Anda di halaman 1dari 21

Composites: Part A 40 (2009) 10901110

Contents lists available at ScienceDirect

Composites: Part A
journal homepage: www.elsevier.com/locate/compositesa

Drop-weight impact of plain-woven hybrid glassgraphite/toughened epoxy composites


Ercan Sevkat a, Benjamin Liaw b,*, Feridun Delale b, Basavaraju B. Raju c
a

Department of Mechanical Engineering Technology, New York City College of Technology, 300 Jay Street Brooklyn, NY 11201, USA Department of Mechanical Engineering, The City College of New York, 140th and Convent Avenue, NY 10031, USA c Tank Automotive Research, Development & Engineering Center (TARDEC), US Army RDECOM, Warren, Michigan, USA
b

a r t i c l e

i n f o

a b s t r a c t
The progressive damage behaviors of hybrid woven composite panels (101.6 mm 101.6 mm) impacted by drop-weights at four different velocities were studied by a combined experimental and 3-D dynamic nonlinear nite element approach. The specimens tested were made of plain-weave hybrid S2 glass-IM7 graphite bers/toughened epoxy (cured at 177 C). The composite panels were damaged using a pressure-assisted Instron-Dynatup 8520 instrumented drop-weight impact tester. During these low-velocity simpact tests, the time-histories of impact-induced dynamic strains and impact forces were recorded. The damaged specimens were inspected visually and using ultrasonic C-Scan methods. The commercially available 3-D dynamic nonlinear nite element (FE) software, LS-DYNA, incorporated with a proposed user-dened damage-induced nonlinear orthotropic model, was then used to simulate the experimental results of drop-weight tests. Good agreement between experimental and FE results has been achieved when comparing dynamic force, strain histories and damage patterns from experimental measurements and FE simulations. 2009 Elsevier Ltd. All rights reserved.

Article history: Received 20 September 2008 Received in revised form 22 April 2009 Accepted 23 April 2009

Keywords: A. Hybrid composites B. Impact behavior C. Finite element analysis (FEA) D. Ultrasonic

1. Introduction Metals have always been the rst choice for engineering design since steel was invented. The need of improved materials with high specic mechanical properties, high stiffness-to-weight ratio and high strength-to-weight ratio has led to the rise of composites. Advanced composites have been used in many different applications, especially in the military eld such as aircraft, tanks, bullet-proof body armors, etc. Composites have also been used in civilian aerospace applications, the automotive industry and transportation [1]. Most of the composite structures and parts used in advanced applications are ber-reinforced composites. Thus, a great deal of research has been conducted to understand the mechanical and thermal behaviors of ber reinforced composites. Recently, bers in textile forms have been introduced for use in composites to exploit the advantages of textile structures, such as better dimensional stability, subtle conformability, and deep draw moldability/shapability [2,3]. Textile structural composite products are usually made from high modulus bers or yarns such as glass, graphite, aramid, etc., to qualify for primary and secondary load-bearing applications. Each material has a particular advantage and deciency relative to other materials. For example, graphite exhibits the highest modulus or stiffness but lacks in toughness. For aircraft and other applications
* Corresponding author. Tel.: +1 212 650 5204; fax: +1 212 650 8013. E-mail address: liaw@ccny.cuny.edu (B. Liaw). 1359-835X/$ - see front matter 2009 Elsevier Ltd. All rights reserved. doi:10.1016/j.compositesa.2009.04.028

where the stiffness-to-weight ratio is important, graphite has become the dominant ber. On the other hand, glass ber is the most dependable material for many textile composites due to its lower cost [3]. Since each ber type has a particular advantage, one may ask the following question: why not develop hybrid composites? Recently, there have been extensive investigations of thin-laminate composites made of hybrid ber composites, which incorporate different types of bers in a single composite system. As expected, the mechanical properties of the hybrid composites lie between the single-ber composites made of constituent bers for the hybrid version [4]. During their service life, composites are subjected to various loading conditions. Their response to such conditions should be clearly understood. For instance, a tool can be dropped onto a composite during maintenance or a ying fragment with low velocity can impact the composite structure. The resultant damage of such impacts is usually in the form of delamination. If the impact energy is high enough, ber breakage and matrix failure may also be observed. Due to their high stiffness-to-weight ratio, graphite ber reinforced composites are commonly used for aerospace applications. However, these composites do not have very good impact resistance. Hosur et al. [5] investigated the drop-weight impact response of twill weave carbon fabric and plain weave S2-glass fabric hybrid laminates. The hybrid laminates of 3 mm thickness were impacted at four different energy levels. They concluded that

E. Sevkat et al. / Composites: Part A 40 (2009) 10901110

1091

by providing S2-glass fabrics on the back surface, impact response of hybrid laminates could be enhanced since S2-glass bers have a higher failure strain (that is, better toughness for impact resistance). Hitchen and Kemp [6] studied the effect of ply hybridization and stacking sequence on the low-velocity impact response of carbon-ber composites. Hybridization was achieved using IM7 and standard strength HTA bers in a toughened epoxy 977 matrix system. They reported that hybrid IM7 composites performed signicantly better compared to those strengthened with HTA bers. It was also claimed that by careful selection of the stacking sequence, the hybrid composite could perform similar to that of high-performance ber. The effect of hybrid face sheet on the drop-weight impact response of sandwich composites had also been studied [7]. The low velocity impact response of composites has been studied extensively. In most studies, the composites were impacted using different shape of impactors at various velocities and the resulting damage was evaluated. The effects of various parameters such as the impact velocity [8], impactor shape [9,10], weaving angles between interlacing yarns [11], stitching [12], water immersion aging [13] on the drop-weight impact responses of composites have also been studied. When composites are exposed to low-velocity impact, the resultant damage and residual strength should be examined carefully. Decision has to be made whether the composite can continue to serve or it has to be replaced. A common method to assess the residual strength of the composite is the compression after-impact test [14]. Shim and Yang [15] employed the four-point bending method to determine the residual mechanical properties of woven carbon fabric reinforced composites. However, subjecting specimens to repeated impacts can also be employed to monitor residual strength and damage progression of composites [12]. Closed form solutions for predicting the response of composite plates are mostly for linear cases and are very limited. On the other hand, nite element (FE) simulations have been performed on composites of different shapes, sizes, compositions, loadings and boundary conditions without the expense and time required for ac-

tual testing. For instance, the low-velocity impact responses of composite laminates (e.g., composite sandwich panels and woven composite plates) were simulated using FE methods [16,17]. In this study drop-weight impact tests were rst conducted on hybrid plain-woven glass-graphite bers/toughened epoxy composites with the time-histories of impact-induced dynamic strains and impact forces being recorded. Post-mortem study on the damaged specimens includes ultrasonic C-scanning and photographing under optical microscope. The commercially available 3-D dynamic nonlinear FE software, LS-DYNA [18], in conjunction with a proposed user-dened nonlinear-orthotropic damage model, was then used to simulate the experimental results of the dropweight tests. 2. Experimental procedure The materials used in this research are woven S2-glass/toughened epoxy, woven IM7-graphite/toughened epoxy and woven S2glass-IM7-graphite bers/toughened epoxy hybrid composites. S2glass fabrics (S2-4533 6000) and IM7-graphite fabrics (IM7-GP 6000) are supplied by Hexcel in 2-D plain-weave form. The matrix, SC-79 toughened epoxy resin, which has Part A (Batch number: SC79A012307) and Part B (Batch number: SC79B012507), was supplied by Applied Poleramic. The composite was processed and machined by EDO Fiber Innovations into 101.6 mm 101.6 mm panels. Vacuum assisted resin transfer molding (VARTM) technique was used to stack the plain-woven ll-warp fabrics together (Fig. 1). The specimens were cured at 177C. Fiber volume fraction for all types was 55%. The nal thickness of the composite panels was approximate 6.35 mm. Four different stacking sequences were tested. The rst type contains 37 layers of glass fabrics and was named the GL specimen (Fig. 1a). The second type had glass fabrics outside and graphite fabrics inside. The glass outer face sheets contain nine layers of glass fabrics at each side for a total of 18 layers. The graphite core contains 16 layers of graphite fabrics. This specimen was designated the GL/GR/GL specimen (Fig. 1b). The third type had graphite

Fig. 1. Plain-woven fabrics and the four composite-panel lay-up sequences: (a) glass only (non-hybrid GL), (b) glass skin graphite core (hybrid GL/GR/GL), (c) graphite skin, glass core (hybrid GR/GL/GR) and (d) graphite only (non-hybrid GR).

1092

E. Sevkat et al. / Composites: Part A 40 (2009) 10901110

skins outside and glass core inside. The graphite skins contained 8 layers of fabrics at each side and the glass core had 16 layers. In total 32 layers of fabrics were used. This type was identied as the GR/GL/GR specimen (Fig. 1c). The fourth type, called the GR specimen (Fig. 1d), contained 28 layers of graphite fabrics. Although the thickness was almost constant for all types, the number of fabric layers used in each composite varied due to the different thicknesses of graphite and glass fabrics. Low impact tests at four different energy levels were carried out using a pressure-assisted Instron-Dynatup 8520 instrumented drop-weight impact tester. Fig. 2 shows the schematic drawings of the experimental set-up of the low-velocity impact tests. During all impact tests, a steel mass was attached to the steel impactor with a hemispherical tup of 16 mm diameter for a total weight of 6.15 kg. A velocity of 3.9 m/s was obtained simply by placing the mass at the highest position of the impact tester and dropping it

freely due to gravity. To obtain higher impact velocities, the mass was held rst at the highest position of the impact tester and the pneumatic assistance option was then deployed to increase the impact energy. In this study, various pressure levels were used to obtain impact velocities of 4.4, 4.8 or 6.3 m/s, respectively. As shown in Fig. 2a, two strain gages (one in the ll or longitudinal direction and the other in the warp or transverse direction) were mounted 25.4 mm away from the center of the composite panel, where the impactor would drop upon. These two strain gages are called SG-1 and SG-2, respectively. The specimen was then clamped circumferentially along a diameter of 76.2 mm in a pneumatic-actuated clamping xture (Fig. 2b). Details of the parameters conducted in the drop-impact tests during this study are given in Table 1. During the test, the time histories of impact force and the ensuing dynamic strains were recorded. For post-mortem observation, optical pictures of the front and back surface of the damaged composites were rst taken. The impacted specimens were then scanned using an immersion ultrasonic system. The through transmission technique using a pair of 5 MHz focused and at transducers was employed to detect the internal delamination in the damaged composite panel. Finally, the specimen was machined carefully into two halves along a central line to reveal the damage inside. 3. Experimental results and discussion Various factors (such as lay-up conguration, laminate thickness, impactor size and shape, constituent properties, temperature, impact velocity and energy, etc.) can affect the impact response and damage pattern. Among the four lay-up sequences tested, the GL specimens had the most resistance to impact. As shown in Fig. 3, for impact velocities up to 4.4 m/s only horizontal and vertical front surface cracks were created in GL specimens and no back surface splitting occurred. When the composites were impacted at higher velocity, e.g., 4.8 or 6.3 m/s, in addition to the front surface cracks, some back surface splitting was also observed. The length of the front surface crack increased with increasing impact velocity. As shown in Fig. 4, drop-weight impact tests of the GL/GR/GL hybrid composites created less horizontal and vertical front surface crack compared to the afore-mentioned GL composite specimens. Only impact tests conducted at 6.3 m/s created some penetration and back surface splitting on the hybrid composite. The remaining three tests did not create severe back surface splitting. On the other hand, drop-weight tests conducted on the GR/GL/GR hybrid

Table 1 Parameters of drop-weight impact tests. Lay-up sequence 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 GL GL GL GL GL/GR/GL GL/GR/GL GL/GR/GL GL/GR/GL GR/GL/GR GR/GL/GR GR/GL/GR GR/GL/GR GR GR GR GR Impact velocity (m/s) 3.9 4.4 4.8 6.3 3.9 4.4 4.8 6.3 3.9 4.4 4.8 6.3 3.9 4.4 4.8 6.3 Impact energy (J) 47 60 71 122 47 60 71 122 47 60 71 122 47 60 71 122

Fig. 2. Schematic drawings of the experimental set-up for drop-weight lowvelocity impact tests: (a) the strain-gage-mounted square composite specimen for drop-weight impact test and (b) schematic diagram of the set-up for drop-weight impact tests.

E. Sevkat et al. / Composites: Part A 40 (2009) 10901110

1093

composite specimens generated extended horizontal and vertical front surface cracks for all specimens tested, as depicted in Fig. 5. Back surface splitting was not severe during the rst three tests conducted at velocities up to 4.8 m/s. However, for the impact test

conducted at 6.3 m/s some penetration and back surface splitting was observed. It should be noted that drop-weight impact tests on both groups of hybrid composites were performed under similar conditions. However, the difference in stacking sequence results

Fig. 3. Fractographs of non-hybrid GL composite specimens after drop-weight impact by an impactor of 6.15 kg with a hemispherical tup of 16 mm diameter at various impact velocities.

Fig. 4. Fractographs of hybrid GL/GR/GL composite specimens after drop-weight impact by an impactor of 6.15 kg with a hemispherical tup of 16 mm diameter at various impact velocities.

Fig. 5. Fractographs of hybrid GR/GL/GR composite specimens after drop-weight impact by an impactor of 6.15 kg with a hemispherical tup of 16 mm diameter at various impact velocities.

1094

E. Sevkat et al. / Composites: Part A 40 (2009) 10901110

Fig. 6. Fractographs of non-hybrid GR composite specimens after drop-weight impact by an impactor of 6.15 kg with a hemispherical tup of 16 mm diameter at various impact velocities.

in appreciable differences in the size of the generated horizontal and vertical front surface cracks as well as the severity of penetration and back surface splitting. Finally, as shown in Fig. 6, the GR non-hybrid composite had the least resistance to impact compared to other congurations

tested. All four tests created penetration and back surface splitting. During impact tests at 4.8 and 6.3 m/s, perforation was observed; impactor did not bounce back and was stuck in the composite. When the horizontal and vertical front surface cracks were studied after impact, it was seen that the size of the cracks

20 18 16 14

20

3.9 m/s 4.4 m/s 4.8 m/s 6.3 m/s

18 16 14

3.9 m/s 4.4 m/s 4.8 m/s 6.3 m/s

LOAD (KN)

LOAD(KN)

12 10 8 6 4 2 0 0 1 2 3 4 5 6 7 8 9 10

12 10 8 6 4 2 0 0 1 2 3 4 5 6 7 8 9 10

TIME(ms)

TIME (ms)

(a) non-hybrid GL
20 18 16

(b) hybrid GL/GR/GL


20 18 16 14

LOAD (KN)

LOAD (KN)

14 12 10 8 6 4 2 0 0 1 2 3 4 5 6 7

3.9 m/s 4.4 m/s 4.8 m/s 6.3 m/s

3.9 m/s 4.4 m/s 4.8 m/s 6.3 m/s

12 10 8 6 4 2 0

10

10

TIME (ms)

TIME (ms)

(c) hybrid GL/GR/GL

(d) non-hybrid GR

Fig. 7. Time histories of the contact forces of drop-weight impacts onto composites.

E. Sevkat et al. / Composites: Part A 40 (2009) 10901110

1095

was very similar to that observed for the GL non-hybrid composites. 3.1. Impact force It should be noted that the Dynatup 930-I data acquisition system, accompanying the Instron-Dynatup 8520 instrumented dropweight impact tester, can only measure the initial velocity of the impactor and force vs. time directly. The remaining parameters, such as absorbed energy, velocity of impactor and deection, are calculated using equations of motion. The shapes of the forcetime histories of the non-hybrid GL composites for the rst three tests conducted at velocities up to 4.8 m/s were quite similar, as depicted in Fig. 7a. An almost linear increase of up to 10 KN was followed by an oscillating increase with a decreasing slope. After the maximum point was reached, the force decreased following almost a linear and non oscillating curve and eventually became zero. The entire impact event took about 4.25 ms and the maximum force occurred around 2.5 ms. On the other hand, for the drop-impact

conducted at 6.3 m/s, the forcetime history exhibited a different pattern. The linear increase was followed by an oscillating increase as it was seen in rst three tests until a sudden drop around 19 KN was observed. Knowing that a sudden drop occurred before 2 ms, which was a little earlier than the other three cases when the maximum force was reached around 2.5 ms. Impact forcetime histories of hybrid GL/GR/GL and GR/GL/GR specimens showed quite similar shapes during the rst three impacts up to 4.8 m/s, Fig. 7b and c. The linear increase at the beginning was followed by an oscillating increase and the force gradually returned to zero after reaching a maximum. Just like the non-hybrid GL case, during the impact tests of both hybrid composites at 6.3 m/s, a sudden drop right after the maximum force was observed. During all four tests conducted on the non-hybrid GR specimens, a sudden drops right after the linear increase was observed (Fig. 7d). In these tests, the oscillating increase that was commonly observed during the tests of the other three composite types was not seen. Drops in the forcetime histories happened very early. Maximum force values were very low

25

20

3.9 m/s 4.4 m/s 4.8 m/s 6.3 m/s


ABSORBED ENERGY (J)

140 120 100 80 60 40 20

3.9 m/s 4.4 m/s 4.8 m/s 6.3 m/s

MAXIMUM LOAD (KN)

15

10

0
GL GL/GR/GL GR/GL/GR GR

GL

GL/GR/GL

GR/GL/GR

GR

(a) maximum impact force


70

(b) energy absorption


3.9 m/s 4.4 m/s 4.8 m/s 6.3 m/s

10 9 8 CONTACT DURATION (ms) 7 6 5 4 3 2

3.9 m/s 4.4 m/s 4.8 m/s 6.3 m/s

60

DELAMINATED AREA ( cm

)
50 40 30 20 10

1 0
GL GL/GR/GL GR/GL/GR GR

GL

GL/GR/GL

GR/GL/GR

GR

(c) delaminated area

(d) contact duration

Fig. 8. Comparison of key characteristics of composites subject to drop-weight impact tests.

1096

E. Sevkat et al. / Composites: Part A 40 (2009) 10901110

Fig. 9. A typical energy curve for drop-weight impact test when rebound of the impactor occurs.

140 120 100


ENERGY (J)

3.9 m/s 4.4 m/s 4.8 m/s 6.3 m/s


ENERGY (J)

140 120 100 80 60 40 20 0

3.9 m/s 4.4 m/s 4.8 m/s 6.3 m/s

80 60 40 20 0

TIME (ms)

TIME (ms)

(a) non-hybrid GL
140 120 100
ENERGY (J)

(b) hybrid GL/GR/GL


3.9 m/s 4.4 m/s 4.8 m/s 6.3 m/s
ENERGY (J) 140 120 100 80 60 40 20 0

3.9 m/s 4.4 m/s 4.8 m/s 6.3 m/s

80 60 40 20 0

TIME (ms)

TIME (ms)

(c) hybrid GR/GL/GR

(d) non-hybrid GR

Fig. 10. Energytime histories for the drop-weight impact tests of composite specimens conducted at impact velocities of 3.9, 4.4, 4.8 and 6.3 m/s, respectively: (a) nonhybrid GL, (b) hybrid GL/GR/GL, (c) hybrid GR/GL/GR and (d) non-hybrid GR.

E. Sevkat et al. / Composites: Part A 40 (2009) 10901110

1097

compared to those observed during the impact of the other three composite types. Once the forcetime histories of the drop impact tests and optical pictures of the impacted composites were evaluated, we can conclude that impact tests, which created back surface splitting, exhibited sudden drops at their forcetime histories; whereas the forcetime histories of the tests, which produced no or very little back surface splitting, did not show any signicant drop. The force generated due to the contact of impactor-composite was linear at the beginning, because the impactor was just pushing the specimen down and no damage had been formed yet. As soon as some damage appeared in the composite, the forcetime history started to show oscillations. This damage could be in the form of delamination, front surface cracks or front surface indentation. If the energy of the impactor was high enough, back surface splitting was also observed. The back surface-splitting reduced the resistance (i.e., the out-of-plane stiffness) of the composite appreciably; thus resulting in the sudden drop of forcetime history. If the remaining resistance of the composite was strong enough, the impactor would be pushed back and rebound happened. During the impact of GR at 4.8 and 6.3 m/s, the composite lost its resistance and the impactor was able to move further and perforated the composite. For comparison, the maximum forces created during impact tests are summarized in Fig. 8a. The results indicate that while the GL composite sustained the highest force, the GR composite resisted the lowest. Hybrid composites performed better than the GR composite but the maximum force produced by these composites was a little less than that of the GL composite. The maximum force

increased with increasing impact velocity during testing of the GL and hybrid composites while the maximum forces produced by the GR specimens were almost the same for the four impact tests conducted at various velocities. 3.2. Impact energy Depending on the level of energy, drop-weight impact tests usually result in three distinct interaction modes between the composite panel and the impactor. If the energy absorbed by the composite is very little, the impactor bounces back. If most of the energy of impactor is absorbed by the composite through various modes of damage, no rebound occurs. Finally, in the case of high energy level, perforation can be observed. A typical energy curve for drop-weight impact test where rebound of the impactor occurs is illustrated in Fig. 9. For this study, most of the energytime histories, as shown in Fig. 10, are similar to the one depicted in Fig. 9, because except for the non-hybrid GR case, rebound occurred for all other composites during the impact tests conducted. Rebound was also observed during the impact tests of GR composites at 3.9 and 4.3 m/s impact velocities. The energytime histories of GR composites when impacted at 4.8 and 6.3 m/s, Fig. 10d, were different than those for the other specimens. At 4.8 m/s the impactor penetrated the composite and no rebound occurred. The test at 6.3 m/s created perforation in the composite and no rebound occurred either. Fig. 8b summarizes the comparison of absorbed energies for the above drop-weight impact tests. As indicated in the gure, the GR

Fig. 11. Fractographs of sectioned composites after drop impact conducted at 6.3 m/s: (a) non-hybrid GL, (b) non-hybrid GR, (c) hybrid GR/GL/GR and (d) hybrid GL/GR/GL.

1098

E. Sevkat et al. / Composites: Part A 40 (2009) 10901110

composite always absorbed more energy compared to the other types. Energy absorption of the composites increased with increasing impact velocity. Energy absorption of GL and hybrid composites was very close during impact tests at the same velocity. 3.3. Delamination Fig. 11 shows sectional views of composites, which high-lighted various modes of damage resulted from drop-weight impact at 6.3 m/s. As depicted in Fig. 11a, delamination in the non-hybrid GL composite was very localized on the impact (compression) side and widened on the back (tension) side of the composite. Similarly, delamination in the non-hybrid GR was much localized on the compression side; however, the damage did not exhibit much spreading in the tension side, Fig. 11b. On the other hand, drop-impact tests created quite extended delamination in the hybrid GL/ GR/GL and GR/GL/GR composites, Fig. 11c and d, respectively, especially at the interface between a GL and a GR layer. Both the hybrid composites delaminated more than the two non-hybrid composites. However, splitting that occurred between hybrid layers was more severe in the GL/GR/GL composite compared to that in the GR/GL/GR composite. It is clear that hybrid composites were more susceptible to delamination, especially at the interfaces of GL and GR layers. This may be due to greater incompatibility between dissimilar layers. The ultrasonic C-scans of the impacted composites, as shown in Fig. 12, further prove that hybrid composites delaminated more than non-hybrid composites. Delamination in hybrid composite where GL was used as an outer skin was more pronounced compared to that in the other hybrid composite when GR was used as the face sheet (Fig. 8c). For the two non-hybrid composites, the GL composites delaminated more than the GR composites. Delamination in the GL and the two hybrid composites increased with increasing velocity. It should be noted that during these tests

rebound was observed. Even impacted at 3.9 m/s, severe delamination was observed in the two hybrid composites and the delaminated area changed very little with changing velocity. However, the impact test of 3.9 m/s created small delamination in the nonhybrid GL and increasing velocity resulted in a signicant increase of delaminated area. Finally, delamination in GR did not increase with increasing velocity. Even though there was some increase between the rst and second tests, impact tests conducted at 4.4 and 6.3 m/s created less delamination than those created by 3.9 and 4.4 m/s impacts. This may be due to the occurrence of deep penetration and perforation in the more brittle GR composite when it was impacted at higher velocities: 4.8 and 6.3 m/s. During these two tests, the impactor penetrated into the composite and no rebound was observed. 3.4. Contact duration The contact durations of drop-impact tests are summarized in Fig. 8d. Contact durations produced by the non-hybrid GL composite were shorter than other types of composite. For the two hybrid composites, the GL/GR/GL type produced a little more contact time compared to the GR/GL/GR type; but both contact durations were denitely longer than that for the GL composite. Contact durations increased with increasing impact velocity for the three composites mentioned here. On the other hand, the GR composite exhibited some inconsistency in contact duration. This again can be explained by the inherent brittleness of the non-hybrid GR composite. It should be noted that rebound was observed during testing of the GL and the two hybrid composites. If the deformation of the composite is not severe, the impactor will travel until its entire energy is spent and the impactor bounces back. Increasing velocity will increases energy of impactor and it will travel more to spend the extra energy.

Fig. 12. Ultrasonic C-scan images of four types of composites impacted at various velocities.

E. Sevkat et al. / Composites: Part A 40 (2009) 10901110

1099

450 400 350 300 250 200 150 100 50 0 0 0.01 0.02 0.03 0.04

700 600 500

(Mpa)

(Mpa)

400 300 200 100 0 0 0.005 0.01 0.015 0.02

E1 = E2
60 50
60 50

E1 = E2

(Mpa)

(Mpa)

40 30 20 10 0 0 0,002 0,004 0,006 0,008 0,01

40 30 20 10 0 0 0,002 0,004 0,006 0,008 0,01 0,012

E3
120 100 80

E3
120 100 80 G12 G13,G23

G12 G13,G23

(MPa)

(Mpa)

60 40 20 0 0 0,02 0,04 0,06 0,08

60 40 20 0 0 0,02 0,04 0,06 0,08 0,1

G12 and G13 = G23

G12 and G13 = G23

(a)

(b)

Fig. 13. Nonlinear longitudinal, transverse and in-plane shear stressstrain curves with their slopes representing Youngs and shear moduli, respectively: (a) glass and (b) graphite.

Table 2 Initial engineering elastic constants, Poissons ratios and density of woven glass and graphite bers-reinforced toughened epoxy. E1 = E2 (GPa) Glass/epoxy composite Graphite/epoxy composite 17.04 36.43 E3 (GPa) 6.51 5.18 G12 (GPa) 2.742 3.080 G13 = G23 (GPa) 2.305 1.816

m12
0.13 0.12

m13 = m23
0.413 0.427

q (g/cm3)
1.756 1.458

1100

E. Sevkat et al. / Composites: Part A 40 (2009) 10901110

This will increase the engaging time of the impactor and the composite. However, during the impact tests of GR severe penetration and even perforation was observed when the composite was impacted at higher velocities: 4.8 and 6.3 m/s. Impact test of GR specimens conducted at 4.8 m/s produced very long contact duration compared to other specimens. If impactor penetrates the composite, the contact duration is mainly governed by the friction between impactor and composite. During penetration, bers and the matrix material are either pushed to the sides or ahead at the tip of the impactor. However, the impactor still cannot easily move forward due to the friction forces. Energy spent for friction within a unit time can be less than the energy spent for damaging the composite and that may cause some delays in contact duration. For the impact GR test at 6.3 m/s, contact duration was recorded as 5 ms. However, the impact force did not return to zero at 5 ms. It actually never returned to zero during the data acquisition time allowed by the equipment. This can be explained by the fact that when the data acquisition time ended, the impactor was still moving and the small force seen in the graph was due to the loose contact between the impactor and the hole perforated in the composite. 4. Dynamic damage-induced nonlinear orthotropic FE simulations 4.1. Proposed nonlinear orthotropic damage model The currently available LS-DYNA material models that can be used for composites with progressive damage are (i) ChangChang model (identied as MAT_22 in LS-DYNA), which is valid for an orthotropic unidirectional composite with brittle failure [19,20], and (ii) YenCaiazzo model [21] (designated as MAT_161 and MAT_162), which is based on Hashins failure criteria for compos-

ites [22] for modeling the progressive failure for unidirectional or woven fabric composites. The use of MAT_161 and 162 requires an additional license from Materials Sciences Corporation and is not available for the researchers of this study. While investigating the interlaminar shear strength of plain weave S2-glass/SC79 composites subjected to out-of-plane high strain rate compressive loading, Gillespie and his associates [23,24] reviewed the YenCaiazzo and other composite failure criteria, such as maximum stress and TsaiHill criteria [1]. Naik et al. [25] applied the TsaiHill quadratic failure criterion to study in-plane damage in woven-fabric composites subjected to low-velocity impact. As shown in Fig. 13, our experimental curves exhibited damageinduced nonlinear behaviors in Youngs moduli along the warp and ll directions as well as the in-plane shear stressstrain relations. Such nonlinear characteristics cannot be accounted for in the ChangChang model. In addition, LS-DYNA also contains a nonlinear orthotropic material (MAT_40), which allows inputting experimental curves by piecewise linear segments and does not have any option for damage. However, this model is not recommended by LS-DYNA due to numerical instability [18]. Thus, the authors of this study chose to develop our own user-dened material subroutine, which can be incorporated into LS-DYNA, through MAT_USER_DEFINED_MATERIAL-MODELS (MAT_43). The mechanical properties of the woven glass and woven graphite composites were obtained from tensile tests (Fig. 13). Youngs modulus at longitudinal direction (E1) was assumed to be equal to the transverse Youngs modulus (E2). The out-of-plane Youngs modulus (i.e., E3 in thickness direction) was computed using the rule of mixtures for transverse property: Em Ef ;t . Here Vf is the ber volume fraction and Em and E3 v Em 1vf Ef ;t f Ef,T are the Youngs moduli of the matrix and the ber along the transverse directions, respectively. In this study, the following values: Vf = 55%, Em (SC-79) = 3.06 GPa [23], Ef,T (S2-glass) = 85 GPa

Fig. 14. Finite element simulations for drop-weight impact tests of composite specimens: (a) non-hybrid GL, (b) hybrid GL/GR/GL, (c) hybrid GR/GL/GR and (d) non-hybrid GR.

E. Sevkat et al. / Composites: Part A 40 (2009) 10901110

1101

[26], Ef,T (IM7-graphite) = 12 GPa [27] were used and resulted in: E3 (S2-glass) = 6.5 GPa and E3 (IM7-graphite) = 5.2 GPa. In addition, curves of varying slope, called Ex, of highly nonlinear stressstrain relations were obtained when tensile tests of the specimens with 45 fabric orientations were conducted. Following the ASTM Standard D-3518 [28,29], the in-plane shear stress, strain and modulus were calculated from the relations: x x s12 r , c12 = ex ey and G12 21E . Here rx is the axial stress 2 mxy (fracPA), ex and ey are the axial and transverse strains, respectively e is the associated Poissons ratio obtained in (ey < 0) and mxy ey x the ASTM D-3518 test. It should be pointed out that if we evaluate the in-plane shear modulus using the formula: G12 4 1 11 2c12 [1], the resultant
Ex

E E
1 2

E1

G12 was very close to the one obtained by the ASTM Standard D3518 method. Similarly, shear moduli in the remaining two planes E3 were obtained through the relations: G13 G23 21 m13 . In this study, all Poissons ratios: m12 and m13 = m23 were measured using three strain gages in the longitudinal, transverse and thickness directions, respectively. The initial elastic constants, Poissons ratios and density of the woven glass and woven graphite composites are tabulated in Table 2. Comparing to the E1 = E2 values, the E3 values are much lower and closer to the Youngs modulus of the matrix material, SC-79 resin, indicating that the through-thickness direction is dominated by the matrix stiffness for these unstitched woven composites.

In LS-DYNA, the user-dened subroutines can be written in two formats: scalar or vectorized. In case of a vectorized subroutine; a group of elements is processed at a time whereas for a scalar subroutine, only one element at a time is processed. The failure criteria and erosion of the elements can be programmed in the subroutine. Furthermore, the MAT_ADD_EROSION option can also be used without including erosion in the LS-DYNA code. In our proposed model, when the stresses in all directions of an element reach their respective critical values, which are determined by the experimental curves shown in Fig. 6, that element failed and is eroded from further calculation. In this study, vectorized subroutine was used and erosion was introduced using the MAT_ADD_EROSION option. The orthotropic failure criteria used in our proposed model, which is similar to the damage-induced nonlinear orthotropic model developed by the authors for unidirectional composites and had been proven to be more suitable than ChangChang model for composites with great degree of nonlinearity due to the use of toughened resin [30], are described as follows:

 Longitudinal failure : r1 P S1 tension jr1 j P C 1 compression 1

where r1 is the normal stress and S1 and C1 are the tensile and compressive strengths in the longitudinal (warp) direction, respectively. After longitudinal failure, E1 (Youngs modulus in the longitudinal

Fig. 15. Comparison of FE and experimental impact forcetime histories for composite plates impacted at 3.9 m/s: (a) non-hybrid GL, (b) hybrid GL/GR/GL, (c) hybrid GR/GL/ GR and (d) non-hybrid GR.

1102

E. Sevkat et al. / Composites: Part A 40 (2009) 10901110

direction), m12, m21, m13 and m31 (Poissons ratios in the 12 and 13 planes) are all set to zero.

 Transverse failure : r2 P S2 tension jr2 j P C 2 compression 2

where r3 is the normal stresses and S3 and C3 are the tensile and compressive strengths in the out-of-plane (thickness) direction, respectively. After out-of-plane normal failure, E3 (Youngs modulus in the out-of-plane direction), m13, m31, m23, m32 are all set to zero.

 In-plane shear failure : js12 j P S12

where r2 is the normal stress and S2 and C2 are the tensile and compressive strengths in the transverse (ll or weft) direction, respectively. After transverse failure, E2 (Youngs modulus in the transverse direction), m12, m21, m23 and m32 (Poissons ratios in the 12 and 23 planes) are all set to zero.

where s12 and S12 are the in-plane shear stress and the associated shear strength. After in-plane failure, G12 (in-plane shear modulus in the 12 plane), m12, m21 (Poissons ratios in the 12 plane) are all set to zero.

 out-plane shear failure : js13 j P S13 or js23 j P S23

 Out of plane normal failure : r3 P S3 tension jr3 j P C 3 compression


70 60 50

where s13, s23 and S13, S23 are the out-plane shear stresses and strengths, respectively. After out-of-plane failure, G13 or G23 (outof-plane shear modulus in the 13 or 23 plane), m13, m31, m23 and m32 (Poissons ratios in the 13 or 23 plane) are all set to zero.

EXPERIMENT FEM

ENERGY (J)

40 30 20 10 0 0 1 2 3 4 5 6

TIME (ms)

(a) hybrid GL/GR/GL


70 60 50

EXPERIMENT FEM

ENERGY (J)

40 30 20 10 0 0 1 2 3 4 5 6

TIME (ms)

(b) hybrid GR/GL/GR


Fig. 16. FE calculated internal energies of laminates and comparison of energytime histories of drop-weight tests for two hybrid composites impacted at 4.4 m/s.

E. Sevkat et al. / Composites: Part A 40 (2009) 10901110

1103

In summary, the proposed orthotropic composite damage model can be viewed as the nonlinear version as the prevailing Chang Chang model [19,20]. In this study, S1 = S2 and S3 were determined using Fig. 13a and b, respectively: C1 = C2 values were not determined experimentally; however, we assumed that compressive failure occurs at very high stress. The in-plane shear strength was max . Here rx max represents obtained using the relation: S12 rx2 the maximum axial stress obtained in the 45 tensile tests of the non-hybrid woven composite specimens. Finally, S3, C3 and S13 = S23 are taken from open literature (e.g., [23,24,31]). In summary, the strength values used for woven glass fabric/SC-79 resin

and woven graphite fabric/SC-79 resin used in this study are as follows: woven S2-glass/SC-79 S3 = 0.055 GPa, C1 = C2 ? 1 S1 = S2 = 0.426 GPa, C3 = 0.285 GPa, S12 = 0.0913 GPa, S13 = S23 = 0.05 GPa woven IM7-graphite/SC-79 597 GPa, S3 = 0.055 GPa, C1 = C2 ? 1 S1 = S2 = 0. C3 = 0.285 GPa, S12 = 0.0965 GPa, S13 = S23 = 0.05 GPa and and

12500

7500

EXPERIMENT (SG-1) EXPERIMENT (SG-2) FEM (SG-1) FEM (SG-2)

2500

STRAIN ()

-2500

-7500

-12500

-17500

TIME (ms)

(a) hybrid GL/GR/GL


EXPERIMENT(SG-1) EXPERIMENT(SG-2) FEM (SG-2) FEM (SG-1)

12500

7500

STRAIN ()

2500

-2500

-7500

-12500

-17500

TIME (ms)

(b) hybrid GR/GL/GR


Fig. 17. Comparison of dynamic strain histories obtained experimentally and by FE of hybrid composite plates impacted at 4.8 m/s: (a) GR/GL/GR and (b) GL/GR/GL.

1104

E. Sevkat et al. / Composites: Part A 40 (2009) 10901110

The criterion for delamination between the composite layers is governed by:

 Delamination :

max0; rn NFLS

2

rs 2
SFLS

P1

where rn and rs are the normal and shear stresses acting on the layer interface, respectively, while NFLS and SFLS are the normal and shear strengths of the layer interface, respectively. This criterion was incorporated into LS-DYNA through the command: CONTACT_AUTOMATIC_SURFACE_TO_SURFACE_TIEBREAK. For the interfaces of glass/glass and graphite/graphite, the above experimentally measured S12 values were taken as their delamination shear strengths, SFLS, which are in line with the values reported in open literature [23,24,31]. Since the interface of glass/graphite was more susceptible to delamination, an SFLS of 0.05 GPa was chosen to match the experimental results through FE simulations. In all types of interfaces, p the delamination normal strength is assumed to be: NFLS 3 SFLS. It should be noted that under drop-weight impact, delamination is dominated by shear failure. Indeed, our sensitivity analyses showed that no signicant changes were observed when NFLS was increased to an articially large value. Finally, ERODING_SURFACE_TO_SURFACE contact model was used between the steel impactor and composite. This model allows elements to be eroded, for the purpose of numerical stability, when certain failure criteria are met. In this study strain-based failure criterion was used for element erosion: that is, when e P eerosion,

element was eroded and removed from calculation. In this study, the much higher Youngs modulus and tensile strength of the steel impactor than those of the woven composites enabled us to model it as a rigid body with a density of the steel while eerosion = 0.4 and 0.3 for S2-glass/SC-79 and IM7-graphite/SC-7 woven composites were chosen, respectively. Note that the eerosion values were much higher than the efailure, which occurred at the end points of the stressstrain curves in Fig. 13. As stated above, the associated stiffnesses and Poissons ratios within the range of efailure and efailure were set to zero. 4.2. Validation of FE models 4.2.1. Comparison of forcetime histories Fig. 14 shows the LS-DYNA nite element meshes used to simulate drop-weight impact onto (a) a woven non-hybrid GL composite (b) a woven hybrid GL/GR/GL composite (c) a woven hybrid GR/ GL/GR composite and (d) a woven non-hybrid GR composite. In these meshes, 18 layers for GL, 14 layers for GR and 16 layers for hybrid composites were created, respectively. Fine meshes around the center and course meshes away from center were used. Since the Dynatup 930-I data acquisition system can only measure force vs. time history and the initial impact velocity just prior to impact directly, the experimentally-obtained contact forcetime histories would be the most reliable data for the validation of the proposed FE models. As shown in Fig. 15, the comparison of FE and experimental results of the impact forcetime histories for

Fig. 18. Comparison of post impact damage patterns of FE and experimental results for drop-weight tests for a non-hybrid GR specimen impacted by a 6.15 kg, 16-mm hemispherical-head impactor at (a) 3.9 m/s, (b) 4.4 m/s (c) 4.8 m/s and (d) 6.3 m/s.

Fig. 19. Comparison of post impact damage patterns of FE and experimental results for drop-weight tests for a hybrid GL/GR/GL specimen impacted by a 6.15 kg, 16-mm hemispherical-head impactor at (a) 3.9 m/s, (b) 4.4 m/s (c) 4.8 m/s and (d) 6.3 m/s.

E. Sevkat et al. / Composites: Part A 40 (2009) 10901110

1105

the drop-weight tests conducted at 3.9 m/s impact velocity exhibits very good agreement. As it can be seen from Fig. 15, the FE results are noisier than their experimental counterparts. This may be due to (1) the inherent stiffness nature in nite element method, which tends to generate the ringing responses and (2) the unavoidable ltering characteristics in electronics used for signal measurement and digital data acquisition. The FE forcetime histories of the non-hybrid GL composite showed a slight delay in both damage initiation and contact duration when compared to experimental result, Fig. 15a. FE calculations produced a little longer contact duration for both hybrid

composites, Fig. 15b and c. The FE predicted forcetime histories of the non-hybrid GR composites showed slight mismatch especially during the period following the damage initiation indicated by the apparent sudden drop in Fig. 15d. The predicted contact duration of this test was also shorter than its experimental counterpart.

4.2.2. Comparison of energytime histories The Dynatup 930-I data acquisition system calculates total energy of the impactor/specimen system at time t using the equation:

Fig. 20. Damage progression and stress counters of FE simulation for a hybrid woven GR/GL/GR composite plate impacted at 4.8 m/s (time interval between each plot is 1 ms).

1106

E. Sevkat et al. / Composites: Part A 40 (2009) 10901110

Et K t Pt At K 0

Here E(t), K(t), P(t) and A(t) are the total energy of the impactor/ specimen system, the kinetic and potential energies of the impactor, and the energy absorbed by the specimen at time t, respectively; mV 2 whereas K 0 1 i is the initial kinetic energy, where Vi is the 2 impact velocity measured by the photodiodes used in the experimental set-up. Since during a drop-weight impact event after the initial contact between the impactor and the specimen, the ensuing change in displacement is very small; thus, the change in potential energy can be neglected. Therefore the energy absorbed by the specimen can be computed as follows:

tained internal energies of the hybrid GL/GR/GL and GR/GL/GR composites. 4.2.3. Comparison of straintime histories Fig. 17 shows the straintime histories of the two hybrid composites: GL/GR/GL and GR/GL/GR impacted by the drop-weight at 4.8 m/s. As shown in the gure, the FE-computed straintime histories at the two strain gage locations, SG-1 and SG-2 (Fig. 2a), are in good agreement with those strain signals captured experimentally. The main difference comes from the slight time delay between the FE and the experimental results and the noisy behavior of the FE curves. As stated in Section 4.2.1. Comparison of forcetime histories above, the difference may result from the inherent stiffness of nite element formulation and the resident ltering electronics in experimental set-up. 4.2.4. Comparison of post-impact damage patterns The comparisons of post-impact damage patterns were made for non-hybrid GR and hybrid GL/GR/GL composites, respectively, (Figs. 18 and 19). Impacted specimens were cut into half at impact

At K 0 K t

1 2 mfV 2 i V t g 2

here V(t) is the impactor velocity at time t. Once the LS-DYNA-computed internal energies of all layers in a composite laminate are summed up throughout the laminate, the resultant FE-calculated internal energy can be used to compare the internal energy obtained experimentally. Fig. 16 shows the good agreement between the FE simulated and experimentally ob-

Fig. 21. FE models for the non-hybrid GR (a) and GL (f) as well as the hybrid GL/GR/GL (b, c, d and e) composites with weights and areal densities.

E. Sevkat et al. / Composites: Part A 40 (2009) 10901110

1107

location. Even though just one quarter of the composite was modeled for each FE simulation due to double symmetry, for comparisons in Figs. 18 and 19 half of the composite is shown for convenience. During the impact tests conducted on the GR composite, rebound was observed at impact velocities of 3.9 and 4.4 m/s; perforation was evident at 6.3 m/s impact velocity and neither perforation nor rebound occurred during the impact test at 4.8 m/s. The FE simulations predicted these behaviors very accurately, as depicted in Fig. 18. It should be noted that due to the erosion option used in FE calculations, element was eroded and removed from further calculations once failure criterion was met. However, during experiment, damaged material still existed after impact event. As shown in Fig. 19 for damage-pattern comparison

for the GL/GR/GL composite, the rst three low-velocity impacts did not create signicant damage to the hybrid composite; whereas impact test conducted at 6.3 m/s did create visible damage. These phenomena were all predicted quite accurately by FE modeling. In addition, optical fractographs of the impacted specimens show delamination in all cases of the non-hybrid GR composite and in the 6.3-m/s case of the hybrid GL/GR/GL case. Delamination was especially profound between hybrid layers due to interfacial mismatch. As seen in Figs. 18 and 19, FE model successfully predicted these delamination characteristics for all experiments. Finally, Fig. 20 shows damage progression and stress counters of FE simulation for hybrid woven GR/GL/GR composite plate impacted at 4.8 m/s.

14 12 10 LOAD (KN) 8 6 4 2 0 0 1 2 TIME (ms) 3 4 5

(a) 3.9 m/s


20 18 16 14 LOAD (KN) 12 10 8 6 4 2 0 0 1 2 3 4 5 TIME (ms) 6 7 8 9 10

(b) 6.3 m/s


Fig. 22. Comparison of FEM predicted forcetime histories of non-hybrid GR composite and GL/GR/GL hybrid composites of various thicknesses impacted at (a) 3.9 m/s and (b) 6.3 m/s.

1108

E. Sevkat et al. / Composites: Part A 40 (2009) 10901110

4.3. Effect of thickness on the performance of hybrid composites It is known facts that graphite reinforced composites are attractive for aerospace applications because of their light weight. But it is also a reality that they are vulnerable to impact. One of our goals in this study is to study the effect of hybridization on the impact performance of graphite reinforced composites. The GL/GR/GL hybrid composite that was used in the above combined experimentalnumerical study has glassber reinforced fabric layers of 1.59 mm thickness on each side and 3.18-mm graphite-ber reinforced fabrics as inner layers. The experimental results showed that using glass fabrics as the outer layers improved the impact performance of graphite composites. However, one of the major concerns in aerospace application is weight reduction and the

use of glass fabrics will denitely increase the weight of the composite. Can the performance of the graphite reinforced composite be improved without increasing the weight too much? To answer this question, experiments may have to be conducted on hybrid composites with myriad lay-up sequences. Another approach, which can be used to predict the results for various stacking sequences more economically, is to use validated FE models without going through the costly experiments. As shown above the time histories of contact force, induced strain, absorbed energy and damage pattern obtained from experiments and FE predictions were in very good agreement. Since the proposed FE model is validated, it can now be employed in engineering design. Here three more FE models are created for GL/GR/GL composites with glassfabric layers of 0.4, 0.8 and 1.2 mm in thickness as the outer face

Fig. 23. FEM predicted damage patterns of (1) GR6.35 mm, (2) GL0.397 mm/GR5.55mm/GL0.397 mm, (3) GL0.794 mm/GR4.76 mm/GL0.794 mm, (4) GL1.19 mm/GR3.97 mm/GL1.19 mm and (5) GL1.587 mm/GR3.17 mm/GL1.587 mm composites impacted at (a) 3.9 m/s and (b) 6.3 m/s.

E. Sevkat et al. / Composites: Part A 40 (2009) 10901110

1109

sheets, respectively. The model with 1.6 mm glass fabrics had already been created for the afore-mentioned validation study. Fig. 21 shows the new FE models for the GL/GR/GL composites (see bd) in the gure) as well as the previously created GL/GR/ GL (e), GR (a), GL (f) composites and their corresponding weights. As it is clear from Fig. 21, adding glass fabrics increases the weight. The response of the GL/GR/GL composites with various lay-up sequences were then predicted using FE models. Fig. 22 shows FE predicted forcetime histories of GR and GL/GR/GL composites of various thicknesses. As expected, adding glass fabrics improves the performance of the composite. The predicted forcetime history of the GR non-hybrid composite impacted at 3.9 m/s indicated a sudden drop that represented the creation of damage in the composite following the linear increase. On the other hand, the addition of a single layer of glass fabrics merely delayed the damage; however, adding 2, 3 and 4 layers of glass fabrics could prevent the damage. The maximum force sustained by the composite slightly decreased by adding more glass fabrics. The GL/GR/GL composite with one outer layer of glass fabrics produced almost the same contact duration (about 4.5 ms) as the GR composite. All remaining GL/GR/GL composites produced around a contact duration of 5 ms. Predicted force time histories of GR composite impacted at 6.3 ms showed sudden drops at around 1 ms (Fig. 22b). Experimental forcetime histories of the same test also showed similar behaviors (Fig. 7d) due to perforation observed during this experiment. Adding more layers of glass fabrics onto the GR composite reduced the damage and perforation was prevented. The GL/GR/GL composites with one and two layers of glass fabrics produced around 9 ms contact duration while the remaining composites resulted in a contact duration around 6 ms. Finally, the FE predicted damage patterns shown in Fig. 23 clearly show improvement for the performance of the composite due to the addition of glass fabric layers. Even adding only one layer of glass fabric onto GR composite made signicant improvement, as proven in Fig. 23a-5). Indeed, the predicted perforation of GR impacted at 6.3 m/s was prevented by adding glass fabrics (Fig. 23b). Hence the improvement in drop-weight impact tolerance is clearly evident: Adding more layers of glass fabrics will reduce the damage at the cost of increasing the weight of the composite.

Acknowledgments This work is supported by US Army Research Ofce through Grant No. DAAD19-03-1-00086. Dr. Bruce LaMattina is the Program Manager. The authors also would like to express their thanks for funding from City University of New York through CUNY Collaborative Incentive Research Grants Program (80209-06 10) and CUNY Research Equipment Grant Competition (80212-12 04) for procuring part of the materials and equipment used in this research. References
[1] Jones RM. Mechanics of composite materials. second ed. Philadelphia, Pennsylvania: Taylor & Francis; 1999. [2] Naik KN. Woven fabric composites. Lancaster, Pennsylvania: Technomic Publishing; 1994. [3] Chou TW, Ko FK. Textile structural composites. New York: Elsevier Science Publishing; 1989. [4] Hancox NL. Introduction to bre composite hybrids. In: Hancox NL, editor. Fibre composite hybrid materials. London: Applied Science Publishers; 1981. [5] Hosur MV, Adbullah M, Jeelani S. Studies on the low-velocity impact response of woven hybrid composites. Compos Struct 2005;67:25362. [6] Hitchen SA, Kemp RMJ. Development of novel cost effective hybrid ply carbonbre composites. Compos Sci Technol 1996;56:104754. [7] Gustin J, Joneson A, Mahinfalah M, Stone J. Low velocity impact of combination Kevlar/carbon ber sandwich composites. Compos Struct 2005;69:396406. [8] Sevkat E, Liaw BM, Delale F, Raju BB. Experimental and numerical studies of S2-glass ber/toughened epoxy composite beams subject to drop-weight or ballistic impact. In: 2007 ASME international mechanical engineering congress and exposition, Paper No. IMECE2007-43966. 2007. [9] Mitrevski T, Marshall IH, Thomson R, Jones R, Whittingham B. The effect of impactor shape on the impact response of composite laminates. Compos Struct 2005;67:13948. [10] Sevkat E, Liaw BM, Delale F, Raju BB. Drop-weight impact responses of woven hybrid glass-graphite/toughened epoxy composites. In: 2008 ASME international mechanical engineering congress and exposition, Paper No. IMECE2008-68835. 2008. [11] Atas C, Liu D. Impact response of woven composites with small weaving angles. Int J Impact Eng 2008;35:8097. [12] Hosur MV, Karim MR, Jeelani S. Experimental investigations on the response of stitched/unstitched woven S2-glass/SC15 epoxy composites under single and repeated low velocity impact loading. Compos Struct 2003;61:89102. ska K, Guillaumat L. The effect of water immersion ageing on low[13] Imielin velocity impact behaviour of woven aramidglass bre/epoxy composites. Compos Sci Technol 2004;64:22718. [14] Davies GAO, Hitchings D, Zhou G. Impact damage and residual strengths of woven fabric glass/polyester laminates. Composites: Part A 1996;27A:114756. [15] Shim VPW, Yang LM. Characterization of the residual mechanical properties of woven fabric reinforced composites after low-velocity impact. Int J Mech Sci 2005;47:64765. [16] Karger L, Baaran J, Temer J. Efcient simulation of low-velocity impacts on composite sandwich panels. Comput Struct 2008;86:98896. [17] Iannucci L. Progressive failure modelling of woven carbon composite under impact. Int J Impact Eng 2006;32:101343. [18] LS-DYNA 971, Livermore Software Technology Corporation (LSTC), Livermore, CA. <http://www.lstc.com>. [19] Chang FK, Chang KY. Post-failure analysis of bolted composite joints in tension or shear-out mode failure. J Compos Mater 1987;21:80933. [20] Chang FK, Chang KY. A progressive damage model for laminated composites containing stress concentrations. J Compos Mater 1987;21:83455. [21] Yen CF, Caiazzo A. Innovative processing of multifunctional composite armor for ground vehicles. ARL-CR-484, US Army Research Laboratory, Aberdeen Proving Ground, Maryland. 2000. [22] Hashin Z. Failure criteria for unidirectional ber composites. J Appl Mech 1980;47:32935. [23] Gillespie Jr JW, Gama BA, Cichanowski CE, Xiao JR. Interlaminar shear strength of plain weave S2-glass/SC79 composites subjected to out-of-plane high strain rate compressive loadings. Compos Sci Technol 2005;65:1891908. [24] Xiao JR, Gama BA, Gillespie Jr JW. Progressive damage and delamination in plain weave S-2 glass/SC-15 composites under quasi-static punch-shear loading. Compos Sci Technol 2007;78:18296. [25] Naik NK, Sekher YC, Meduri S. Damage in woven-fabric composites subjected to low-velocity impact. Compos Sci Technol 2000;60:73144. [26] Hexcel technical fabrics handbook, reinforcements for composites, March 2009. [27] Cox BN, Flanagan G. Handbook of analytical methods for textile composites, NASA CR 4750, March 1997. [28] Adams DF, Carlsson LA, Pipes RB. Experimental characterization of advanced composite materials. 3rd ed. Boca Raton, FL: CRC Press; 2003.

5. Conclusions Some conclusions can be drawn based on the above study.  Among the four lay-up sequences tested, GL composites had the most, while GR had the least resistance to impact. Both hybrid specimens performed better than GR but not as well as GL. Between the two types of hybrid composites, the glass-skin/ graphite-core type performed a little better than the graphiteskin/glass-core type.  The process of hybridization gives us control of certain parameters such as impact force and damage to the composite. However, hybrid composites are prone to delamination, especially between dissimilar layers.  Nonlinear orthotropic user-dened damage model was successfully implemented into LS-DYNA. Dynamic force, strain and absorbed energy histories and post-impact damage patterns obtained from experiment and FE simulations were in good agreement.  The FE predicted forcetime histories and damage patterns showed that in order to reduce impact damage, surface of GR composites can be covered by glass fabrics. Adding more fabrics improved the performance of the GR composite; but this results in weight increase.

1110

E. Sevkat et al. / Composites: Part A 40 (2009) 10901110 [31] Reinhart TJ, et al. High-strength medium-temperature thermoset matrix composites. Composites, engineered materials handbook, vol 1. Metals Park (OH): ASM International; 1987.

[29] Pipes RB, Blake RA, Gillespie Jr JW, Carlsson LA. Test methods, delaware composite design encyclopedia. In: Carlsson LA, Gillespie Jr JW editors. vol. 6. Lancaster, PA: Techonomic Publishing; 1990. [30] Sevkat E, Liaw BM, Delale F, Raju BB. A combined experimental and numerical approach to study ballistic impact response of S2-glass ber/toughened epoxy composite beams. Compos Sci Technol 2009;69:96582.

Anda mungkin juga menyukai