You are on page 1of 37

Journal Pre-proof

Non-linear material characterization and numerical modeling of cross-ply basalt/epoxy


laminate under low velocity impact

Mohit Gupta, R.T. Durai Prabhakaran, Puneet Mahajan

PII: S0142-9418(19)30426-X
DOI: https://doi.org/10.1016/j.polymertesting.2020.106349
Reference: POTE 106349

To appear in: Polymer Testing

Received Date: 11 March 2019


Revised Date: 9 December 2019
Accepted Date: 11 January 2020

Please cite this article as: M. Gupta, R.T.D. Prabhakaran, P. Mahajan, Non-linear material
characterization and numerical modeling of cross-ply basalt/epoxy laminate under low velocity impact,
Polymer Testing (2020), doi: https://doi.org/10.1016/j.polymertesting.2020.106349.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. 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.

© 2020 Published by Elsevier Ltd.


Non-linear material characterization and numerical modeling of cross-ply basalt/epoxy
laminate under low velocity impact
Mohit Gupta a, R.T. Durai Prabhakaran b, Puneet Mahajan a
a
Department of Applied Mechanics, Indian Institute of Technology Delhi
b
The BioComposites Centre, Bangor University, Gwynedd, LL57 2UW, UK

Abstract

The low velocity impact behavior of basalt/epoxy composites, seen as an eco-friendly

replacement of glass-epoxy composites, has not been studied systematically so far. Here, the

elastic elasto-plastic properties, strengths, intralaminar and interlaminar fracture energies were

determined. The intralaminar energies were determined using compact tension and compression

tests. The elasto-plastic properties needed in the plastic potential were determined using off-axis

test. These properties are used in Finite Element (FE) code with an elasto-plastic damage model

developed earlier to simulate the impact response of cross-ply laminates basalt/epoxy laminates.

Low velocity impact (LVI) experiments at 10 J, 20 J and 30 J are performed on these

composites. The FE simulation is successful in capturing force, energy, deflection histories and

damage zones showing a close match to the experiments. A comparison of impact force history

and damage area (ultrasonic C-scan) of basalt-epoxy laminates with glass epoxy laminates

having same volume fraction shows nearly similar peak forces but the major axis of the

ellipsoidal damage zone was bigger in glass/epoxy laminates.

Keywords: Polymer-matrix composites (PMCs); Impact Behaviour; Damage Mechanics;

Mechanical Testing

1. Introduction

In recent years, the environmental issues have triggered attention towards eco-friendly materials,

motivating researchers to develop products using nature friendly raw materials. This has

1
prompted the composite industry to study mineral fibers such as basalt which do not require

additives during processing as a replacement for glass fibers. Colombo et. al. [1] compared

mechanical properties of basalt, glass and carbon fibers and concluded that basalt can fill the gap

between glass and carbon. Surprisingly, although impact loadings are among the most critical for

composites, impact behavior of basalt-epoxy composites and properties needed for numerically

simulating it have not been systematically studied. Few experimental and numerical studies on

impact of basalt-epoxy composites exist but validation of numerical predictions of impact force

or damage area with experiments is not available. Also, plasticity behavior and some fracture

properties needed for simulations are sparsely available in literature for these composites.

On the other hand, impact behavior of fiber reinforced polymeric composites made of glass,

carbon and aramid and their susceptibility to damage has been widely studied using

experimental, analytical and numerical methods. It is seen that unidirectional (UD) laminates of

above composites are more susceptible to damage in transverse direction due to their lower

transverse strengths, which results in initiation of matrix cracking and delamination at

comparatively low loads. In recent years, numerical methods have been used to predict inter-

laminar and intra-laminar damage in these composites and these predictions have been validated

with experimental force-time and damage measurements [2,3,4]. Most of the proposed numerical

models in literature are based on CDM approach and degrade the stiffness of the laminae using

damage variables of different modes of failures [4,5]. Lately, plasticity combined with damage

has been used to study loading-unloading during impact [6,7] of composites. Plasticity arises

from the non-linear behavior of resin [8] and damage in composite under shear and transverse

loadings [9]. The use of these numerical models requires as input elasto-plastic properties of

composites, damage initiation strengths and fracture properties in different modes. Determining

2
these properties needs extensive experimentation. For validation of the numerical results, impact

experiments and techniques to predict the extent of damage are also needed.

Studies on basalt UD fabric composites have separately evaluated tensile, fatigue, flexural and

compressive and ILSS properties [1,10,11,12]. Recently Scalici et. al. [18] prepared laminates

from UD fabric of 580 gsm and in a single paper determined tensile tests, flexural tests,

Interlaminar shear strengths (ILSS), and interlaminar fracture energies. Lopresto et. al. [14]

investigated and compared plain woven composite of basalt and glass of 200 gsm and 290 gsm

respectively for epoxy matrix and found compressive strength, Young's modulus and flexural

behaviour of basalt were superior to glass, whereas glass composite had higher tensile strength.

Several other researchers [15-18] have compared basalt fiber composites with other reinforcing

materials especially with glass fibers and concluded that it has higher modulus, bending and

compressive strengths.

A lot of work on low velocity impact (LVI) for CFRP and GFRP is available in literature

[19,20], but basalt fiber reinforced polymers (BFRP) have caught interest mostly in the last

decade. Najafi et al. [21] used hybrid basalt and carbon fabric composite and found that as the

weight fraction of basalt is increased in the laminate the energy absorption increases and was

maximum for a 0.83:0.17 fraction of basalt-carbon respectively. Dehkordi et. al. [22] used basalt

and nylon yarns to fabricate plain woven basalt, nylon and intraply basalt-nylon hybrid fabric for

studying LVI performance. Basalt fabric has also been investigated using plain woven fabric and

making hybrid laminates with flax and glass fibers [23]. Being a mineral fiber many suggests to

use it as outer layers with inner layers of low strength plant fibers. LVI study on basalt UD fabric

laminates were performed by Suresh et. al. [24] and Andrew et. al [25], they used cross ply and

UD laminates respectively for studying LVI. They compared basalt laminates with glass and

3
hemp laminates and their hybridized forms concluding that hybridized laminates have better

impact energy absorption capabilities compared to single fabric laminates. Sfarra et al. [13]

compared the damage features on woven glass and basalt laminates and observed an increased

directionality of impact damage in BFRP compared to GFRP. In 2016 Boria et. al. [26] simulated

LVI on basalt/flax hybrid laminates in LS-Dyna using only tensile and flexural properties.

From the literature review it is evident that UD basalt laminates have not been investigated much

in depth for static mechanical characterization, LVI experiments and numerical simulation.

While tensile, compression, ILSS, Interlaminar fracture energies and flexural properties are

available at different places, no single study provides data for different fiber volume fractions.

To simulate an impact phenomenon using CDM in present study laminates basalt are fabricated

using vacuum infusion technique and their static elastic moduli and strength properties are

determined. Double Cantilever Beam (DCB) test, Compact tension and compression tests are

performed to determine the interlaminar and intralaminar fracture energies. Off axis tension tests

to find the parameters [27] needed in the plasticity potential are also implemented. These

properties were further used in the numerical model based on continuum damage mechanics

coupled with plasticity [6]. The interlaminar fracture energies are needed in predicting

delamination using cohesive zones. The model is able to capture interlaminar and intralaminar

damage, plastic deformations due to indentation and delamination. The numerical predictions of

force and extent of damage from simulations are compared with force-time curves from impact

experiments and measurements from ultrasonic c-scan. Since, basalt is seen as replacement for

glass, a comparison of impact behavior with glass with similar volume fraction is also given.

4
2. Materials and mechanical characterization

2.1 Manufacturing

UD Basalt Fabric of aerial density 320 g/m2 were used to prepare a layup of 8 layers [0/90]2S for

impact tests. Volume fractions were found to be 49.2 % for basalt and 48.3 % for glass

laminates. Basalt fabric UD tows are stitched together with the sewing nylon threads but without

any backing fibers commonly seen in glass fabrics. Low viscosity epoxy from HEXION (RIM

135) and hardener (RIMH 137) were mixed in the ratio 100:30 by weight. Laminates were

prepared using vacuum assisted resin infusion molding technique (VARIM).

2.2 Modulus and Strength Properties:

For characterization of composite laminates, the samples were cut with water jet to ASTM

dimensions. Digital Image Correlation (DIC) technique with images obtained using cameras

with a resolution of 5 megapixels was used for measuring displacements. These were analyzed in

the software VIC-3D [28]. Tension, compression and Iosipescu shear experiments were

performed to determine the moduli and strengths for basalt and glass laminates and results are

reported in table 1.

The elastic properties of Basalt/epoxy and Glass/epoxy were very close except for transverse

tensile strength (S2t) and longitudinal compressive strength (S1c). Transverse tensile strength is

higher for glass/epoxy laminate as glass fabric has 5% backing fiber bundles in transverse

direction. The lower S1c for basalt/epoxy could be explained by the lower microbuckling loads

due to smaller diameter of basalt monofilament (12.90 µm) compared to E-glass fibers (16.70

µm). Scanning electron microscopy and X-ray tomography images of specimens showed. kink

band- failure in glass/epoxy specimens whereas basalt samples broke in two pieces without the

kink band being captured.

5
For a UD laminate of volume fraction 62 % Scalici et. al. [18] reported E11= 40.4 GPa, E22= 8.58

GPa, S1t = 619 MPa, S2t = 6.16 MPa, G12 = 2.31 GPa. Zhang et. al [29] used basalt UD laminate

of v.f. 22.1 % for high strain rate characterization, quasi-static longitudinal strengths reported is

521 MPa.

Table 1: Elastic and strength properties for composite UD laminate.

Property Average (% Std. Dev)


Basalt/Epoxy Glass/Epoxy
E11 [fiber direction] (GPa) 37.95 (± 2.56 %) 34.95 (± 3.16 %)

E22, E33 [transverse direction] (GPa) 9.82 (± 5.75 %) 10.93 (± 4.36 %)

G12,G13 [shear modulus in-plane] (GPa) 2.8 (± 4.67 %) 2.5 (± 5.62 %)

G23 [shear modulus out of plane] (GPa) 4.14 5.12

υ12, υ13 [Poisson’s ratio in-plane] 0.30 (± 7.94 %) 0.27 (± 6.61 %)

υ23 [Poisson’s ratio out of plane] 0.19 0.07

Density (kg/m3) 1860 1805

S1t [tensile strength fiber direction] (MPa) 746.90 (± 6.10 %) 730.00 (± 2.56 %)

S2t=S3t [tensile strength transverse direction] (MPa) 20.28 (± 4.83 %) 85.39 (± 5.43 %)

S1c [Compressive strength fiber direction] (MPa) 311.01 (± 7.60 %) 450.29 (± 6.21 %)

S2c=S3c [Compressive strength transverse direction] (MPa) 90.45 (± 4.55 %) 105.80 (± 5.62 %)

2.3 Interlaminar (Mode-1) fracture toughness characterization:

Interlaminar fracture energies are required to predict propagation of delamination cracks Double

cantilever beam test was performed in accordance with ASTM D 5528 [30] on 5 kN DAK UTM

at crosshead speed of 5 mm/min to calculate the mode I interlaminar fracture toughness.

Specimens were 24 layered UD laminates with initial crack length ao=40 mm. As seen in Fig.

6
1(a) no fiber scissoring was observed during the tests. Modified Beam Theory (MBT) data

reduction scheme was used to calculate the GIC (energy release rate) using the Eq. (1).

3Pd (1)
G IC =
2b(a+|∆|)

'P' and 'd' are the load and cross head travel and 'b' is the width of specimen. Instantaneous crack

length 'a' was obtained by image analysis and a reference scale was used to calculate the crack

length. '∆' is obtained by linear regression of the cubic root of the compliance (C1/3) vs the

instantaneous crack length 'a'. The fracture toughness curves shown in Fig. 1(b), indicate a

fracture initiation value of 0.7 kJ/m2 and average propagation value of 1.3 kJ/m2 (shown with a

dotted line) in Fig. 1(b). These values are higher than those reported by Scalici et. al [18].
2
S-1
1.8 S-2
S-3
1.6
Extended UTM 1.4
Display 1.2
GIC (kJ/m2)

1
0.8
0.6
Reference
0.4
Scale
0.2
Crack
0
Front 40 60 80 100 120 140
Crack length, 'a' (mm)
(a) (b)
Fig. 1: (a): Mode-1 DCB test setup. (b) Fracture toughness values for mode-1 DCB tests.

2.4 End Notch Flexure characterization:

End Notch Flexure (ENF) test was performed on the composite specimen as per ASTM standard

D7905 [31] to determine the mode II Interlaminar fracture energy. In ENF test, the Mode II

critical energy release rate is given by:

9a 2 Pc2 C
G IIC = (2)
2b(2L3 +3a 3 )

7
where GIIC is the mode II interlaminar fracture energy, PC is the critical load and represents the

maximum load on the load-displacement curve, 'a' is the initial crack length, 'b' is the specimen

width, 'L' is the half of span length and 'C' is the compliance, calculated as the ratio of load point

displacement to the applied load. The initiation fracture toughness was calculated to be 1.505

kJ/m2. The interlaminar fracture energies calculated above can also be used for intralaminar

matrix dominated transverse damage growth.

2.5 Characterization of Intralaminar Fracture Toughness in fiber direction:

For UD laminae the material response in longitudinal direction is dominated by fibers. In

transverse and through thickness it is dominated by matrix properties. For the crack to propagate

through the laminae in longitudinal and transverse direction requires the fracture energies which

are obtained from fracture toughness tests. The crack may propagate either cutting through the

fibers or along the fibers through the matrix. Thus two types of fracture energies are required to

evaluate the parameter for intralaminar damage to propagate. Now a crack can propagate in

tension or compression thus enabling four different crack propagation ways i.e. in longitudinal

(tension and compression) and transverse (tension and compression) to the fiber length. For

predicting damage propagation in the composite in longitudinal direction intralaminar failure

energies in fiber tension and compression modes are needed. Energies consumed in longitudinal

modes are much larger compared to interfacial fracture energies and are important at high impact

energies. Compact Tension (CT) and Compact Compression (CC) tests are used to evaluate these

and were performed according to the dimensions and procedure described by Pinho et. al. [32].

The critical strain energy release rate can be calculated using the change in compliance with

corresponding crack length as:

PC2 dC
(Γ )
Lam
C i
=
2t da
(3)

8
Pc is the maximum measured load, t is the specimens thickness, Γ Lam
IC is the fracture toughness of

laminate, i = ‘t’ for tension mode, ‘c’ for compression mode.

(a) (b)

Fig. 2: (a) Schematic diagram showing the CT test specimen dimensions. (b): Instantaneous
crack detection using DIC for CT test.

300
CT S-1
250 CT S-2
Fracture Toughness (kJ/m2)

200

150

100

50

0
26 28 30 32 34 36 38 40
Crack length, a(mm)
(a) (b)

Fig. 3: (a) Fracture surface showing fiber pull-out of a compact tension test. (b) R-curve for
the tensile fracture toughness tests using CT specimen.

A 24 layer laminate with layup [(904/0)2/902]S with 0° fibers aligned along the loading direction.

When the specimen is loaded the crack propagates at the crack tip perpendicular to the loading

direction cutting across the fibers of 4 layers of 0° and through the matrix parallel to the fibers in

9
remaining 20 layers. The layup schematic diagram with dimension for both the specimens is

shown in Fig. 2(a) and 4(a). This propagation of crack length was monitored using fine speckles

over the surface of specimen and is seen as straight red line in Figs. 3(b) and 4(b).

(a) (b)

Fig. 4: (a) Schematic diagram showing the CC test specimen dimensions. (b) Instantaneous
crack detection using DIC for a CC test.

140

120
Fracture Toughness (kJ/m2)

100

80

60
CC-S1
40 CC-S2
CC-S3
20

0
15 20 25 30 35 40 45
Crack length, a (mm)
(a) (b)

Fig. 5: (a) 3D-reconstructed model from the Micro-CT scan images of the CC tested specimen.
The cut-section slice image shows a 53.2° fractured plane formed due to compression. (b) R-
curve for the tensile fracture toughness tests using CC specimen.

The fractured surface for the CT specimen was a flat surface as shown in Fig. 3(a). The CC

specimen failed due to transverse failure of plies at an angle of 53.2° as shown in Fig. 5(a) on the

right side with a 3D-reconstructed model obtained from Micro-CT scan.

10
The compliance calibration method (CCM) suggested by Laffan et. al. [33] was employed to

calculate the dC/da in Eq. (3) using the experimental C vs a. data for CT and CC tests shown in

Fig. 3. Once, the fracture energy of the laminate was known, the fracture energy of 90o layers is

subtracted to find out the fracture toughness of 0o as shown in Eq. (4) below

(Γ ) t
Lam Lam
- ( Γ mC ) t m
(Γ ) f
C i =
C i

tf
i (4)

where ΓfC is fracture toughness for 0° fibers and Γ Cm is for 90° fibers, where values of ( Γ mC ) , ( Γ mC )
t c

are interlaminar fracture toughness values obtained from interlaminar mode-I (1.3 kJ/m2) and

mode-II (1.505 kJ/m2) experiments respectively in section 2.3 and 2.4. The calculated

intralaminar fracture energy values for compact tension and compact compression are 151 kJ/m2

and 101 kJ/m2 respectively as shown in Figs. 3(b) and 5(b).

2.6 Off axis tension test and plasticity parameters:

Composite specimens in transverse and shear loading exhibit plasticity [8,9]. Sun and Chen [27]

proposed a single parameter model to describe the plastic behaviour of UD laminate. From the

small strain decomposition theory, total strain, ( ε ) in the material can be decomposed into

plastic ( ε p ) and elastic ( ε e ) parts as:

ε = εe + ε p (5)

Assuming negligible plastic behavior in the fiber direction, Sun and Chen [27] showed that the yield

condition could be written in terms of equivalent stress σ e and flow stress as σ f

p
f yield (σ e , σ f ) = σ e − σ f (ε ) = 0
1
3 2
where σ e =  ( (σ 22 − σ 33 ) 2 + 4σ 23
2
+ 2a 66 (σ 122 + σ 132 ) )  (6)
2 
p
and ε is the effective plastic strain.

11
The coefficient a 66 is the single parameter in the equivalent stress σ e which needs to be determined

experimentally. For plane stress condition Eq. (6) reduces to

1
3 2
f =  (σ 222 + 2a 66σ122 )  − σ f (ε ) = 0
p
(7)
2 

Using associative flow rule, incremental plastic strains can be written in terms of the plastic

potential 'f' as

∂f
d εijP = dλ (8)
∂σ ij

For plane stress, the plastic strain increment and effective plastic strain increment are given by

 d ε11P   0 
 P  
d ε22  =  σ 22  d λ (9)
d γ P  2a σ 
 12   66 12 

1/2
P 2 2 
d ε =  (σ 22 + 2a66σ 122  d λ (10)
3 

To determine a 66 Sun and Chen proposed an off axis tension test with loading direction X at an
angle ϕ to the fiber direction, so that
1
3  2
σ f =  (( sin ϕ )4 + 2a 66 (sin ϕ ) 2 (cos ϕ )2 )  σ x (11)
2 

1
3  2
ε = ε /  ((sin ϕ )4 + 2a 66 (sin ϕ )2 (cos ϕ )2 )
p p
(12)
2
x

12
(a) (b)
Fig. 6: (a) Schematic depicting a loaded off-axis tensile specimen with stresses at an element
within at an angle φ. (b): Strain contours showing from DIC for a 20° off-axis specimen
loaded in tension at failure.

An off-axis specimen is cut at angle from a UD laminate and when loaded the stresses that

generate at the center are as shown in Fig. 6(a). Longitudinal strains (εxx) were captured using

DIC and the contours over the speckled region obtained after the post processing at the point of

failure are as shown in Fig. 6(b).

Curves of σ x and ε x are plotted for different values of ϕ and a 66 is so chosen that these curves

collapse into a single master curve. In practice, the value of a66 is increased from zero till the R-

square value of the fitting data reaches a maximum value. This master curve can be expressed as

ε P = A( σ f ) n (13)

13
90 90

80 80

70 70

Effective Stress (MPa)


Stress (MPa) [σx]

60 60

50 50

40 40
20 deg 20 deg
30 30 deg 30
30 deg
40 deg
20 40 deg
20

10 10

0 0
0.000 0.002 0.004 0.006 0.008 0.010 0.000 0.002 0.004 0.006 0.008 0.010 0.012
Total Strain [εx] Effective Plastic Strain
(a) (b)

Fig. 7: (a) Stress vs Total strain for 20°, 30° and 40° off axis specimens. (b) Effective Stress-
effective plastic strain plots for a 66 =2.1

For basalt/epoxy, tests were performed at three different angles ( ϕ = 20°, 30° and 40°) and

curves of σ x and ε x for these are shown in Fig. 7(a). Values of A and n are found by fitting Eq.

(13) using the experimental data from the collapsed master curve in Fig. 7(b). The values of a66,

A and n are given in table 2.

Table 2: Off-axis test plasticity parameters, Interlaminar and Intralaminar Fracture toughness
values.

Plasticity Material Intralaminar Fracture Interlaminar Fracture


Parameters Energies Properties
a 66 2.1 (Γ fC ) t 151 kJ/m2 G IC 1.3 kJ/m2
A 2.15xe-07 (Γ fC ) c 101 kJ/m2 GIIC 1.505 kJ/m2
n 2.46 (Γ Cm ) t 1.3 kJ/m2
(Γ Cm ) c 1.5 kJ/m2

14
3. Numerical Model:

3.1 Constitutive model of Composite

Damage growth in a material starts at microscopic level with nucleation of defects which

deteriorates the load bearing capacity of the material. In damaged material the effective stress

(σe) is termed as the stresses being applied over the intact cross-section of the material.

For a three dimensional stress state, effective stress is related to the nominal stress by introducing

a second order tensor M as shown in Eq. (14).

s = M:s e (14)

where s = [s s s s s s ]T
11 22 33 23 13 12

M is the damage effect tensor written as:

1 − D11 0 0 0 0 0 
 0 1 − D 22 0 0 0 0 
 
 0 0 1 − D33 0 0 0  (15)
 

M= 0 0 0
1
2
( (1 − D22 ) + (1 − D33 ) ) 0 0 

 
 0 0 0 0
1
( (1 − D11 ) + (1 − D33 ) ) 0 
 2 
 
 0

0 0 0 0
1
2
( (1 − D11 ) + (1 − D 22 ) ) 

where D11 , D22 , D33 are damage variables in fiber, in plane transverse and out of plane

transverse directions. Taking the nature of loading into consideration damage variables can be

further elaborated as;

D11 = 1 − (1 − d1t ) (1 − d1c )


D22 = 1 − (1 − d 2t ) (1 − d 2 c ) (16)
D33 = 1 − (1 − d 2 t ) (1 − d 2 c )

where d1t and d1c are damage variables for fiber tension and fiber compression respectively, d2t

and d2c are damage variables for matrix tension and matrix compression. These variables are

assumed to be constant throughout the ply thickness.


15
Using the principle of strain equivalence, the constitutive relationship between stress and strain

for damage model is represented as:

s = M : Eo : ε e (17)

σe = Ε(d) : (ε − ε p ) (18)

Where E(d) is second order constitutive tensor for damaged laminate and Eo is for undamaged

composite material. εp is the plastic strain obtained by integrating Eq. (10).

3.3 Intralaminar Damage initiation criteria

Physically based damage models are proposed in order to predict initiation and propagation of

each failure modes in laminated fiber-reinforced composites. Failure modes are classified as

intra-laminar (matrix cracking and fiber pullout/breakage) and interlaminar (delamination)

damage. The Intralaminar initiation criteria given by Hashin [34] for fiber tension, fiber

compression and matrix tension and by Puck and Schurmann [35] for matrix compression

damage initiation are used. These are described in table 3.

Table 3: Damage initiation criteria's for tension and compression in longitudinal and transverse

fiber direction.

2 2 2
σ  τ  τ  Fiber tension
R1t =  11  +  12  +  13  ≥ 1 forσ 11 ≥ 0
 S1t   S12   S13 
2
σ  Fiber compression
R1t =  11  ≥ 1 forσ 11 ≤ 0
 S1c 
2 2 2
 σ   σ   σ 2 −σ σ   σ 2 +σ 2  σ 
R2 t =  22  +  33  +  23 2 22 33  +  12 2 13  ≥ 1R1t =  11  ≥ 1
 S2t   S2t   S 23   S12   S1c  Matrix Tension

forσ 22 + σ 33 ≥ 0

16
2 2
 σn   σ ln 
R2 c =  A t n n  +  ≥ 1 for σ 22 + σ 33 < 0
 S + µt σ n   S12 + µl σ n 
n n
 23 
σ n = σ 22 cos θ + σ 33 sin 2 θ + 2σ 23cosθ sinθ
n 2

σ tn = −σ 22cosθ sinθ + σ 33cosθ sinθ + 2σ 23 (2 cos 2 θ − 1)


Matrix compression
σ ln = σ 12cosθ + σ 13 sinθ
S 2 c  1 − sinψ 
S 23A =  
2  cosψ 
ψ = 2θ − 90°
µtn = tan(2θ − 90°)

3.4 Intralaminar Damage propagation:

Once the loading reaches the damage surface, damage is initiated and its evolution is governed

by an exponential damage law proposed by Matzenmiller et. al. [36] as following

 1 
d i = 1 − exp  (1 − Ri ) m  i ∈ [1t ,1c, 2t , 2c] (19)
 me 

where ‘m’ is material softening parameter and ‘e’ is Napier’s constant. This parameter 'm'

controls the post damage propagation behavior and has different value for different damage

propagation modes. In the damaged region continuous degradation of stiffness is observed with

the increase in strain. Strain localization problem in FE analysis arises due to strain softening

behaviour of material during damage growth, thus making the solution element size dependent.

For mesh size independent solution, based on Bazant’s crack band theory [37] Singh and

Mahajan [6] have proposed that for different modes ‘m’ can be calculated from the Eq. (20).

1 ∞  1 
ΓC = σ 0ε 0 Lc + ∫ exp  (1 − Ri ) m  E 0ε d ε Lc (20)
ε
2 0  me 

Here Lc is the characteristic length of the element for a particular damage mode and Γ C is the

fracture energy at failure for the same damage mode. For a 3-D, 8 node brick element

corresponding Lc values with their fracture energies are shown in table 4.

17
Table 4: Characteristic lengths and their corresponding fracture energy values for different
intralaminar fracture modes

Characteristic Fracture Calculated


Damage Mode
length Energy value of ‘m’
Fiber tension Lc=Lx ΓC =(ΓCf )t 0.335

Fiber Compression Lc=Lx ΓC =(ΓCf )c 0.6

Matrix tension Lc=Ly ΓC =(ΓCm )t 0.17

Matrix compression Lc=Ly / cos(θ) ΓC =(ΓCm )c 0.4

where Lx and Ly are the lengths of element along fiber and transverse directions respectively.

The values of corresponding fracture energy were found in section 2.3. It is assumed that for

matrix compression the failure occurs at a fracture angle, θ with the direction of applied load.

Generally, the value of θ is observed to be equal to 53° as per [35].

To determine ‘m’, FE analysis of single element of same size as used in impact simulations later

was performed. For each propagation mode using the values of Lc from table 4 and Γ C from table

2, the value of ‘m’ was adjusted in such a way that Eq. (20) is satisfied.

3.5 Inter-Laminar damage using cohesive surfaces:

To model the interface between laminas inbuilt ABAQUS cohesive surface formulation was

used. Prior to failure initiation, traction displacement law can be written as:

σ 11   K 33 0 0   δ33 
    
σ 22  =  0 K 31 0   δ31  (21)
σ   0 0 K 32  δ32 
 33  

where 1, 2 and 3 denoted direction of local co-ordinate system, in which 3 is the out of plane and

1 and 2 are in plane directions. For present work,

18
Ci
Ki = i Î [33,31,32] (22)
10te

where Ci is the tensile modulus (E33) for out of plane stiffness and shear modulus (G13 and G23)

for in plane stiffness value. te is the element thickness of the adjacent layers. The damage

initiation is defined based on maximum nominal stress. Once the damage has initiated,

subsequent traction-displacement is governed by damage variable D given by

∆ mf σ eff
D=∫ dδ (23)
∆0m Gc − G 0

where , 0 signifies damage initiation and f for final failure

δm = δ33
2
+ δ31
2
+ δ32
2
(24)

σ eff = σ 33 + σ 312 + σ 322


2
(25)

The fracture energy Gc is given by Benzegagh-kenane (BK) law.

η
 G +G 
G C =G IC +(G IIC +G IIIC -G IC )  IIC IIIC  (26)
 G IIIC 
Here, η is a material parameter, often used values are 1 or 2. In present study η =1 is adopted and

found suitable for current work. GIIIC and GIIC are assumed equal. The values of fracture

energies GIC and GIIC used in simulation are given in table 2.

3.6 Computational procedure:

The above model is implemented in ABAQUS/Explicit using the user-defined subroutine

VUMAT. The following procedure was applied for updating nominal stress which is based on

Euler's forward integration scheme. Computational procedure is shown in Fig. 8.

19
Fig. 8: Computational procedure flow chart showing the stress update
procedure for an integration point.

4. Drop-weight low velocity impact (LVI) experiments

The proposed formulation is further used to study the low velocity impact (LVI) response of

cross-ply basalt/epoxy composite. Impact tests were carried out on square plates clamped from

all four sides using an INSTRON Dynatup 9250 HV drop tower in accordance with the method

prescribed in ASTM D7136. 175 mm × 175 mm laminates were used for impact tests. The

square specimen was subjected to an out-of-plane impact (perpendicular to the plane of the

laminate) using a drop- weight device with a hemispherical striker tup with mass 13 kg and

diameter 12.7 mm. The specimen was placed in the rigid fixture and securely fixed using four

screw clamps between upper

20
and lower frame which resulted an open effective area of 125 mm × 125 mm. The close up of the

impact machine with tup and fixture for gripping laminate is shown in Fig. 9. The force data

obtained from output of load cell installed inside the impactor tup is recorded digitally.

The laser flag in the machine captures the incident velocity just before the impactor hits the

plate, defined as v1 . Force as a function of time is integrated for the impact duration to calculate

the velocity v 2 of the falling mass at any time t2 using the relation below

1 t2
v2 = v1 +
mI ∫
t1
( F − mI g )dt (27)

where mI , v and F are the impactor mass, impactor velocity and contact force. The change in

kinetic energy of the falling mass provides the absorbed and recovered kinetic energy as shown

in Fig. 9. The closed area inside the force-displacement curve (loop) signifies the energy

absorbed by the plate and the area under the curve during the rebound phase signifies recovered

energy. All the incident impact energy is not recovered and a fraction of that is dissipated in the

form of plate damage, plasticity and friction between various interfaces. Three samples each

were tested for glass and basalt laminates for the condition’s specified in table 5.

21
Fig. 9: Impactor tup and fixture to hold the laminate.

Force F (t)
Sensor

1 t2
v2 = v1 +
mI ∫
t1
( F − m I g ) dt

v2
∆d = ∫v1
vdt

mI v12 mI (v2 (t ))2


EI = −
2 2

EI = ER + ED

Fig. 10: Schematic of impact experiment showing stages from force data collection to

energy time plots.

22
Table 5: Specification of Impact Tests

Impactor Sample Dimensions Impact Velocity Energy


mass (kg) [ velocity flag ](m/s) Impacted (J)
1.25 10
13 (175 mm x 175 mm x 2.08 mm) 1.78 20
2.15 30

5. Results and discussions

5.1 Impact Experiments on basalt/epoxy and comparison with glass/epoxy composites

Three samples of basalt/epoxy laminates were tested following the procedure and conditions

described in section 4. Three laminates were tested at each energy level and Fig. 11 shows load-

time plots for these laminates. A sudden load drop is observed in all the curves between the zone

of 0.1 to 0.5 kN loads (at the beginning of the strike of impactor). This disturbance in the force

values is named as Hertzian failure which denotes the initiation of damage in the form of

interlaminar delamination [3]. Sevkat et al. [20] proposed that the initial undulations in the load

drop was an indication of an abrupt transition of the specimen from an intact state to a damaged

state. Simulation confirmed these start of undulations is the initiation of delamination. The 10 J

curve is almost symmetrical and induces only matrix cracking in the laminate. 20 J experiments

start showing some irregularities at the top which depicts matrix cracking along with initiation of

fiber damage, whereas in the 30 J there is prominent fiber breaking as shown by a steep fall in

the forces.

Fig. 12 and table 6 compare impact behavior of basalt/epoxy versus glass/epoxy laminates

having same volume fraction. The load-time graphs in fig. 12 are shifted on the time (x) axis to

provide a clear view to individual curves without any overlap. Basalt laminates behave little

stiffer as evident from the rising slope of the load-time curves and resulted in slightly higher

23
loads compared to glass for 10 J and 20 J. No fiber damage is observed till 20 J in both the

laminates, but at 30 J fiber damage is observed in basalt laminates which is evident by a sudden

drop in force whereas glass laminates don't show any fiber damage.
6
30 J

5 20 J

4
Load (kN)

10 J
3

0
0 5 10 15 20 25 30 35 40
Time (ms)

Fig. 11: Force-time plots for three impact Fig. 12: Force-time plots for basalt/epoxy
experiments each at 10 J, 20 J and 30 J on and glass/epoxy laminates for 10 J, 20 J and
basalt/epoxy laminates. The average value of 30 J.
these is used in Fig. 14 (a). (Note: The curves
are shifted on the time axis)

Table 6 shows comparison of impact data of basalt/epoxy (B.E.) and glass/epoxy (G.E.)

laminates. Energy absorption ranges from 34% to 82% for basalt laminates and for glass ones is

46% to 73%. At impact energy of 10 J, absorption is found less in basalt laminates but as the

impact energy is increased to 20 J and 30 J basalt laminates start experiencing fiber damage

initiation and fiber failure in few layers which increases the energy absorption compared to glass.

Till fiber failure is dominant, peak forces are higher and maximum central deflections are lower

in basalt laminates due their higher in-plane stiffness. At 30 J, basalt laminates show higher

maximum central deflection due fiber failure in few layers at bottom which makes the laminate

softer. Damage areas calculated using the area of ellipse shows that basalt has lower damage

spread compared to glass at all energy levels.

24
LVI study on cross-ply (12 ply) were performed by Suresh et. al. [24] who used 200 gsm fabric

to make laminates by hand-layup technique. Samples were impacted at 2.17 J using a circular

fixture with internal opening of 40 mm and a peak force of 1.82 kN was achieved with 40 %

energy absorption. Andrew et. al. [25] used a 16 layered UD basalt and glass laminates for a

similar energy impact and concluded that basalt absorbed more energy compared to glass with a

peak force of 1.293 kN and 1.202 kN for glass and basalt respectively.

Table 6: Comparison of Impact data analysis for Basalt/epoxy vs glass/epoxy.

Max. Damage area


Impacted Energy Peak
% Energy central (Ultrasonic C-scan) Area
Energy Absorbed force
absorption deflection [π x a x b]
(J) (J) (kN)
(mm) a(mm) b(mm)
B.E. 10.16 03.46 34.06 3.09 08.18 19.00 14.50 865.07
10 J
G.E. 10.40 04.84 46.54 2.88 08.79 27.80 13.00 1134.80
B.E. 20.64 13.86 67.15 4.67 10.41 27.50 17.00 1468.00
20 J
G.E. 20.06 12.10 60.30 4.59 10.89 38.30 14.80 1779.90
B.E. 30.10 24.96 82.92 5.43 12.52 49.50 18.50 2875.50
30 J
G.E. 30.16 22.20 73.61 5.67 12.40 55.00 18.80 3246.76

5.2 Numerical Simulation of impact of basalt/epoxy laminates

Eqs. 15 to 24 were implemented as a subroutine (VUMAT) in ABAQUS and the results were

validated against those from LVI experiments performed on Basalt/epoxy laminates. Material

properties needed for simulations were taken from the tests performed in section 2. The layup

and test parameters are similar to as shown in table 5. The eight layered composite laminate,

steel fixture for gripping and a spherical steel impactor of 12.7 mm diameter were meshed with

three dimensional 8-noded brick elements and single element across the thickness is taken in

each ply as shown in Fig. 13. Taking the geometrical configuration and loading conditions into

consideration, quarter symmetry is used in simulation to reduce the computational time. The

25
preload applied on the frame for tightening the plate is modeled as equivalent pressure applied at

the top face. Bottom face of the frame is fixed in all directions.

Fig. 13: Details of the Simulation model.

5.3 Comparison between experiments and simulation

The output from the simulations was compared against the experimental results at three different

energy levels. Fig. 14(a) and 14(b) shows force-time and energy-time curves for 10 J energy

level

obtained from the numerical simulation and the experiments The curves match reasonably well.

The damage was due to delamination and matrix cracking for low impact energies and no fiber

damage was observed. When the energy is increased to 20 J the predicted peak force is lower by

6%. Energy absorbed in damage as predicted by simulation is less compared to that in

experiments. For 20 J impact energy, fiber damage initiates in some elements but does not reach

the critical damage value of “1” in any element.

As energy is increased from 20 J to 30 J fiber damage is observed at the bottom side of the plate

and damage grows subsequently on the impactor side, as seen in the sequence of inset images in

26
Fig. 16. For the 30 J case a sudden drop in force is observed from approximately 5.8 kN load to

4.2 kN in the experiments and the same is also observed in the numerical simulations as seen in

Fig. 14(a). These sudden drops indicate catastrophic fiber failure in some layers.

(a) (b)
Fig. 14: Comparison of experimental and simulation results of (a) force-time history and (b)
impactor energy-time history.

(a) (b)
Fig. 15: Comparison of experimental and simulation results of (a) Central deflection-time
history and (b) Force-Deflection history.

27
Central deflection plots from simulations were also compared with the experiments as shown in

Fig. 15(a). Maximum deflection is almost accurately predicted from the numerical simulations,

with a small deviation (0.5 ms) in peak on time axis for 30 J energy level. Fig. 15(b) shows the

Fig. 16: Fiber damage extent at different energy levels.

10 J 20 J 30 J

(a)

(b)

(c)

Fig. 17: Damage area comparison between (a) Glass/Epoxy and (b) Ultrasonic C-scan of

28
Basalt/Epoxy Laminates (c) Numerical Simulation prediction of Basalt/epoxy laminates.

force versus central deflection plots for experiments and simulations. The area inside the loop

shows the energy dissipated in plastic deformation and damage. The closed form suggests that

the complete perforation of plate did not occur and the impactor rebounded with the left over

strain energy in the plate imparted back to it.

As the plate bends during impact besides intra-laminar damage there is sliding between the

layers and the bond between them is broken and energy is absorbed in delamination. To quantify

the damage, ultrasonic C-scans were performed for basalt/epoxy laminates whereas partially

translucent nature of glass laminates helped in capturing the damage area directly. A 20 MHz

probe was used to scan the damage area for basalt laminates. From Figs. 17 (a) and (b) it’s

evident that along the major axis of the damage area in basalt laminates is shorter and minor axis

are longer as compared to glass ones. This major axis is along the fiber direction of the bottom

most layer.

29
Fig. 18: Micro-CT scan of the impacted laminate at 30 J impact energy.

The reason for this higher eccentricity of damage ellipse in basalt laminates is due to its lower

transverse strengths compared to glass laminates leading to more cracks in transverse direction.

Fig. 17(c) shows the numerical predictions of the maximum damage areas which matches well

with that obtained in the C-scan images for the lower energies (10 J, 20 J) but for higher energy

(30 J) the length of major axis of damage is over predicted.

A sample of 80mm x 30mm was cut from the center of a BUD 30 J impacted laminate. The

sample was scanned using MicroCT at 90 KVA. The scanned image shows the formation of a

dent on the impacted side and fiber breakage on the non-impact side as shown in Fig. 18.

6. Conclusion

30
The elastic moduli and strengths, intralaminar (Compact Compression and Compact Tension)

and interlaminar fracture toughness (DCB and ENF) needed for impact simulations of basalt

epoxy composites have been systematically characterized. A single parameter plasticity model

was used to express the incremental stress-strain in basalt/epoxy composite. The single

parameter was determined by performing off-axis test.

A comparison of elastic properties with glass-epoxy composite having similar volume fraction

shows that the two have similar properties except that glass/epoxy composites have higher

transverse strengths in tension and compression because of backing fibers and larger

monofilament diameter.

Impact experiments performed on both basalt and glass laminates shows that peak forces are

higher in case of basalt laminates till fiber breakage (10 J and 20 J). At impact energy of 30 J

fiber failure is dominant in basalt laminates and peak force drops compared to glass laminates.

At higher energy levels fiber failure was observed earlier in basalt compared to glass. Energy

absorption ranged from 34% to 82 % for basalt laminates whereas for glass laminates was from

46 % to 73 %. Damage area was found less at all energy levels for basalt laminates.

A constitutive law for UD composites with provision for intralaminar damage was implemented

as subroutine in FE code Delamination was modeled using cohesive interaction between plies.

The properties determined above were used as input to the FE code. Results of impact

experiments performed on basalt laminates were compared with numerical predictions of force-

time and extent of damage. The delamination was measured experimentally using ultrasonic C-

scan. Force-time curves and central deflections from numerical predictions and experiments

match well.

Acknowledgements

31
Authors are grateful to the NDT group of CSIR-National Metallurgical Laboratory (CSIR-NML)

Jamshedpur, India for their kind permission to use the ultrasonic C-scan facility.

References

1 C. Colombo, L. Vergani, M. Burman, Static and fatigue characterisation of new basalt


fibre reinforced composites. Composite structures 2012;94(3):1165-74.
2 M.V. Donadon, L. Iannucci, B.G. Falzon, J.M. Hodgkinson, S.F. de Almeida, A
progressive failure model for composite laminates subjected to low velocity impact
damage. Computers & Structures 2008;86(11-12):1232-52.

3 Y. Shi, T. Swait, C. Soutis, Modelling damage evolution in composite laminates subjected


to low velocity impact. Composite Structures 2012;94(9):2902-13.

4 L. Raimondo, L. Iannucci, P. Robinson, P.T. Curtis, A progressive failure model for mesh-
size-independent FE analysis of composite laminates subject to low-velocity impact
damage. Composites Science and Technology 2012;72(5):624-32.

5 L. Iannucci, J. Ankersen, An energy based damage model for thin laminated composites.
Composites Science and Technology 2006;66(7-8):934-51.

6 H. Singh, P. Mahajan, Modeling damage induced plasticity for low velocity impact
simulation of three dimensional fiber reinforced composite. Composite Structures 2015;
131:290-303.

7 H. Singh, M. Gupta, P. Mahajan, Reduced order multiscale modeling of fiber reinforced


polymer composites including plasticity and damage. Mechanics of Materials 2017;
111:35-56.

8 W.V. Paepegem, I.D. Baere, J. Degrieck, Modelling the nonlinear shear stress–strain
response of glass fibre-reinforced composites. Part I: experimental results. Composites
science and technology. 2006 Aug 1;66(10):1455-64.
9 C. Bouvet, S. Rivallant, J.J. Barrau, Low velocity impact modeling in composite laminates
capturing permanent indentation. Composites Science and Technology. 2012 Nov
16;72(16):1977-88.

10 T. Czigány, Special manufacturing and characteristics of basalt fiber reinforced hybrid


polypropylene composites: mechanical properties and acoustic emission study. Composites
science and technology. 2006 Dec 18;66(16):3210-20.
11 S. Sfarra, C. Ibarra-Castanedo, C. Santulli, A. Paoletti, D. Paoletti, F. Sarasini, A. Bendada
, X. Maldague, Falling weight impacted glass and basalt fibre woven composites inspected
using non-destructive techniques. Composites Part B: Engineering. 2013 Feb 1;45(1):601-

32
8.

12 M.H. Lapena, G. Marinucci, O.D. Carvalho, Mechanical characterization of unidirectional


basalt fiber epoxy composite. 2014; ECCM.

13 W. Chen, H. Hao, M. Jong, J. Cui, Y. Shi, L. Chen, T.M. Pham, Quasi-static and dynamic
tensile properties of basalt fibre reinforced polymer. Composites Part B: Engineering. 2017
Sep 15;125:123-33.
14 V. Lopresto, C. Leone, I.D. Iorio, Mechanical characterisation of basalt fibre reinforced
plastic. Composites Part B: Engineering 2011;42(4):717-23.

15 S. Carmisciano, I.M. De Rosa, F. Sarasini, A. Tamburrano, M. Valente, Basalt woven fiber


reinforced vinylester composites: Flexural and electrical properties. Materials & Design
2011;32(1):337-42.

16 V. Manikandan, J.W. Jappes, S.S. Kumar, P. Amuthakkannan, Investigation of the effect of


surface modifications on the mechanical properties of basalt fibre reinforced polymer
composites. Composites Part B: Engineering 2012;43(2):812-8.

17 Q. Liu, M.T. Shaw, R.S. Parnas, A.M. McDonnell, Investigation of basalt fiber composite
mechanical properties for applications in transportation. Polymer composites
2006;27(1):41-8.

18 T. Scalici, G. Pitarresi, D. Badagliacco, V. Fiore, A. Valenza, Mechanical properties of


basalt fiber reinforced composites manufactured with different vacuum assisted
impregnation techniques. Composites Part B: Engineering 2016;104:35-43.

19 L.S. Sutherland, C.G. Soares, Impact on low fibre-volume, glass/polyester rectangular


plates. Composite structures 2005;68(1):13-22.

20 E. Sevkat, B. Liaw, F. Delale, B.B. Raju, Drop-weight impact of plain-woven hybrid


glass–graphite/toughened epoxy composites. Composites Part A: Applied Science and
Manufacturing 2009;40(8):1090-110.

21 M. Najafi, S.M.R. Khalili, R. Eslami-Farsani, Hybridization effect of basalt and carbon


fibers on impact and flexural properties of phenolic composites. Iranian Polymer Journal.
2014 Oct 1;23(10):767-73.
22 M.T. Dehkordi, H. Nosraty, M.M. Shokrieh, G. Minak, D. Ghelli, The influence of
hybridization on impact damage behavior and residual compression strength of intraply
basalt/nylon hybrid composites. Materials & Design 2013;43:283-90.

23 R. Petrucci, C. Santulli, D. Puglia, F. Sarasini, L. Torre, J.M. Kenny, Mechanical


characterisation of hybrid composite laminates based on basalt fibres in combination with
flax, hemp and glass fibres manufactured by vacuum infusion. Materials & Design

33
2013;49:728-35.

24 C.S. Kumar, V. Arumugam, H.N. Dhakal, R. John, Effect of temperature and hybridisation
on the low velocity impact behavior of hemp-basalt/epoxy composites. Composite
Structures 2015;125:407-16.
25 J. Andrew, C. Ramesh, Residual strength and damage characterization of unidirectional
glass–basalt hybrid/epoxy CAI laminates. Arabian Journal for Science and Engineering
2015;40(6):1695-705.
26 S. Boria, A. Pavlovic, C. Fragassa, C. Santulli, Modeling of falling weight impact behavior
of hybrid basalt/flax vinylester composites. Procedia engineering 2016;167:223-30.

27 C.T. Sun, J.L. Chen, A simple flow rule for characterizing nonlinear behavior of fiber
composites. Journal of Composite Materials 1989;23(10):1009-20.

28 C. Solutions, Vic-3D Testing Guide. Correlated Solutions. Inc., Columbia, SC. 2010.

29 H. Zhang, Y. Yao, D. Zhu, B. Mobasher, L. Huang, Tensile mechanical properties of basalt


fiber reinforced polymer composite under varying strain rates and temperatures. Polymer
Testing 2016;51:29-39.

30 ASTM Standard, D5528-01, 2001. Test method for Mode I interlaminar fracture toughness
of unidirectional fiber-reinforced polymer matrix composites. West Conshohocken, PA:
ASTM. Google Scholar. 2007.

31 ASTM. Standard test method for determination of the mode II interlaminar fracture
toughness of unidirectional fiber-reinforced polymer matrix composites. ASTM D7905.
2014.

32 S.T. Pinho, R. Robinson, L. Iannucci, Fracture toughness of the tensile and compressive
fibre failure modes in laminated composites. Composites science and technology
2006;66(13):2069-79.

33 M.J. Laffan, S.T. Pinho, P. Robinson, L. Iannucci, Measurement of the in situ ply fracture
toughness associated with mode I fibre tensile failure in FRP. Part I: Data reduction.
Composites Science and Technology 2010;70(4):606-13.

34 Z. Hashin, Failure criteria for unidirectional fiber composites. Journal of applied mechanics
1980;47(2):329-34.

35 A. Puck, H. Schürmann, Failure analysis of FRP laminates by means of physically based


phenomenological models. Composites Science and Technology 2002;62(12-13):1633-62.

36 A.L. Matzenmiller, J. Lubliner, R.L. Taylor, A constitutive model for anisotropic damage
in fiber-composites. Mechanics of materials 1995;20(2):125-52.

34
37 Z.P. Bažant, B.H. Oh, Crack band theory for fracture of concrete. Matériaux et
construction 1983;16(3):155-77.

35
HIGHLIGHTS
Non-linear material characterization of unidirectional basalt-epoxy laminates. Properties
include tensile, compression shear modulus and strengths, interlaminar and intralaminar
strengths; Off-axis tests were performed for evaluating plasticity parameters; These
properties are used as an input to the material model developed in VUMAT for simulating
impact phenomenon based on CDM approach; Low velocity impact tests at 10J, 20J and 30J
were carried out on basalt/epoxy and glass/epoxy laminates for similar thickness and
volume fractions; Damage growth was captured using ultrasonic C-scan and force-time,
deflection time and energy time results gave a good match against the experiments;

You might also like