Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Composites Part B 171 (2019) 310–319

Contents lists available at ScienceDirect

Composites Part B
journal homepage: www.elsevier.com/locate/compositesb

A Split-Hopkinson Tension Bar study on the dynamic strength of


basalt-fibre composites
G.C. Ganzenmüller a, b, *, D. Plappert b, A. Trippel b, S. Hiermaier a, b
a
Fraunhofer Ernst-Mach-Institute for High-Speed Dynamics, EMI, Freiburg i. Br, Germany
b
Albert-Ludwigs Universit€
at Freiburg, Institute for Sustainable Systems Engineering, Freiburg i. Br, Germany

A R T I C L E I N F O A B S T R A C T

Keywords: This paper investigates the strain rate sensitivity of laminated composites made of plies of unidirectional basalt
Mechanical testing fibres and epoxy resin. We consider laminates with quasi-isotropic [0∘, 45∘,þ45∘,90∘]s and orthogonal
Strength [þ45∘, 45∘]4 layup. A Split-Hopkinson Tension Bar is used to generate accurate stress/strain data at elevated
Basalt fibre
rates of strain of �3 � 102/s. Moderate strain rate effects are observed with strength increase of �3.5% per
decade of increased loading rate for both laminate types.

1. Introduction only exhibit strength increases of approximately 1%–10% per order of


magnitude of increased strain rate [8–14]. In another study, the strain
The purpose of this work is to present reasonably accurate dynamic rate sensitivity of woven basalt/epoxy composites was studied, and a
tensile strength data for composites made of plies of endless, unidirec­ similarly exceptional high strain rate sensitivity was observed [15]. Is
tional basalt fibres and epoxy resin. Over the last two decades, basalt there something special about basalt fibres which makes them very
fibres have come into consideration as potential reinforcement for strain rate sensitive? To shed some light on this issue, we investigate the
composite materials requiring high strength and temperature resistance. strain rate dependency of the tensile failure strength and strain for
In contrast to the infamous asbestos fibre, which meets these application quasi-isotropic [0∘, 45∘,þ45∘,90∘]s and orthogonal [þ45∘, 45∘]4 layups
requirements from a mechanical point of view, basalt is not considered of unidirectional basalt fibres in a thermoset epoxy matrix within the
carcinogenic [1]. It is a mineral belonging to the group of silicates and as strain rate regime 10 3 – 3 � 102/s. To obtain accurate data at high rates
such similar to glass in its chemical composition but with more iron and of strain, we employ a Split-Hopkinson Tension Bar, for reasons we feel
less calcium content. While glass fibres are completely amorphous important to briefly review in the following:
solids, basalt fibres can feature some degree of cristallinity [2]. Basalt’s Acquiring accurate stress/strain data at high rates of strain >100/s is
mechanical properties strength and stiffness are lower than those of not a trivial task, as has been discussed extensively in the literature [16,
carbon fibres, and more akin to those of S-2 glass fibres. The quasi-static 17]. The difficulty may be attributed to three issues: (i) The testing
properties of basalt fibres and its composites are already well known [3, apparatus must achieve a well defined, e.g., constant velocity at which
4]. A promising field of application for basalt fibre composites are the sample is loaded, typically on the order of a few m/s. This requires
structural elements in the automotive sector [5], where it is important to strong acceleration and good velocity control of the gripping devices
predict the dynamic material behaviour at elevated strain rates relevant used to interface the specimen with the testing apparatus. (ii) The stress
at crash, which are on the order of 10–100/s [6]. However, as of now, state within the specimen needs to be inferred from a force sensor, which
only a few studies are available on dynamic properties. This is particu­ is usually not mounted on the specimen itself but rather located in-line
larly true for composites composed of plies of endless, unidirectional with the loading axis at some distance from the specimen. During the
(UD) fibres. To our knowledge, only one, very recent study is available short time span of a dynamic experiment, propagation effects of elastic
[7], where the authors report that the dynamic tensile strength of a pure waves must be considered. The force sensor will, in general, not sense
0∘ basalt/epoxy composite doubles its strength when strain rate is the force experienced by the specimen at a given instant of time, but
increased from 10/s to 300/s. This amount of strain rate sensitivity instead report another value which is influenced by its own acceleration.
appears exceptionally high, as UD composites of carbon or glass fibres Furthermore, the dynamic bandwidth, i.e., its frequency response

* Corresponding author. Fraunhofer Ernst-Mach-Institute for High-Speed Dynamics, EMI, Freiburg i. Br, Germany.
E-mail address: georg.ganzenmueller@inatech.uni-freiburg.de (G.C. Ganzenmüller).

https://doi.org/10.1016/j.compositesb.2019.04.031
Received 22 January 2019; Received in revised form 2 April 2019; Accepted 26 April 2019
Available online 4 May 2019
1359-8368/© 2019 Elsevier Ltd. All rights reserved.
G.C. Ganzenmüller et al. Composites Part B 171 (2019) 310–319

function, must be adequate to accurately describe the specimen’s force stress/strain data for composites, see e.g. Refs. [20,21], and a strong
signal. Typically, dynamic bandwidths on the order of 1 MHz are theoretical basis is available which rationalizes these beliefs [22].
required for experiments at strain rates between 102 – 103/s [18]. (iii) In contrast, Split-Hopkinson Tension (SHTB) Bar experiments are
Determination of the specimen strain state requires fast and robust more difficult to perform. The specimen needs to be attached to the ends
measurement techniques. Fast, because for sampling the strain evolution of the input and output bars without introducing significant additional
of a specimen with failure strain of 5% at a strain rate of 100/s with only mass, e.g., specimen grips as used in conventional testing machines. The
100 data points, already an acquisition rate of 200 kHz is required. presence of such additional mass means additional changes in acoustic
Robust, because the sampling method must yield reliable measurements impedance, causing unwanted wave reflections. This violates the as­
for a rapidly accelerating and deforming specimen. While strain gauges sumptions made in the classic analysis based on simple wave propaga­
meet the requirement of fast acquisition rates, it can be difficult to tion in media with constant acoustic impedance. The situation is
maintain adhesion on the surface of a specimen until failure. Alterna­ fortunate for specimens with cylindrical shape, as these can be threaded
tively, high-speed imaging and optical strain analysis is a universal yet at the ends and screwed into matching interior threads in the bar ends.
expensive solution if high spatial resolution is required. Thus, no additional mass is introduced for affixing the specimen. Com­
For conducting experiments at high rates of strain, two popular posites, however, are typically tested in a planar strip form. Until now,
experimental methods are available. The first method is given by the the only solution has been has been to introduce slots either into the bar
universal testing machine, powered by a servo-hydraulic actuator. While ends or into adapters, and bond the specimens into these slots with high-
these machines can operate over large ranges of testing velocities and strength adhesives [9,12,23]. This process requires intensive labour and
forces with typical upper limits of 10 m/s and 250 kN, difficulties arise expensive machining. Recently, two of the current authors introduced a
for accurately measuring the force signal of a specimen under rapid clamping device which allows reversible mounting of strip-type speci­
loading. Taking reference to the points (i) and (ii) from above, this is due mens without violating the requirement of constant acoustic impedance
to the following reasons: (I) The apparatus needs to achieve a high [24]. Here, we utilize this clamping device to obtain SHTB data of
testing velocity before the specimen is loaded. This is realized via a lost reasonable accuracy and compare this data with results from quasi-static
motion device, which lets the hydraulic piston travel freely under ac­ and moderate strain rates.
celeration. Once the piston is at speed, its movement is coupled The remainder of this article is organized as follows: The composite
instantaneously to the specimen’s grips. The instantaneous acceleration investigated here and the experimental methods employed are described
of the – typically heavy – specimen grips causes a strong inertial force in the next section. We then report our findings for strength and failure
response which overlays the true force signal of the specimen’s stress strains, analyse the strain rate dependency and conclude with a critical
response. This effect may be theoretically visualized by considering the review of our data in the light of other researchers’ findings.
frequency response of a velocity signal with a step-like acceleration. Its
Fourier transform shows amplitudes over an infinite frequency spec­ 2. Materials and methods
trum. (II) Force measurement is conventionally performed via a piezo-
electric load cell with finite stiffness and its own limited frequency 2.1. Material specimens
response. Such force sensors act as low-pass filters. The fact that the
quasi-infinite frequency spectrum associated with the step loading Sheets of cured UD basalt fibre composites with layups
character of the slack adapter is not present in the output signal is caused [0∘, 45∘,þ45∘,90∘]s and orthogonal [þ45∘, 45∘]4 were provided tested.
by this filtering effect. While servo-hydraulic universal testing machines This composite is made from a prepreg of basalt fibres with filament
have their warranted uses for specific applications, acquiring high- diameter 17 μm and a low-viscosity epoxy resin system. Laminates were
quality stress-strain data, in particular for lightweight, non-metallic cured in an autoclave process resulting in fibre volume fraction of 60%.
materials, is challenging. Naively applying such a machine at high To illustrate the mechanical properties of this material, we quote the
rates of strain, one does not measure the true specimen response but nominal strength and stiffness of pure 0∘ specimens according to ISO-527
instead a convolution of the specimen response with the acceleration as 1310 MPa and 44 GPa [25].
response of the specimen grips and elastic wave reflection effects with Testing specimens were cut from the sheets using a diamond coated
visible oscillatory artifacts, see e.g. Ref. [19]. However, it needs to be saw blade to obtain good cut surface quality. The dimensions of the
pointed out that high-quality data may also be obtained using this quasi-isotropic (QI) specimens are length � width � height ¼
approach if its limitations are considered carefully [17]. 100 � 3 � 1.6 mm3 with a nominal gauge length of 20 mm. The di­
A more straightforward alternative for high rate testing is given by mensions of the �45∘ specimens are 100 � 5 � 2.0 mm3 with the same
the Split-Hopkinson Bar method, also known as the Kolsky Bar method. nominal gauge length. The same specimen geometry was used across all
It is out of the scope of the present work to detail this technique strain rates. Specimens were bonded to grooved 7075 aluminium tabs
extensively, instead the reader is referred to Ref. [18]. Here, the spec­ with a high-strength, thermally activated epoxy resin according to
imen is sandwiched between two long and straight bars of constant cross Fig. 1. We use split tabs as these can be easily manufactured in large
section, the input and output bars. An elastic wave of well defined quantities using a CNC router. Additionally, the split tabs provide good
amplitude is created in the input bar, either by impacting the bar with accessibility to the interior which makes it trivial to correctly apply the
another long bar, the striker, or by pre-tensioning the bar and employing adhesive.
a release mechanism. The elastic wave then passes through the specimen The tabs feature holes for M4 through-bolts to affix the assembled
into the output bar, creating secondary transmitted and reflected waves specimen to the aluminium grips as shown in Fig. 2. The clamps are
at the interfaces between specimen and input/output bars. These waves designed to fulfill three task: (i) they provide an ISO M12 � 1.5 thread to
are measured using strain gauges mounted on the bars, and analysed interface with either the Split-Hopkinson bars or the universal testing
using the theory of one-dimensional elastic wave propagation to yield machine. (ii) The transition from the circular interfacing end with the
both specimen strain and stress. Due to the simple geometry, all 1D wave testing apparatus to the rectangular shape used to attach the specimen
effects are accounted for, and the resulting stress/strain curves can be of maintains a constant cross section area. For dynamic testing, the Split-
high quality. However, this fortunate situation applies only to Hopkinson bars are also made of aluminium, such that constant acous­
compression experiments performed with the Split-Hopkinson Pressure tic impedance is obtained. This avoids unwanted wave reflections due to
Bar (SHPB). Here, a small amount of grease is sufficient to stick a cy­ the specimen gripping device, see Ref. [24] for details. (iii) The ø4.1 mm
lindrical sample between the faces of input- and output bar, and no through holes allow for clamping the specimen inside the grip. To this
explicit specimen grips are required. Due to this simplicity, the SHPB end, lightweight M4 aluminium bolts with strength class ISO 5.6 may be
method is widely accepted for reliably producing accurate compression used, yielding an axial force of 1.8 kN per bolt. With 4 bolts and a

311
G.C. Ganzenmüller et al. Composites Part B 171 (2019) 310–319

Fig. 1. CAD rendering of the specimen with its split aluminium tabs. Also shown is the fixture which is used to align all parts during assembly and curing of the epoxy
adhesive. All dimensions in mm.

Fig. 2. Clamping fixture maintains cross section area and thus provides constant acoustic impedance which is required for reflection-free wave transmission. All
dimensions in mm.

312
G.C. Ganzenmüller et al. Composites Part B 171 (2019) 310–319

conservatively assumed friction coefficient μ ¼ 1.0 between the clean a Wheatstone bridge circuit to eliminate bending information. The
and unlubed aluminium surfaces, this translates into a useful upper force Wheatstone bridge circuit is driven in constant voltage mode and its
limit of �14 kN for this clamp. This limit exceeds the expected forces in output is increased by a factor of 100 using an amplifier with 1 MHz
this work by a factor of 4. We note that those parts of the through-bolts bandwidth. This signal is recorded by a data acquisition (DAQ) card
and nuts protruding from the surfaces introduce additional mass, which operating at 10 MHz and 16 bit resolution. The conversion factor from
causes unwanted wave reflections. However, this effect is so minor that strain to force is established via a calibration procedure, wherein a
its effects cannot be observed in our data. dedicated force sensor is placed between input- and output bars, while a
static load is applied onto the bars. The signal of the force sensor is
2.2. Quasi-static and low strain rate testing recorded using the same amplifier and data acquisition card as during a
real experiment. The force sensor, in turn is calibrated against the
For the quasi-static and low strain rates of 10 3/s and 10 1/s, we use already calibrated load cell in the universal testing machine described
a screw-driven universal testing machine with 100 kN capacity and a above. Thus, the entire force measuring system consisting of strain
maximum testing velocity of 1500 mm/min. Specimens were attached gauges, amplifier and DAQ card is calibrated, which compensates for
using the gripping device described in section 2.1, which itself was any eventual misalignment of the strain gauges or similar constant
mounted to the testing machine using cardanic joints to prevent off-axis systematic errors.
loading. The testing velocity was adjusted in pilot experiments such that In addition to the line-scan camera, we also use a high-speed area
the target strain rates were obtained. Force was measured via the ma­ camera to obtain images of the deformation process. The highest sam­
chine’s internal load sensor, which is calibrated to accuracy class 1/ISO pling rate of this camera is 38.65 kHz at a resolution of 336 � 96 pixels.
7500–1. Strain was measured optically using a black/white camera with This is not enough for accurate strain determination, but serves well to
resolution of 1280 � 1024 pixels and a maximum frame rate of 1500 fps obtain a qualitative impression of the failure behaviour, see Appendix A.
at this resolution. To this end, white contrast marks were painted on the Both line-scan and area cameras are triggered by the DAQ card to
specimen at the edges on the gauge section with a paint marker. The start recording when the incident wave created by the striker reaches
displacement of the contrast marks was recorded and converted into a strain gauge no. 1, c.f. Fig. 3. In the case of the line-scan camera, a
strain time series using a motion analysis software based on Digital synchronization signal is fed back to the DAQ card, so precise timing
Image Correlation (DIC). Both force and strain time series were com­ information is available to correlate strain and force signals. It is
bined into nominal stress/strain graphs by eliminating the time infor­ important to realize that even though strain and force signals are
mation. We note that for this universal testing machine in combination perfectly well synchronized in this manner, these signals are measured
with the tested specimens, a strain rate of 10 1/s is the most dynamic in different locations: strain is measured on the specimen but force is
experiment that can be achieved before acceleration effects cause inferred from the strain gages on the output bar, in our case 200 mm
ringing artifacts in the force signal can be observed. upstream in the transmitted elastic wave direction. The time taken by
the elastic wave to travel this distance must be accounted for by shifting
the force signal by Δt ¼ 200 mm/5090 mm/ms ¼ 0.039 ms, where
2.3. Dynamic testing apparatus
5090 mm/ms is the longitudinal wave speed in aluminium.
In classic Split-Hopkinson Bar experiments, a check of dynamic
The Split-Hopkinson Tension Bar (SHTB) used here is sketched in
equilibrium is possible by placing strain gages symmetrically around the
Fig. 3 and described in detail in Ref. [26]. Compared to other SHTBs, this
specimen, i.e., also on the input bar near the specimen. This allows to
setup is optimized for low velocities, low forces and a long pulse dura­
check when the forces on both sides of the specimen are equal in
tion of 1.2 ms. Specimens were attached to the bars using the grips
magnitude, which marks the beginning of measured data validity and
detailed in section 2.1. Shortly before performing the experiment, white
constant strain rate. Here, we cannot provide this information as no
contrast marks were painted on the specimen at the edges of the gauge
additional strain gauge was mounted on the input bar. However, the
section using a paint marker. We have observed that it is crucial for the
onset of dynamic equilibrium may also be calculated analytically, see e.
paint to be still compliant. If it is fully dried it will come off the specimen
g. Ref. [22]. Usually, the time taken for the elastic wave to perform 4 to 5
before actual failure occurs, thus voiding the strain measurement. For
round-trips within the specimen is sufficient [28,29]. In case of the
determining the strain, we employ a line-scan camera to track the
quasi-isotropic specimen, the longitudinal wave speed
displacement of the contrast marks. If only uniaxial displacement is of
is � 3300 mm/ms, and the gauge length is 20 mm, resulting in an
interest, line-scan cameras are advantageous over area cameras due to
equilibration time of <60 μs.
higher 1D resolution and increased light sensitivity due to larger pixel
Our SHTB experiments were set up such that the incident wave at­
size. We employ a model with 1 � 4096 pixel resolution and a line scan
tains a force amplitude of 20 kN, corresponding to a particle velocity of
frequency of 200 kHz. The strain is ultimately obtained by
7.2 m/s. This marks the upper end of what is safely attainable with our
post-processing the line-scan data using a pattern matching algorithm
aluminium setup due to its mechanical strength. Fig. 4 shows the
with sub-pixel accuracy [27]. Force is measured via a pair of conven­
recording of such an experiment on a quasi-isotropic specimen. The
tional strain gauges mounted on the output bar, connected diagonally in

Figure 3. Sketch of the SHTB setup employed in this work. All dimensions in mm. Input and output bars are 16 mm diameter aluminium rods. The striker is a hollow
aluminium tube of 40 mm outer diameter and 20 mm inner diameter. Two strain gauge stations on the input bar and output bar measure the incident wave, εinc, and
transmitted wave, εtra.

313
G.C. Ganzenmüller et al. Composites Part B 171 (2019) 310–319

Figure 5. Force and strain time series for a single Split-Hopkinson Tension Bar
Figure 4. Bottom plot: Recordings from a Split-Hopkinson Tension Bar experiment on a quasi-isotropic specimen at strain rate 300/s. Force data (black
experiment on a quasi-isotropic specimen. Both incident and transmitted waves solid line) is sampled at 10 MHz, local specimen strain (red circular symbols) is
are direct recordings from the strain gauges at the locations described in Fig. 3, sampled at 200 kHz using a line scan camera. The solid red line is a linear
scaled to units of force. The reflected wave is computed from the expression regression to the local strain data points with slope 0.295/ms, corresponding to
εref ¼ εinc εtra . Top plot: strain rate as obtained from the reflected wave and its the target strain rate of � 300/s. The classic, but erroneous, Split-Hopkinson
time integral, which is the specimen strain. estimate of the specimen strain as computed from the reflected wave is
also shown.

incident wave has a pronounced rise time of �0.2 ms, which is due to the
Fig. 5 compares the local specimen strain measured using the line
use of a pulse shaper between striker and transfer flange, a rubber disk of
scan camera with the strain computed from the reflected wave. The
1 mm thickness. This pulse shaper significantly reduces oscillations in
locally measured strain is 4.9% and thus approximately only 65% of the
the transmitted force signal, which would otherwise completely domi­
classic estimate. Additionally the local strain rate as obtained from a
nate. However, the pulse shaper also implies that the strain rate is not
linear fit to the specimen strain is 300 � 20/s, which is half of what is
constant at the start of loading but instead slowly ramps up. In this case,
estimated by the classic approach. We would like to emphasize that this
the pulse shape is chosen such that a nearly constant strain rate is ob­
difference is not due to the two-wave approximation, but instead caused
tained before failure. The strain rate may be obtained in two different
by the compliance of the specimen holding fixture.
ways, either directly from the strain gauge signals or by measuring
Our SHTB setup with its long pulse duration allows to reach rela­
locally on the specimen using an independent method. In the following,
tively low strain rates compared to what is common in the SHTB com­
we will compare these two approaches and argue why it is important for
munity. In principle, we could have realized even lower strain rates for
the SHTB to only use the independent local measurement.
same-sized specimens: in the quasi-isotropic case, only 0.25 ms of the
As a result of one-dimensional wave propagation theory, the classi­
available pulse duration of 1.2 ms is used, and for �45∘ specimens,
cally used strain rate is proportional to the reflected wave originating at
failure is reached at approx 0.55 ms. However, such experiments would
the interface between input bar and specimen [18],
deliver little additional insight as strain rate effects typically depend
2 c0 εref logarithmically on the strain rate. It was therefore our aim to achieve the
ε_ classic ¼ ; (1)
L0 highest possible strain rate. The quality of our data compares well with
other Split-Hopkinson studies [8,9,12,23], and we feel confident that
where c0 is the wave speed in the bar and L0 is the initial separation of quantitative conclusions can be drawn from these data.
input and output bar, which is equal to the specimen length. This strain
rate thus measures the relative movement of the specimen-bar in­ 3. Results and discussion
terfaces. With our setup, we cannot measure the reflected wave εref
directly, as the geometry does not allow an isolated reading of the re­ 3.1. Data reduction
flected pulse because the striker has a larger generalized impedance than
the input bar which causes superpositions of waves. Instead, we Fig. 6 shows the stress/strain curves obtained at each combination of
approximate the reflected wave by assuming the condition of dynamic strain rate and laminate layup. Each testing series consists of N ¼ 5 valid
equilibrium, also referred to as the two-wave approximation: experiments. An experiment is considered valid, if failure occurred
εref ¼ εinc εtra (2) within the gauge section, i.e., between the holding tabs. Within one
series, the individual experiments exhibit scatter which necessitates an
The strain rate computed in this fashion, along with the resulting averaging procedure before multiple testing series can be compared to
strain is depicted in the upper part of Fig. 4. This strain rate attains a each other. The characteristic properties maximum stress, σ max and
value of ε_classic�600/s near failure and the failure strain is 7.6%. It is strain at maximum stress, εmax are represented using straightforward
known that this approach works well in the case of a pressure bar, but we
averages and uncertainty estimates, e.g. for the stress: σmax ¼
claim that is not apt in the case of the SHTB, where the specimen PN qffiP
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
N
mounting introduces additional compliance. In the present case, the j¼1 σ max;j =N and sðσ max Þ ¼ j¼1 ðσ max;j σmax Þ2 =ðN 1Þ.
specimen is glued into a holder, and the deformation of the adhesive In contrast, computing an average stress/strain curve for an entire
contributes erroneously to the nominal strain measured using the re­ testing series is not as straightforward. Within each testing series, the
flected wave. It is therefore necessary to employ a local measurement of sequences of strain values εi are not the same for all experiments; they
specimen strain.

314
G.C. Ganzenmüller et al. Composites Part B 171 (2019) 310–319

Figure 6. Stress/strain curves for quasi-isotropic and �45∘ specimens at different strain rates. Black lines show individual experiments and red lines represent
average behaviour obtained with the histogram approach (see text). Error bars indicate the standard deviation of the average. An artificial steep drop in the average
curves indicates our definition of failure strain, which is the strain at maximum stress.

Fig. 7. Comparison of average stress/strain data for quasi-isotropic laminates Fig. 8. Comparison of average stress/strain data for �45∘ laminates tested at
tested at different strain rates. Errorbars indicate the standard deviation and are different strain rates. Errorbars indicate the standard deviation and are only
only shown for the highest strain rate results to avoid clutter. shown for the highest strain rate results to avoid clutter.

span different ranges and are sampled at different locations. Thus, a


naïve averaging process over j ¼ 1.N datasets, such as σðεi Þ ¼
PN
j¼1 σ j ðεi Þ=N, cannot work. Instead, we employ a histogram-based
Table 1
approach. First, a strain axis which extends from the smallest observed Average stress and strain values for quasi-isotropic and �45∘ laminates, for the
strain value (typically 0) to the average strain at maximum stress is different strain rates considered in this work. The discrete standard deviation is
defined and discretized with sufficiently high resolution – 1000 points in used as uncertainty estimate following the � symbol. Note that for the �45∘
our case. Each individual stress/strain dataset is then interpolated using laminate, maximum strain is defined as the strain at maximum stress.
linear functions and sampled at every histogram bin which is covered by strain rate 10 3/s 10 1/s 3 � 102/s
the dataset. Now, each histogram bin contains between 1 and N samples,
QI max. strain [%] 3.3 � 0.4 3.9 � 0.2 4.8 � 0.1
which allows the definition of an average and an uncertainty estimate, QI max. stress [MPa] 428 � 43 452 � 42 524 � 38
for which we use the discrete standard deviation. �45∘ max. strain [%] 15.0 � 1.4 14.0 � 1.2 13.9 � 0.8
�45∘ max. stress [MPa] 179 � 6 194 � 7 217 � 7

315
G.C. Ganzenmüller et al. Composites Part B 171 (2019) 310–319

3.2. Quasi-isotropic specimens

Stress/strain curves for the quasi-isotropic specimens are shown in


Fig. 6, along with their representative average curves. We note that the
results obtained with the universal testing machine show some un­
dulations at strains <1%, which is due to initial movement of the car­
danic joints. Nevertheless, we are confident that the average curves are
reliable. The material exhibits almost linear-elastic, brittle character.
Maximum stress and strain at maximum stress coincide with the failure
stress and strain. Comparison among the different strain rates shows that
the failure stress and increases with strain rate from approximately 428
to 524 MPa, i.e. by 22 � 8%, see Table 1. To illustrate that the
comparatively small specimen size used here allows representative re­
sults to be obtained, we note that the corresponding quasi-static strength
according to ISO-527 using much larger, 25 mm wide specimens is
436 � 10 MPa [25]. Failure strain increases significantly from approxi­
mately 3.3%–4.8%. Classic visco-elastic behaviour, i.e., increased stiff­
ness and decreased failure strain with strain rate, behaviour is not
Fig. 9. Strain rate sensitivity analysis of failure stress for quasi-isotropic (solid observed. Instead, and according to Fig. 7, all curves are superimposed
circles) and �45∘ specimens (crosses). Error bars denote one standard deviation.
but extend to higher stress and strain as the strain rate is increased.
Straight lines are fits using a sensitivity model which is proportional to the
Within the uncertainty estimate, no distinction in stiffness can be
logarithm of the strain rate, see text.
observed, see Fig. 10. The location at which failure occurred exhibited
little systematic behaviour, independent of strain rate. Experiments

Fig. 10. Photographs of representative specimens after testing. The upper part shows the �45∘ specimens with zoomed-in regions for the two different strain rates
10 3/s and 300/s. The lower part shows the same for quasi-isotropic specimens. The edge length of a grey square on the background paper is 10 mm.

316
G.C. Ganzenmüller et al. Composites Part B 171 (2019) 310–319

were only considered valid, if failure occurred in the gauge region. The arguments may be misleading, as the composite’s strength is also a
failure mode could not be identified as the specimens disintegrated function of the fibre-matrix interface strength. To the knowledge of the
instantaneously at failure due to elastic energy release. This process was authors, no published information is available about the loading rate
too fast to be captured even with the high-speed camera at a frame rate sensitivity of interface properties in the case of basalt fibres and epoxy
of 38.65 kHz, see Fig. A11. resins.

3.3. �45∘ specimens 4. Discussion and conclusions

The �45∘ specimens exhibit classic strain rate dependent pseudo- This work reports the strain rate dependency of failure stress and
viscoplastic behaviour, see Fig. 8 Following an initial linear response, strain for composites made of basalt fibres and epoxy resin, produced as
which is the same at all strain rates within the uncertainty estimate, a laminates with quasi-isotropic [0∘, 45∘,þ45∘,90∘]S and orthogonal
rounded transition to an inelastic regime with constant strain hardening [þ45∘, 45∘]4 layups. Three different strain rates in the range 10 3 –
is observed. According to Table 1 the stress level of the inelastic regime 3 � 102 are considered. Data at the highest strain rate is obtained using a
increases significantly with strain rate: the maximum stress increases Split-Hopkinson Tension Bar. Our data is of sufficient quality to conduct
from approximately 179 to 217 MPa, i.e. by 21 � 4%. The corresponding a meaningful strain rate sensitivity analysis of strength, i.e., our mea­
quasi-static results for conventionally-sized specimens of 25 mm width surements indicate significant changes in strength – relative to the ex­
according to DIN ISO-527 is 203 � 8 MPa [25]. This differs slightly but periments’ standard deviation – as a function of strain rate. We quantify
finite-size effects are expected especially for this specimen type as the the strain rate sensitivity of strength as approximately 3.5% per order of
amount of shear deformation that can be sustained is proportional to the magnitude of loading rate, both for quasi-isotropic and �45∘ specimens.
specimen width. Failure strain (strain at maximum stress) is approxi­ This is in agreement with existing high-quality studies on comparable
mately 14% and thus much higher than in the quasi-isotropic case. The carbon and glass fibre epoxy composites in tension [8–13] and also
failure strain appears to decrease slightly with increased strain rate, similar to what is observed in compression [14,20,21]. Our results
although this observation cannot be fully justified given the uncertainty contradict the findings of a recent study on a similar basalt/epoxy
estimate. No qualitative distinction could be made in failure behaviour composite [7], who report a much more pronounced strain rate sensi­
at different strain rates. Failure occurred typically in the center of the tivity. In their study, a doubling of strength for a strain rate increase
gauge region between the holding tabs, and was less explosive in char­ from 10/s to 300/s was measured using a servo-hydraulic universal
acter compared to the quasi-isotropic specimens, see Fig. 10. testing machine. We argue that the Split-Hopkinson Tension Bar should
be used as the standard method for testing dynamic behaviour of com­
3.4. Analysis of strain-rate effects posites: Whilst servo-hydraulic machines are detrimentally affected by
pronounced wave propagation effects on short time scales, the Hop­
We analyse the strain rate dependency by plotting failure stress kinson Bar method incorporates this very phenomenon as the basis for
against the decadic logarithm of strain rate, see Fig. 9. Both quasi- its measuring principle. To facilitate the use of the Split-Hopkinson
isotropic and �45∘ datasets are well described with a simple model, Tension Bar for composites, we introduce a novel system for easily
which is proportional to the logarithm of the strain rate, i.e., mounting specimens of strip shape to the Hopkinson bars. This new
� � �� mounting system provides constant acoustic impedance and does not
ε_
σ fail ¼ A � 1 þ B log : hinder wave propagation. Our results show that the strain rate sensi­
ε_ 0
tivity is nearly identical for quasi-isotropic and �45∘ laminates. This is
Here, ε_0 ¼ 1/s is a reference strain rate which serves to render the somewhat surprising, as �45∘ specimens load the thermoplastic and thus
argument of the log function dimensionless; A is the failure stress at the strain rate sensitive polymer matrix in shear, whereas much of the load
reference strain rate and B is the slope, i.e., the strain rate sensitivity. in the quasi-isotropic specimens is carried by the 0∘ plies, which are not
Fitting this function to the data points from Table 1 using the Levenberg- expected to exhibit pronounced strain rate sensitivity due to their
Marquardt algorithm yields the parameters A and B including their mineral constitution. To further investigate this issue, future work
uncertainty estimates. As is obvious from Fig. 9, the failure stress at ε_0 is should address accurate testing of 0∘ specimens with Split-Hopkinson
higher for the quasi-isotropic specimens than for the �45∘ specimens, methods. This, however, is challenging because i) such specimens
A ¼ 477.2 � 5.0 MPa vs. A ¼ 200.1 � 0.5 MPa. More interestingly, the require much stronger holding forces than the laminates considered
strain rate sensitivity is identical within the uncertainty estimate: We here, and ii) the small failure strain reduces the time available for
observe B ¼ 0.038 � 0.005 in the quasi-isotropic case and equilibration such that accurate measurements might prove difficult.
B ¼ 0.035 � 0.001 for the �45∘ specimens. This is somewhat surprising,
as most of the load within the quasi-isotropic specimen is carried by the Funding
fibres of the 0∘ ply, and the strain rate sensitivity of the fibres itself is
believed to be weak [30]. In contrast, in the �45∘ specimens, the epoxy This work was supported by the Gips-Schüle-Stiftung and the Carl-
matrix is loaded in shear, and it is well known that epoxy resins exhibit Zeiss-Stiftung [Sonderlinie 2017/2018 Grundlagenwissenschaften mit
pronounced strain rate sensitivity [31,32]. However, these simple Anwendungsbezug].

317
G.C. Ganzenmüller et al. Composites Part B 171 (2019) 310–319

Appendix A. Example Split-Hopkinson High Speed Images

Fig. A11. Sequence of high-speed images showing the deformation behaviour of a quasi-isotropic basalt-fibre/epoxy composite test in our Split-Hopkinson Tension
Bar at a loading rate of 300 /s.

References [6] May M, Hesebeck O, Marzi S, B€ ohme W, Lienhard J, Kilchert S, Brede M,


Hiermaier S. Rate dependent behavior of crash-optimized adhesives – experimental
characterization, model development, and simulation. Eng Fract Mech 2015;133:
[1] Kogan FM, Nikitina OV. Solubility of chrysotile asbestos and basalt fibers in
112–37. https://doi.org/10.1016/j.engfracmech.2014.11.006.
relation to their fibrogenic and carcinogenic action. Environ Health Perspect 1994;
[7] Chen W, Hao H, Jong M, Cui J, Shi Y, Chen L, Pham TM. Quasi-static and dynamic
102(Suppl 5):205–6.
tensile properties of basalt fibre reinforced polymer. Compos B Eng 2017;125:
[2] De�ak T, Czig�any T. Chemical composition and mechanical properties of basalt and
123–33. https://doi.org/10.1016/j.compositesb.2017.05.069.
glass fibers: a Comparison. Textil Res J 2009;79(7):645–51. https://doi.org/
[8] Gilat A, Goldberg RK, Roberts GD. Experimental study of strain-rate-dependent
10.1177/0040517508095597.
behavior of carbon/epoxy composite. Compos Sci Technol 2002;62(10):1469–76.
[3] Liu Q, Shaw MT, Parnas RS, McDonnell A-M. Investigation of basalt fiber composite
https://doi.org/10.1016/S0266-3538(02)00100-8.
mechanical properties for applications in transportation. Polym Compos 2006;27
[9] Eskandari H, Nemes JA. Dynamic testing of composite laminates with a tensile split
(1):41–8. https://doi.org/10.1002/pc.20162.
Hopkinson bar. J Compos Mater 2000;34(4):260–73. https://doi.org/10.1177/
[4] Jamshaid H, Mishra R. A green material from rock: basalt fiber – a review. J Text
002199830003400401.
Inst 2016;107(7):923–37. https://doi.org/10.1080/00405000.2015.1071940.
[10] Daliri A, Vijayan A, Ruan D, Wang CH. High strain rate tensile properties of basalt-
[5] Wu JQ, Zhong ZL. The applications of basalt fiber in the automotive industry in the
fibre reinforced polymer composites. In: 17th Australian international aerospace
future. Adv Mater Res 2011;(332–334):723–6. https://doi.org/10.4028/www.
congress : AIAC 2017; 2017. p. 164.
scientific.net/AMR.332-334.723.

318
G.C. Ganzenmüller et al. Composites Part B 171 (2019) 310–319

[11] Spronk S, Gilabert Villegas FA, Sevenois R, Garoz G� omez D, Van Paepegem W. [22] Foster J. Comments on the validity of test conditions for Kolsky bar testing of
Tensile rate-dependency of carbon/epoxy and Glass/polyamide-6 composites. In: elastic-brittle materials. Exp Mech 2012;52(9):1559–63. https://doi.org/10.1007/
ECCM18 - 18th european conference on composite materials; 2018. s11340-012-9592-6.
[12] G�omez-del Río T, Barbero E, Zaera R, Navarro C. Dynamic tensile behaviour at low [23] Harding J, Welsh LM. A tensile testing technique for fibre-reinforced composites at
temperature of CFRP using a split Hopkinson pressure bar. Compos Sci Technol impact rates of strain. J Mater Sci 1983;18(6):1810–26. https://doi.org/10.1007/
2005;65(1):61–71. https://doi.org/10.1016/j.compscitech.2004.06.004. BF00542078.
[13] Jacob GC, Starbuck JM, Fellers JF, Simunovic S, Boeman RG. Strain rate effects on [24] Ganzenmüller GC, Langhof T, Hiermaier S. A constant acoustic impedance mount
the mechanical properties of polymer composite materials. J Appl Polym Sci 2004; for sheet-type specimens in the tensile split-hopkinson bar. EPJ Web Conf 2018;
94(1):296–301. https://doi.org/10.1002/app.20901. 183. https://doi.org/10.1051/epjconf/201818302064. 02064.
[14] Daniel IM, Werner BT, Fenner JS. Strain-rate-dependent failure criteria for [25] Plappert D. Quasi-static testing of basalt-fibre reinforced epoxy composites
composites. Compos Sci Technol 2011;71(3):357–64. https://doi.org/10.1016/j. according to DIN ISO 527, Tech. rep., to be published elsewhere. 2018.
compscitech.2010.11.028. [26] Ganzenmüller GC, Blaum E, Mohrmann D, Langhof T, Plappert D, Ledford N,
[15] Zhang H, Yao Y, Zhu D, Mobasher B, Huang L. Tensile mechanical properties of Paul H, Hiermaier S. A simplified design for a split-hopkinson tension bar with long
basalt fiber reinforced polymer composite under varying strain rates and pulse duration. Procedia Engineering 2017;197:109–18. https://doi.org/10.1016/
temperatures. Polym Test 2016;51:29–39. https://doi.org/10.1016/j. j.proeng.2017.08.087.
polymertesting.2016.02.006. [27] Guizar-Sicairos M, Thurman ST, Fienup JR. Efficient subpixel image registration
[16] Ramesh KT. High rates and impact experiments. In: Springer handbook of algorithms. Opt. Lett., OL 2008;33(2):156–8. https://doi.org/10.1364/
experimental solid mechanics. Boston, MA: Springer; 2008. p. 929–60. https://doi. OL.33.000156.
org/10.1007/978-0-387-30877-7. [28] Frew DJ, Forrestal MJ, Chen W. A split Hopkinson pressure bar technique to
[17] Eriksen RNW. Of DCAMM special reports. High strain rate characterisation of determine compressive stress-strain data for rock materials. Exp Mech 2001;41(1):
composite materials, vol. 179. DTU Mechanical Engineering; 2014. 40–6. https://doi.org/10.1007/BF02323102.
[18] Chen WW, Song B. Split Hopkinson (Kolsky) bar: design, testing and applications. [29] Kimberley J, Ramesh KT. The dynamic strength of an ordinary chondrite.
Springer Science & Business Media; 2010. Meteoritics Planet Sci 2011;46(11):1653–69. https://doi.org/10.1111/j.1945-
[19] Wang W, Makarov G, Shenoi RA. An analytical model for assessing strain rate 5100.2011.01254.x.
sensitivity of unidirectional composite laminates. Compos Struct 2005;69(1): [30] Yazici M. Loading rate sensitivity of high strength fibers and fiber/matrix
45–54. https://doi.org/10.1016/j.compstruct.2004.04.017. interfaces. J Reinf Plast Compos 2009;28(15):1869–80. https://doi.org/10.1177/
[20] Ochola R, Marcus K, Nurick G, Franz T. Mechanical behaviour of glass and carbon 0731684408090379.
fibre reinforced composites at varying strain rates. Compos Struct 2004;63(3–4): [31] Amos Gilat, Goldberg Robert K, Roberts Gary D. Strain rate sensitivity of epoxy
455–67. https://doi.org/10.1016/S0263-8223(03)00194-6. resin in tensile and shear loading. J Aerosp Eng 2007;20(2):75–89. https://doi.org/
[21] Hsiao HM, Daniel IM, Cordes RD. Strain rate effects on the transverse compressive 10.1061/(ASCE)0893-1321(2007)20:2(75).
and shear behavior of unidirectional composites. J Compos Mater 1999;33(17): [32] Werner BT, Daniel IM. Characterization and modeling of polymeric matrix under
1620–42. https://doi.org/10.1177/002199839903301703. multi-axial static and dynamic loading. Compos Sci Technol 2014;102:113–9.
https://doi.org/10.1016/j.compscitech.2014.07.025.

319

You might also like