et al. Drop-weight Science and Technology). Willian Andrew Applied Science Publishers; 2015. p.
impact behavior of woven hybrid basalt-carbon/epoxy composites. Compos Part: B 37–72.
2014;59:204–20. [57] Mahltig B, Kyosev Y. Basalt fibers in Inorganic and composite fibers; production,
[32] Lopresto V, Leone C, Iorio ID. Mechanical characterization of basalt fiber reinforced properties and applications. Woodhead Publishing; 2018. p. 195–217.
plastic. Compos Part B 2011;42:717–23. [58] Martynova E, Cebulla H. Glass fibers in Inorganic and composite fibers; production,
[33] Scalici T, Pitarresi G, Badagliacco D, Fiore V, Valenza A. Mechanical properties of properties and applications. Woodhead Publishing; 2018. p. 131–63.
basalt fiber reinforced composites manufactured with different vacuum-assisted [59] Tempelman E, Shercliff H, van Eyben BN. Resin Transfer Molding. Manufacturing
impregnation techniques. Compos Part B 2016;104:35–43. and design: understanding the principles of how things are made. Elsevier Ltd.;
[34] Chairman CA, Babu SPK. Mechanical and abrasive wear behavior of glass and ba­ 2014. p. 171–86.
saltfabric-reinforced epoxy composites. J Appl Polym Sci 2013. https://doi.org/10. [60] Wypych G. Fillers − origin, chemical composition, properties, and morphology in
1002/APP.39154. handbook of fillers. Chem Tec Publishing; 2016. p. 14–266.
[35] Dorigato A, Pegoretti A. Flexural and impact behavior of carbon/basalt fibers hy­ [61] Bhat T, Fortomaris D, Kandare E, Mouritz AP. Properties of thermally recycled
brid laminates. J Compos Mater 2014;48(9):1121–30. basalt fibers and basalt fiber composites. J Mater Sci 2018;53(3):1933–44.
[36] Carmisciano S, Rosa IMD, Sarasini F, Tamburrano A, Valente M. Basalt woven fiber [62] Liu Q, Shaw MT, Parnas RS, McDonnell AM. Investigation of basalt fiber composite
reinforced vinyl ester composites: Flexural and electrical properties. Mater Des Mechanical properties for applications in transportation. Polym Compos
2011;32:337–42. 2006;27:478–83.

10

You might also like