Você está na página 1de 19

IMPACT LOADING IN COMPOSITES

SPECIAL PROBLEMS
AE 8900
DR. KARDOMATEAS

APRIL 28, 2015


MANSI AGRAWAL
ABSTRACT

Composite materials have emerged as the materials of choice for increasing the performance and reducing the
weight and cost of military, general aviation, and transport aircraft and space launch vehicles. Major
advancements have been made in the ability to design, fabricate, and analyze large complex aerospace
structures. The recent efforts by Boeing and Airbus to incorporate composite into primary load carrying
structures of large commercial transports and to certify the airworthiness of these structures is evidence of the
significant advancements made in understanding and use of these materials in real world aircraft.

NASA has been engaged in research on composites since the late 1960’s and has worked to address many
development issues with these materials in an effort to ensure safety, improve performance, and improve
affordability of air travel for the public good.

In addition to good quasi-static performance different structures have dynamic impact requirements. For a road
vehicle this might be crash worthiness, an aircraft or spacecraft it can be the ability to bird strikes or space debris
impact and a naval ship needs to be protected against blast or ballistic loading. In this report important aspects
of dynamic loading of composite and sandwich structures are addressed. The current state of the technology is
reviewed and key issues requiring additional research identified. Also, grand challenges to be solved for
expanded use of composites in aero structures are identified.
INTRODUCTION
Composites are materials that are comprised of strong load-carrying materials (known as reinforcement) imbedded in a
weaker material (known as matrix). Reinforcement provides strength and rigidity, helping to support structural load. The
matrix, or binder, maintains the position and orientation of the reinforcement, balance loads between the reinforcements,
protect the reinforcement from environmental degradation, and provide shape and form to the structure. [1]

The three most common types of reinforcing fiber are fiberglass, carbon and Aramid (Kevlar). Glass fibers are the heaviest,
have the greatest flexibility and the lowest cost. Aramid is a moderate cost and stiffness fiber and is the lightest weight.
Carbon is moderate to high in price, slightly heavier than Kevlar but lighter than glass, and features certain varieties that
have exceptionally high stiffness. Hybrids are composites with more than one reinforcing material. Advanced composite
is a term used to describe composites reinforced with very high performance fibers such as carbon and Kevlar. Fiberglass
is usually included.

In addition to general stiffness and strength requirements, one of the important factor affecting composites is impact
behavior. A car needs to be able to protect the passengers in the event of a collision. Aircrafts need to be able to withstand
impact loading events when subjected to bird strikes or hail storms. In an impact event, the material and the structure are
deformed at higher rates which potentially could give a completely different behavior compared to a quasi-static (low
speed) loading scenario. [2]

The energy absorbed during impact process is often very large. That energy is mainly dissipated by a combination of matrix
damage, fibre fracture and fibre-matrix debonding. These facts lead to the significant reductions in the load-carrying
capabilities in such structures. In ballistic impacts (short contact between impactor and target) the damage is localized
and clearly visible by external inspection, while low velocity impact involves long contact time between impactor and
target, which produces global structure deformation with undetected internal damage at points far from the contact
region [4].

Studies conducted previously

In their study Karakuzu et al. investigated the stacking sequence and thickness effects on the impact behavior of the
glass/epoxy composite plate at equal energy (40 J), equal velocity (2m/s) and equal impactor mass (5kg). They chose five
different orientations as [0/±15/90]s, [0/±30/90]s, [0/±45/90]s, [0/±60/90]s and [0/±75/90]s and three different plate
thickness as 2.9mm, 5.8mm and 8.7 mm. 3DIMPACT code was used as the solver and to verify it’s accuracy experimental
studies were carried out with the [0/30/60/90]s orientation which was subjected to impactor with 20J impact energy and
5 kg mass and velocity of 2.828 m/s [5]. The comparison is shown in Figure 1 and it can be seen that the two curves are
close to each other.
20J
2m/s

Figure 1: Contact force-contact time Figure 2: Delamination area for [0/30/60/90]s


history for [0/30/60/90]s laminate [5] oriented plates [5]

Figure 2 shows the delamination area for [0/30/60/90]s oriented plates subjected to 20J impact energy at 2 m/s velocity
and 10 kg mass and their superimposition. We can see that it is pretty similar to the experimental result. After an
agreement between the two results, numerical analysis was carried out for various cases and following conclusions were
drawn by Karakuzu et al.

1. The contact force increases by increasing ±θ at equal energy, equal mass and equal velocity.
2. The lower impactor mass with higher impact velocity causes greater contact forces. However, the lower impact
energy with lower impact velocity and lower impact energy with lower impact mass cause lower contact forces.
3. The lower mass with higher velocity causes higher deflection while the lower energy with lower velocity and lower
mass with lower energy cause lower deflection
4. Higher plate thickness cause higher contact force and lower deflection. But the contact force and deflection rates
decrease by increasing the plate thickness.
5. The overall delamination are increases by increasing impact energy. However, it does not significantly change by
increasing the fiber orientation.
6. The overall delamination area decreases by increasing the plate thickness.
Figure 3 and 4: Maximum contact force and maximum deflection variation with different orientation for 2.9 mm thickness at
a) equal energy (40 J) b) equal mass(5 kg) and c) equal velocity(2 m/s).

Figure 5 and 6: Variation of maximum contact force and maximum deflection by plate thickness at a) 5 kg mass and 4m/s velocity (40
J) and b) 5 kg mass and 2m/s velocity (10 J).[5]
Figure 7: Overall delamination areas at equal energy (40 J) for a a) 5 kg, b) 10 kg, c)15 kg and d) 20 kg [5]

Figure 8 and 9: Overall delamination areas at equal impactor mass (5 kg) and at equal velocity (2 m/s) for a) 10 J, b) 20 J, c)
30 J and d) 40 J[5]
B. Z. Jang et al. made an effort to study the impact response of the fabric, fabric-epoxy single ply, and corresponding
five-ply laminate samples. Prevailing deformation and fracture mechanisms in these materials under various loading
conditions were studied, with the major material parameters dictating such mechanisms identified. The effects of fiber
hybridization on the mechanical behavior of composites were also investigated. Impact behavior was assessed using a

Dynatup Model 730 drop weight facility manufactured by the General Research Corp. Weights collected in a box with an
impacting nose tip of 1.27cm diameter, were dropped through 1.5 m to hit the laminate at the center of the span. The
weights and, occasionally, the drop height as well were varied to give a range of incident energies up to 110 J. The load
and energy at a maximum load point (end of damage initiation phase) are symbolized by Pm and Em. With Et representing
the total energy absorbed by the specimen in complete penetration, the energy absorbed in the propagation phase will
be Ep = Et – Em [2]. The impact resistance of a composite is determined by the total energy dissipated in the material
before final failure occurs. The total energy absorbed by the material is the sum of the shock or dynamic wave energy
dispersed and the energy consumed during plastic deformation, plus the energy needed for creating new surfaces. The
latter includes delamination and through-the-thickness cracks. The microfailure mechanisms possibly operating during
impact loading include matrix cracking, fiber-matrix debonding, fiber breakage, and fiber pull-out.[2]

It can be seen from Table 2 that the impact toughnss of fiber composites is the following descending order:
1. PE, 2. Polyester, 3. Nylon, 4.Glass, 5. Aramid and 6.Graphite
Also it can be seen that maximum load implies high total energy absorbtion level. Except for the aramid composite, the
[2]

ductility index [D.I.) of composites decreased with an increasing total impact energy. This suggests that, for tougher
laminates, more energy will be dissipated before the maximum load is reached than that afterwards. This observation
again points to the direction that tough fibers are essential to achieving composite impact toughness.

The impact properties of single-layer fabric samples are summarized in Table 3. When fully impregnated with epoxy
resin, most single-layer composites (lamina) behaved quite differently than the corresponding bare fabric specimens. As
summarized in Table 4, adding epoxy resin dramatically reduced the maximum load and total energy of either graphite
or Kevlar fabric in impact. In the case of Kevlar fabric a 4 times reduction in strength and 11 times reduction in total
energy absorbed were measured. The Kevlar lamina was through penetrated without exhibiting any deflection. On the
[2]

contrary, the impact properties of both polyester and PE monolayers were significantly enhanced by the epoxy resin. As
in the case of bare fabric, the polyester lamina was perforated. But the lamina exhibited a moderate deflection and
extensive fiber-matrix interfacial debonding. The PE lamina also showed extensive deformation and visible debonding
phenomenon. But no through penetration or fiber breakage was observed in the impact loaded PE lamina. Both plastic
deformation and interfacial debonding seemed to be effective energy absorbing mechanisms. The PE based materials,
either a layer of fabric alone or with epoxy impregnation, undoubtedly top this group of materials studied in both
strength and ductility rating. [2]
The following conclusions were reached by B. Z. Jang et al.:[2]
1. The high-strength Polyethylene fibers have been shown to absorb large amounts of energy prior to failure of
composites. These fibers can be used in hybrids to improve the impact resistance of various composite materials.
2. A large fiber modulus or a great fiber failure strain alone does not provide adequate fiber toughness for improved
impact resistance of composites. The fibers must have both high strain and reasonably high modulus (and strength). PE
fibers appear to have such needed characteristics. The matrix should be of comparable ductility to ensure that the
composite strain does not become a limiting factor in constraining the stretching of fibers.
3. A myriad of failure mechanisms in hybrid laminates can occur during impact loading. Laminates with an alternating
stacking sequence usually exhibit a combination of through penetration and delamination, the latter being in the dagger
shape and visible from both sides. For unsymmetric hybrid laminates containing two or three layers of PE fabric, the
impact energy depends on which side facing the impact direction. In general, this unsymmetric configuration possesses
better impact resistance than its alternating-sequence counterpart. In most cases, failure in these materials involved
perforation, delamination and some tearing of the non-PE layers followed by deflection of PE layers provided PE side
faced the impact direction. When the non-PE side was struck first, these non-PE layers were perforated with a lesser
degree of delamination created. With some exceptions, this process was followed by through penetration of PE layers,
leading to an inferior energy absorbing capability.
4. Except for the aramid-based hybrid laminates, the ductility index decreases with increasing energy absorbing
capability of a composite, implying that, for tougher composites, more energy will be dissipated to reach the maximum
load than afterwards. This observation again suggests the significance of fiber properties in controlling the strength and,
therefore, the impact resistance of composites.

Geoffrey Roy Goodmiller in his studies has worked on composite patch performance under low-velocity impact loading.
Fiber reinforced polymer (FRP) composite patches are adhesively bonded to a metallic structure as a repair method to
either restore the load carrying capacity of a damaged structure or to increase damage tolerance in an undamaged
structure as reinforcement. The adhesive transfers loads between the two and provides a stiff connection due to its
large area for load transfer, despite its own relatively low stiffness. [6]
He says that low-velocity impacts on the adhesive between the composite patch and metal can be a possible initiator of
disbond. As delaminations in composites tend to occur between plies of differing orientations due to interlaminar shear
stresses, so might disbonds be expected to occur in a brittle adhesive between the composite and metal substrate due
to their differing impact behaviors. Baker [6] found that the structural adhesives typically used in aircraft repair have
significantly higher toughness 18 than most composites and so fatigue failure often nucleates at the surface of the
composite and not within the adhesive layer. Denney and Mall [8] found the location of the disbond is important to find
its effect on patch performance; disbonds occurring over cracks resulted in greater crack growth and fatigue damage
than disbonds occurring a distance away from cracks. Al-Zubaidy et al. [9] conducted experiments on the bond strength
between CFRPs and steel with epoxy adhesive under impact tensile loading; their research indicates that bond strength
increases with increasing loading rates until offset by the onset of delamination within the CFRP. Fatigue loading is also
reason of disbond propagation in adhesively bonded joints.[7]

The Hashin plane stress failure criteria in Geoffrey’s research are as follows [7]:

Table 5[7]
In these equations, Xt and Xc represent the longitudinal tensile and compressive strength, Yt and Yc represent the
transverse tensile and compressive strengths, and SL and ST represent the longitudinal and transverse shear strengths. σij
represents the effective stress in the corresponding direction.
Lapczyk introduced damage variables, dtf , dcf , dtm , and dcm, whose values range from 0 at or before damage initiation in
a given mode to 1 at complete failure in that mode. These values are combined to determine damage variables for matrix,
fiber, and shear (dm , df , and ds ):

These damage variables are then used in damage operator, M:

This matrix is applied to the true stress tensor, σ, to determine the effective stress tensor, σ :̂

If the material is completely undamaged, M is equal to the identity matrix. The effective stress tensor is used by Abaqus
in calculating the Hashin damage initiation criteria, as denoted in Equations 1 to 4.
To compare Numerical and experimental results, Geoffery used a finite element model consisting of 9 separate plies, each
0.44 mm thick, with a 0.01 mm thick cohesive layer sandwiched between each ply. Plies were modeled with SC8R: 8-node,
quadrilateral, reduced integration, continuum shell elements, with enhanced hourglass control, a maximum degradation
of 85%, and a Hashin damage viscous stabilization factor of 1x10-7 in order to prevent excessive element distortion. The
ply material studied was E-glass/epoxy; the material properties were previously provided in Table 5. At the center of the
mesh, elements are 0.5 mm x 0.5 mm; their size increased with distance from the impact zone in order to reduce
computational demand. The interlaminar layers were modeled with COH3D8, 8-node cohesive elements, with a regular
mesh of 0.5 mm x 0.5 mm, as determined by the interlaminar damage validation. Maximum degradation for these
elements was set at 99% and linear bulk viscosity was set at 0 as suggested in the Abaqus user manual. These cohesive
layers were tied to the ply layers using surface-to-surface ties. The impactor was modeled as a rigid body with a mass of 2
kg with a velocity of 4.472 m/s, providing initial impact energy of 20 J. Contact between the impactor and composite
layers, as well as contact between plies suffering delaminations, was modeled with a friction penalty of 0.65, based on
friction between composite parts. The plate was given fixed boundary conditions along the bottom and top edges,
representing the circular clamps, while the motion of the impactor was restricted to only the normal dimension. The
impact simulation was run in Abaqus/Explicit, with a time span of 0.005 seconds. The linear bulk viscosity parameter was
set at the recommended value of 0.06, and the quadratic bulk viscosity parameter was the recommended value of 1.2. [7]

A comparison of results from the FEA with experimental data showed good agreement between the two in terms of
absorbed energy and contact force. Figure 10 graphs absorbed energy vs. time for numerical and experimental results.
Acceleration due to gravity was not modeled; the additional energy related to the conversion of potential energy due to
the change in position of the impactor and composite during deflection was neglected. Factors which might have led to
higher final levels of absorbed energy included delamination in the composite beyond the clamped section, which was not
modeled, and vibrations in the testing apparatus. Figure 11 compares the contact force in the FEA and experimental data.
The Hertzian force, corresponding to the first onset of delamination, was experimentally determined to be 1370.5 N; FEA
gave a result of 1725.7 N. The maximum contact force, measured at 5170.4 N, was predicted by FEA to be 5602.9 N. A
major difference in the graphs of the contact force was the much larger drop after maximum contact force during
numerical analysis. This behavior corresponded to massive intralaminar damage occurring at this time. [7]

Figure 10. Comparison of absorbed energy for validation-20 J [7] Figure 11. Comparison of contact force for validation-20 J [7]
Composite materials are increasingly used for primary load bearing components in aerospace applications. A major
concern for their wider use is the potentially catastrophic effect on the material of impact induced damage. This particular
form of damage can often be difficult to detect by visual inspection (e.g. Barely Visual Impact Damage) and frequently
gives rise to internal or back face damage in the form of delamination and matrix cracking. J. W. C et al. develop a novel
composite system which employs a biomimetic approach to perform a self-repairing function. Such a system can perform
two functions; the visual enhancement of impact damage by the bleeding action of a highly conspicuous medium such as
fluorescent dye, and the restoration of mechanical properties by a healing agent, stored within hollow fibres, infiltrating
the damaged area and acting to ameliorate the effect of the damage. Impact indentation followed by four-point bend
flexural testing was conducted to evaluate the strength restoration after self-repair. The results of mechanical testing have
shown that a significant fraction (∼97%) of strength is restored by the self-repairing effect.[11]

The indentation in the virgin specimens and repair agent infiltrated specimens showed no significantly different load-
displacement response, Figure 12. A distinctive characteristic of the behaviour is an initial rise in load followed by two
sudden falls at approximately 750 and 1050 N (denoted A and B in Figure 11) during the indentation process. These
slight reductions in load correspond to damage progression through the thickness of the laminate in the form of
delamination, matrix cracking and hollow fibre fracture occurring at the upper (A) and lower (B) solid/hollow ply
interfaces.

Fig. 12. [11]: Typical load–displacement curves for indentation on infiltrated


and uninfiltrated specimens. Two transient load events (at points A and B)
occur around 750 and 1050 N.

Interface delamination, matrix cracking and hollow fibre fracture are shown in Figure 13, which shows a cross-section
through an uninfiltrated specimen (Group A) after indentation. The majority of the impact induced damage is localised
within or adjacent to the hollow fibre plies, thus creating an ideal situation for self-repair by the uncured resin within the
hollow fibres. The fracture of hollow fibres in the 0 and 90° plies and the mixing of resin and hardener allows initiation of
the curing process whilst simultaneously promoting infiltration of the local matrix cracks and delamination by capillary
action. A key aspect of this whole self-healing process is that the impact energy must be of a sufficient threshold value to
fracture hollow fibre plies. This threshold value can be tailored for any application by the constituents, number and
positioning of the repair agent bearing layers within the laminate stack.[11]
Figure 13. Optical micrograph of cross-section through impact damaged hybrid solid
glass/hollow glass/epoxy laminate.[11]

Figure 14 and table 6 show the results of flexural testing of the four specimen groups. Specimen groups A and B
effectively control as neither have been subjected to resin infiltration and in addition, group B specimens have not been
subjected to indentation. It is clear that impact has a serious effect on flexural strength, as specimen group A (damaged,
uninfiltrated) shows an approximate 14% lower value compared to group B (undamaged, uninfiltrated). These two sets
of results effectively provide the lower and upper bounds, respectively, against which the self-repairing specimens can
be compared.

Figure 14: Average flexural strength of the four groups of specimens


tested.[11]

Table 6. Results from four-point bend flexural testing


Modulus of rupture
Specimen No. of MOR standard Failure Weibull
(flexural strength), σf
group specimens deviation (MPa) strain (%) modulus (m)
(MPa)
A 8 548.1 50.7 2.7 8.2
B 8 623.9 6.0 3.1 94.7
C 8 603.6 19.4 3.1 27.4
D 8 554.8 50.2 2.5 9.3

Specimen group D had been subjected to a healing regime of 1.5 h at 40 °C immediately after indentation which has
restored only 1% of flexural strength compared with group A. However, specimen group C, which were subject to 24 h at
room temperature immediately after indentation show a regain of 10% in strength compared to group A and also
compare very favourably with the results for group B, being only marginally lower. This suggests that slower (24 h)
curing of the healing agent allows a much more effective repair to take place compared to the temperature accelerated
regime. This is probably attributable to a more extensive penetration of any damage crack paths (a time-dependent
process) by the repair resin before the viscosity rise associated with cure progression precludes this process.
In order to more fully appreciate and validate the effect of the self-repairing process on restoring mechanical strength,
Weibull analysis was undertaken on the results from the four-point bend flexural testing. The Weibull modulus, m, can
be used to define the shape of the failure distribution curve and the width of the probability distribution. In Figure 15,
specimens group B (undamaged, uninfiltrated) shows the highest value of Weibull modulus (94.69), indicating a narrow
distribution of failure strength. After indentation of identical material (specimen group A), a significant decrease in
Weibull modulus to 8.21 is observed. For both repaired specimen groups (C and D), the Weibull modulus showed an
increase (compared to group A), with the highest value being observed for group C (27.38). This suggests that the self-
repair process is particularly effective at removing the scatter in strength which arises as a result of the distributed
damage associated with an impact event.

Figure 15: Weibull distribution of the four-point bend flexural testing for the four groups of
specimens.[11]

A specimen similar to those from group C and D, where the hollow plies contain an UV fluorescent dye and resin repair
agent, was subjected to flexure until failure. The resulting bleeding of dye from within the fractured hollow fibres into
the damaged area was visually observed under an UV light source, Figure 16, thus demonstrating the effectiveness of
the bleeding action as a method for damage visualisation.

Fig. 16. Fracture of a hybrid solid/hollow fibre reinforced plastic showing bleeding of UV fluorescent dye (Ardrox 985
dye penetrant) along crack paths. (×45 magnification).[11]

Figure 17 a shows a conventional image of a test specimen from group C after impact damage via indentation and
flexural testing. External surface damage as a result of the failure in flexure can be clearly seen. Figure 17 b shows the
same specimen irradiated under UV light. The extent of internal damage is now more clearly visible. The bleeding action
of the dye into the numerous cracks and fissures created by the initial indentation serves to decorate any damage sites,
increasing their ease of detection. This could be of particular benefit, where rapid visual inspection of large surface areas
(e.g. wing skin panels) is required. Work is currently underway to quantify and correlate the visible damage area with
the energy of impact.
Figure 17:(a) Conventional image of impact damage in a hybrid solid/hollow glass fibre reinforced plastic
specimen showing surface damage (×18 magnification). (b) Same impact site as shown in (a) irradiated with UV
light highlighting internal intra and interlaminar damage. (×20 magnification).

The results of flexural testing have shown that for the lay-up investigated, a significant fraction of lost mechanical
strength is restored by the self-repairing effect of a repair agent stored within hollow fibres. The ‘self-repair’ is
dependent upon uncured resin (in the 0° plies) combining with the hardener (in the 90° plies) as a result of fibre fracture
in both these layers. This self-repairing mechanism is not proposed as a permanent measure to eradicate the effects of
damage within a composite but to provide a means to inhibit further damage propagation.[11]

M. V. Hosur et al.[12] in their study investigated repair strategies for thick section woven S2-glass/vinyl ester composites
subjected to ballistic impact loading. Three different schemes were adopted to repair the panels. Different steps
involved in the repair include:
1.evaluation of damage (visual and ultrasonic);
2.removal of damage;
3.condition the surface of repair (chemlok 7700);
4.cutting patches (of different shape and size);
5.layup of repair patches (vinyl-ester 411-350);
6.curing;
7.assessment of repair (ultrasonics).

While most of the steps outlined are similar in the three methods, step #5 shows certain deviation in each method. In
the first method, repair is carried out by means of patching which is the most recognizable repair method in use today.
For purpose of the current research the patching method is extended to include multi-layered patching. In the second
method, combination of fiber fragment and resin is used. In this process the resin system is reinforced before
application by fibers to strengthen the repair. Third method of repair consists of the insertion of a single layer lamina at
different depth levels along with fabric patches. All the laminates repaired were inspected by ultrasonic C-scan
techniques to evaluate the effectiveness of the repair.[12]

Three of the panels repaired using the three methods described above were subjected to ballistic impact using FSP
projectiles. The results of post-repair impact tests are presented in Table 7. Also included in the table is the ballistic limit
of the same samples tested in their virgin state. It can be seen that the post-repair ballistic limit for the three samples as
compared to that of the samples in their virgin state is about 25%, 112% and 64% respectively, which gives the
qualitative indication of the effect of the three methods. From, this it can be inferred that the methods 2 and 3 are very
effective in restoring the integrity of the laminate. Projectile clearly penetrated the sample 15 (repair method #1) with
some of the patches still intact, whereas for the samples #19 and #17 the repair patches were debonded. Hence, the
energy absorbed was higher for samples #19 and #17.[12]

Table 3. Post-repair impact properties[12]

First impact Impact after repair


S. no.

Absorbed energy (J) Projectile type Absorbed energy (J) Projectile type

15 427 FSP 106 FSP

19 547 Cylindrical 614 FSP

17 433 Conical 279 FSP

To evaluate the static performance of the repaired panels, three-point flexural tests were carried out. For each repair
method, rectangular samples were cut from the repair portion and also from the undamaged region. Samples from the
repair section were tested to evaluate the effectiveness of the repair method whereas the samples from the undamaged
region were tested to evaluate the static flexural strength that represented the virgin samples. Flexural tests were
carried out in displacement control mode using 20 kips capacity MTS testing machine. Three samples each from
undamaged region and repaired section were tested for each repair method. In all the cases, initiation of damage was
from the edges of the patch region. Typical load–deflections plots are given in. Figure 18a-f. Figure 18a and b represent
the load–deflection plots of the undamaged and repaired samples (method #1), whereas Figure 18c ad d represent the
load–deflection plots of the undamaged and repaired samples (method #2) and Figure 18e and f represent the load–
deflection plots of the undamaged and repaired samples (method #3). Almost all the repaired samples exhibit multiple
load drops with a major load-drop indicating the catastrophic initiation of the damage at the edge of the patch. Flexural
strength of the samples is given in Table 8. From the table it can be seen that the average flexural strength was
recovered by about 51–58%. The results indicate that the repair methods outlined in the current work, which do not
adopt any expensive tooling or materials for repairing seem to recover both static and dynamic properties to a
resonable level.
Fig. 12. Load–deflection plots of samples from repaired and
undamaged sections under flexure.[12]

Table 4. Static flexural strength[12]

Flexural strength (MPa)


Repair method Sample # Strength retention (%)

Undamaged section Repaired section


1 1 393 166 53
2 411 262
3 – 218
Average 402 215

2 1 552 320 51
2 475 117
3 511 349
Average 512.7 262

3 1 586 274 58.5


2 464 263
3 – 385
Average 525 307
[1] Strong, Dr. A. Brent.( 1989) “Fundamentals of Composites Manufacturing; Materials; Methods, and

Applications,” Society of Manufacturing Engineers, Dearborn, MI.

[2] Sohrab Kazemahvazi. (2010). Impact Loading of Composite and Sandwich Structures. Doctoral Thesis Stockholm,
Sweden

[3] B. 2. JANG*, L. C. CHEN, L. R. HWANG, J. E. HAWKES, and R. H. ZEE. The Response of Fibrous Composites to Impact
Loading. Materials Engineering Program Auburn University, Alabama 36849

[4] Kreculj, D.; Rasuo, B. Impact Problem on Aircraft Constructions from Composite Materials. (2009). // Technics,
Mechanical Engineering, Belgrade, pp. 1-8, (in Serbian).

[5] Ramazan Karakuzua, Emre Erbila and Mehmet Aktasb. (2010). Damage prediction in glass/epoxy laminates subjected
to impact loading. aDepartment of Mechanical Engineering, Dokuz Eylul University, 35100, Izmir, Turkey. bDepartment of
Mechanical Engineering, Usak University, 64300, Usak, Turkey.

[6] Baker, A.A., Introduction and overview, in Advances in the bonded composite repair of metallic aircraft structure,
A.A. Baker, L.F. Rose, and R. Jones, Editors. 2003, Elsevier: Oxford, UK. p. 1-18.

[7] Geoffrey Roy Goodmiller.(Dec 2013). Investigation of Composite Patch Performance Under Low-Velocity Impact
Loading. University of Tennessee – Knoxville.

[8] Denney, J.J. and S. Mall(1997) Characterization of disbond effects on fatigue crack growth behavior in aluminum
plate with bonded composite patch. Engineering Fracture Mechanics. 57(5): p. 507-525.

[9] Al-Zubaidy, H.A., X.-L. Zhao, and R. Al-Mahaidi.(2012). Dynamic bond strength between CFRP sheet and steel.
Composite Structures. 94(11): p. 3258-3270.

[10] L.B. Greczuk. (Sept 1973)Foreign Object impact damage to composites. A symposium sponsored by ASTM
Committee D-30 on High Modulus Fibers and their Composites AMERICAN SOCIETY FOR TESTING AND MATERIALS
Philadelphia, Pa.,

[11]J.W.C. Pang, I.P. Bond.(2005). Bleeding composites’—damage detection and self-repair using a biomimetic approach.
7th International Conference on the Deformation and Fracture of Composites (DFC-7)

[12] M.V. Hosura, , , U.K. Vaidyab, D. Myersa, S. Jeelania. (2003). Studies on the repair of ballistic impact damaged S2-
glass/vinyl ester laminates. Selected Papers from the Symposium on Design and Manufacturing of Composites

Você também pode gostar