Sei sulla pagina 1di 12

Thin-Walled Structures 119 (2017) 235–246

Contents lists available at ScienceDirect

Thin-Walled Structures
journal homepage: www.elsevier.com/locate/tws

Full length article

Design, analysis, fabrication, and testing of composite grid-stiffened panels MARK


for aircraft structures

Sonell Shroff, Ertan Acar, Christos Kassapoglou
Aerospace Structures & Computational Mechanics, Department of Aerospace Engineering, Delft University of Technology, The Netherlands

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

Keywords: Grid stiffened panels promise increased damage tolerance with reduced weight. Recent progress in automated
Grid structures manufacturing of composites has made it possible to produce such panels at low cost and thus has made them a
Composites competitive alternative to traditional skin-stiffened or sandwich panels. Fuselage skin panels typical of a 150
Manufacturing passenger aircraft were designed using an iterative process combining finite element models and local special
Failure
purpose analysis methods. 40 cm by 30 cm panels, representing the final design, were fabricated using vacuum
Impact damage
assisted resin transfer moulding. Pristine and impact damaged panels were tested in compression to failure. A
progressive failure model was used to predict the extent and type of damage during impact. It was then com-
bined with the global finite element model to obtain residual strength predictions. Analytical predictions were
very close to test results. The fabrication method showed good quality and consistency and, when automated,
can be used for production.

1. Introduction fibres are used across grid intersections with no local height increase,
the nodes end up with a higher fibre volume fraction than the rest of the
Grid stiffened panels can be used as an alternative to conventional grid due to doubling locally the amount of material. In case the panel is
stiffened skin design for fuselage of civil aircraft because of significant produced with pre-impregnated fibre tapes, the nodes also have excess
weight savings as a result of efficient load distribution between the resin that must be compacted in the mould adequately or removed after
grids that stiffen the skin and the skin itself. In early applications [1], curing [7]. When produced with the resin infusion technique, in-
the geodesic fuselage, the predecessor of the grid-stiffened skin, was adequate vacuum leads to voids and slow manufacturing process [8].
built with a fabric shell to behave as an aerodynamic cover while the With the development of automated fabrication methods of composites,
load was carried by the Duralumin geodesic frame. More recently, with grid stiffened panels are back in the forefront as a viable alternative to
the use of composites, the grid and skin can be designed such that there conventional designs [9,10]. This means that issues of concern such as
is an efficient load distribution between the two, where the grid bears the ones mentioned above must be well understood and quantified. To
most of the loads, and the skin is an aerodynamic cover that is able to that end, a program was undertaken to evaluate the structural perfor-
resist localised loads such as impact [2,3]. Grid stiffened structures are mance of composite grid-stiffened structures with emphasis on their
inherently highly damage tolerant due to multiple load paths present. damage tolerance.
Damage is typically limited to a single cell in the grid and grid-skin
debonding remains confined to only a small section of the structure [4]. 2. Baseline definition
Smaller bay areas in the grids are known to increase the impact re-
sistance according to experimental studies when compared to conven- A 150-passenger aircraft is considered as the baseline example. The
tional structures [5]. A higher in-plane stiffness and 28% lower de- conventional stringer-frame architecture is replaced by a grid-stiffened
flections when compared to an equivalent panel made of sandwich design. The fuselage components, such as the floor beams, frames and
structures [6] suggests it can be a good alternative design concept in attachments are optimized to relieve the loads in the fuselage skin.
certain applications. First, panels with low, intermediate and high loads are isolated and
Composite grid-stiffened panels were hard to implement in the past designed separately using a grid-stiffened configuration. Each panel
because of high manufacturing cost and concerns about quality control. typically has several critical load cases, maximum compression, max-
They have an inconsistent fibre volume fraction because, if continuous imum shear, or combined biaxial loading and shear, depending on


Corresponding author.
E-mail address: c.kassapoglou@tudelft.nl (C. Kassapoglou).

http://dx.doi.org/10.1016/j.tws.2017.06.006
Received 23 September 2016; Received in revised form 19 May 2017; Accepted 7 June 2017
0263-8231/ © 2017 Elsevier Ltd. All rights reserved.
S. Shroff et al. Thin-Walled Structures 119 (2017) 235–246

aircraft manoeuvres and mission scenarios. For each of these load cases, Moulding (VARTM) is a suitable alternative for manufacturing of grid-
different configurations are used and the geometry is optimized for stiffened panels. In this process, the resin is drawn into a preform solely
minimum weight checking for (a) strength, (b) buckling of the panel as under vacuum without the use of high amount of heat or pressure. This
a whole, (c) buckling of skin between ribs, (d) column buckling of ribs makes the process inexpensive to produce large, complex and high
and (e) crippling of ribs (local buckling followed by failure in the post- quality parts in one step. VARTM can be successfully used to produce
buckling regime). Note that all checks involving strength are based on grid-stiffened panels of high quality with minimum void content, if the
first ply failure using the Tsai-Hill failure criterion where the strength infusion process can be controlled and the parameters influencing the
values are reduced to account for material scatter, environmental ef- infusion process are understood.
fects, and damage. These are simplified failure predictions as opposed Coupled with VARTM, automated dry fibre placement followed by
to more detailed criteria used later in Section 5. During this process, out-of-autoclave curing for production processes is expected to be the
isogrid and anisogrid pattens are examined. It is interesting to note that most cost effective way in the near future. Fibre placement heads with
the grid-stiffened designs came out lighter than traditional skin stif- advanced capabilities and variable tow width handling are being de-
fened panels for load cases with significant loading in more than one veloped and improved constantly [10]. To support the technique, new
directions. Axial and transverse loads are more effectively transferred materials are being developed by major composite material suppliers.
with grid members in two directions and, for high shear loads, inclined Considering these possibilities for future automation and mass pro-
ribs allowed removing skin plies and led to a weight lower than that for duction of complex grid geometries and shapes of parts, dry fibre pla-
the traditional skin-stiffened panel. The main challenge with some of cement followed by vacuum infusion was chosen as the manufacturing
the grid-stiffened designs was to have acceptable manufacturing cost. method in this study. The panels were produced as a proof of concept,
The resulting designs for the low, medium, and high load cases are and therefore, did not make use of a robotic fibre placement head. The
manually blended with each other so a single pattern can be applied to fibre used was continuous and placed by hand. The fibre was also held
all. This involves adjustments for manufacturability such as matching in tension to maintain its form and shape due to lack of tack or “binder”
the angle of grids which are not aligned with the fuselage axis and between the adjacent layers.
fixing the grid height to a single value across panels irrespective of load An open-mould infusion process was selected with a female mould,
intensity. As a result, a small weight penalty is incurred compared to to place the fibres into, to prepare a preform, followed by infusion of a
the individually optimized panels but the ease of fabrication is greatly room-temperature resin so that the flow was easy and uniform without
improved. Finally, a highly loaded panel section of the fuselage is ex- the application of heat. This allowed close monitoring of the process to
tracted for detailed design and the loads acting on this panel are in- facilitate improvement for consecutive panels. An out-of autoclave
vestigated. This demonstration panel is shown in Fig. 1. curing process was used such that the panel was cured at room tem-
The extracted local panel is loaded under combined compression perature followed by a post cure cycle in an oven. Literature has shown
and shear of 430 N/mm and 75 N/mm respectively. The compression that the quality of the parts produced by either curing process, is similar
load is found to be roughly six times higher than the shear load, and is [11] and in fact, in some applications, the out-of autoclave cycle has
therefore considered to be the more significant one to evaluate the been reported to have lower void content [12]. In addition, other
design and fabrication process selected here. benefits over autoclave curing such as no tearing or leaking of vacuum
bag under autoclave pressure and easier compaction of thin skin sec-
tions make this an attractive alternative.
3. Fabrication process
A 3D printed plastic lattice or grid was used to produce a silicone
rubber mould with the Silastic S green elastomer (Fig. 2a), due to its
Automated fibre placement techniques are developed for a variety
superior flexibility and strength combination. The thermal expansion
of multidirectional, steered and tailored composite panels. The most
coefficient measured for the silicone rubber tooling is approximately
common processes use pre-impregnated fibre tows that can be placed or
342 × 10−6 mm/mm-°C resulting in a maximum expansion of
stacked to build a grid of the required height, however, there is an
0.376 mm in the thickness direction and 0.94 mm in the width direction
excessive material build-up at nodal intersections that has to be avoided
for large mould pockets when post- cured at 80 °C. This mould was
by highly compacting material at the nodes and/or spreading the tows
cured at room temperature for 12 h before fibre placement. The same
around the intersection area to maintain a consistent grid height. The
rubber mould was used to produce all 6 panels and remained intact for
fibre volume fractions at each intersection tend to vary and the panel,
further repeated use showing no discernible damage.
therefore, lacks consistency. Another difficulty with this process is the
The carbon fibre tows used in this process were continuous and
requirement of an autoclave for the curing processes, which limits the
since they were hand laid, a suitable pattern was used so as to prevent
size and shape of the part that can be manufactured, and increases the
breaking the fibre and to save time. This “weaving” or placement pat-
cost of the production process. Vacuum Assisted Resin Transfer
tern was facilitated with a table with pins at desired locations. A CAD
model of this table with the mould assembly is shown in Fig. 2b. The
dotted lines with arrows show the pattern used in weaving the fibre
through the mould grooves. The carbon fibre was wrapped around
hand-held spools which were used to place continuous dry fibres in the
grooves. The fibres were wound around the pins on the placement table
to keep them straight and flat between adjacent layers. Placement of the
fibres took place in a clean room with controlled temperature and
moisture. In this process 6 mm wide dry Torayca T700S carbon fibre
tows were used.
After the placement was complete, the dry fibre skin layup was
placed on the mould to cover the grid. One meter wide unidirectional
Torayca T300 carbon tape plies were cut to size for the skin layup. The
assembly was prepared with bleeder and vacuum tubes and was placed
on a plate and vacuum bagged (Fig. 2c). The mould was kept under
vacuum for 6 h to ensure removal of most of the trapped air bubbles
Fig. 1. Generic grid stiffened fuselage, with the optimized half and pull-out of the local between the fibre layers.
panel selected for demonstration. A room temperature resin, Epikote 04908 was used for infusion of

236
S. Shroff et al. Thin-Walled Structures 119 (2017) 235–246

Fig. 2. Manufacturing Process. (a) Mould is made from silicone


rubber with an assembly of 3D printed grids; (b) Winding table is
prepared and dry carbon fibre tows are placed in the grooves; (c)
VARTM is carried out with a room temperature resin; (d) Cured
panel.

the panel. The complete infusion process lasted 2 h at room tempera- hence the grids were narrower than required.
ture. The panel was left to cure for 12 h. Subsequently, post curing was It is interesting to point out that, unlike the grid height and width
carried out at 80 deg C for 4 h. values which were more or less equally distributed above and below the
nominal or reference value, the skin thickness was mostly higher than
4. Destructive and non-destructive evaluation of part quality the reference value (see Fig. 4). A possible reason for this one-sided
thickness variation is the lack of sufficient pressure on the silicone
The quality of the fabricated parts was investigated by measuring mould during the infusion and curing: The silicone rubber expands by a
the variation in grid height, width and skin thickness. In addition, fibre small amount because the temperature is higher than the room tem-
volume fraction of the panels was measured at several locations and perature. Over larger skin pockets, the pressure on the skin is greater
micrographs of section cuts provided additional information at critical than over smaller skin pockets. This difference in pressure allows larger
regions such as nodal intersection. resin rich areas or air pockets to remain in the skin. This pressure can
The variation in grid height, width and skin thickness over the be, however, made more uniform by applying a caul plate minimizing
length of the fabricated panels is shown in Figs. 3 and 4. Summary the problem and resulting in a skin thickness closer to the reference
results are also given in Table 1. Panels 1 and 2 consistently show high value.
deviations from the reference height (see Fig. 3) indicated by the hor- The part quality improved as the process parameters such as
izontal line at 21.37 mm for all three grid types. The standard deviation number and location of inlets, vacuum outlets, and, flash to remove
of the grid heights for the six panels (see Fig. 5 and Table 1) shows a excess air in the preform, were improved. The standard deviation in the
steep improvement as compared to Panel 1 in the dimensional tolerance grid width improves towards the last panel, Panel 6. One of the most
of the part. The standard deviation of the grid height for the last panel important effects on the grid cross-sections in terms of dimensions is
(panel 6) is between six and seven times smaller that of the first panel that of the silicone rubber mould. The stiffened panel can either be post-
(panel 1). Panels 4, 5 and 6, fabricated in that order, show steadily cured at a higher temperature of 40° with the silicone rubber mould and
improving dimensional tolerance and consistency along the length and a caul plate to minimize rubber expansion, or post-cured without the
width of the panels as manifested by the decrease in scatter around the rubber mould. The fabrication results show that the latter option is
nominal value. better for the panel dimension, because, panels 5 and 6 were both post-
The width variation for the grids is shown in Fig. 3b, wherein, the cured without the silicone rubber mould around them and show a
width of Panel 2 is seen to have a maximum variation along the length maximum coefficient of variation of 1.4% and 2.5% for the grid height
of the grid, as well as far from reference width. Panel 2 has been found and width respectively and this deviation is consistent for all three grid-
to have the lowest quality due to excessive air bubbles accumulated types as shown in Fig. 5.
inside the vacuum setup. This caused zones of lower resin content, and The fibre volume fraction was measured using the burn-out process

237
S. Shroff et al. Thin-Walled Structures 119 (2017) 235–246

Fig. 3. Variation in grid height and width compared to nominal (reference) value. a. Variation in grid height. b. Variation in grid width.

of cured reinforced resin using ASTM standard D2584-02. An average wider allowance was provided in the tooling at the nodes and in a
fibre volume fraction in the grid away from nodal intersections was transition region 5 mm into and out of each node to spread the tows of
found to be approximately 35% with an average of 0.75% voids. material and minimize the tendency for thickness build-up at the node
Maximum void content was measured as 0.8%. At the intersections or itself. The fibre volume measurements in this region did not include the
grid nodes, the average fibre volume fraction was found to be 47% with extra 1 mm worth of material and this reduces the fibre volume. The
0.5% voids. Usually, the fibre volume at grid intersections is higher highest void content there was measured as 1.2%. This fibre volume
than this value. However, in the present design, up to an extra 1 mm fraction must be improved for better strength properties of the grid and

238
S. Shroff et al. Thin-Walled Structures 119 (2017) 235–246

pressure and debulked to remove the trapped air between the different
layers. In that case, more fibre layers can be placed inside the grid
groove. A higher tension force in the fibre tows during placement is
expected to reduce the waviness in the fibre tows as is seen in the
current process (Fig. 7a). The fibre tension was not measured and
hence, there was no way of checking for consistency in the current
process.

5. Analysis model

The finite element model was created in ABAQUS and care was
taken to model the grid stiffened panel as manufactured along with its
material properties. The fibre volume fractions from the destructive
evaluation of the panel were used to calculate the moduli, strength and
Fig. 4. Variation in skin thickness compared to reference thickness.
other material properties of the panel. For strength, this was done using
micromechanics and assuming plane stress conditions, no voids, and
intersections. The low fibre volume fraction is partly a result of using a
perfect bonding between fibres and matrix. The resulting strength va-
hand-placement method with slit tapes that were slightly wider than lues for 35% fibre volume were as shown in Table 2 below.
the grooves in the mould because the standard tow available from the
The grid intersections needed accurate representation, since there is
supplier was slightly different than the width of the grid members de- double the number of carbon fibre layers stacked on top of each other at
termined by optimization. Folding of the tape, insufficient compaction
this location. The grid layup was modelled in two ways to investigate
and debulking are some of the possible causes of lower fibre volume the effect of the modeling technique. The first included calculating
fraction than the target value of 50%. Better control of the process can
equivalent properties for a sub-laminate of two plies of adjacent or-
lead to a higher fibre volume fraction. The fibre, matrix and void vo- ientation angles. For example, for an intersection of 0° and 90°, prop-
lume fractions at various locations in panel 6 are reported in Fig. 6.
erties for an equivalent ply of layup 0/90 were calculated and used as
Eighty (80) micrographs were made at various locations in the input. In the second approach, the grid height was divided into double
panel, especially the grid intersections to investigate voids and wavi-
the number of elements through the thickness, with a single through-
ness of the fibres. The micrographs of three grid intersections are shown the-thickness element for each ply orientation. In the second case, it
in Fig. 7a. The waviness of the stacked fibres is visible in the left most
was also necessary to gradually reduce the properties of the regions in
image, whereas the other two images are for an angled intersection and the grid adjacent to the grid intersections. Comparison between the two
show the fibre pattern to be more uniform. The micrograph shown in approaches showed that the first method was simpler to model and
Fig. 7b belongs to a longitudinal grid section far away from an inter- more reliable. It did not show any stress discontinuities for a converged
section. Some small voids, and large resin rich areas can be seen in this mesh as opposed to the second method. Therefore, the first method was
image. implemented in the final model.
The non-destructive evaluation was done on the panel using a phase The grid stiffened panel was modelled with two independent parts –
array ultrasonic system [13] and a C-scan setup to investigate part the skin and the grid. The skin was a 14-ply laminate with a stacking
quality on pristine panel and to determine the level of damage after sequence of (45/−45/04/90)S and the unidirectional grids are shown
impact testing. The phase array ultrasonic system clearly showed the in Fig. 8. The skin and grid are connected with a tie constraint. The
poorly infused areas of the earlier fabricated panels which were later mesh was created using reduced integration solid elements with one
investigated with section cuts to verify the ultrasonic scans. This element through the thickness of each ply. A distortion control was used
method of investigation helped determine the problem areas to improve to prevent the elements at the free edges from exhibiting large distor-
the fabrication process further. tions causing a premature failure of the analysis. For the compression
The manufacturing process can be improved in the future with the test, the edges of the panel up to the first grid transverse to the loading
goal of increasing the fibre volume fraction by using the width of the
were constrained to move only in the loading direction to represent the
fibre tows as the grid width for the design and an automated tape fibre potting added to the edges of the panel in the experiments as shown in
placement process, such that the tows are compacted using sufficient
the Fig. 8. The edges parallel to the load were left free whereas the

Table 1
Nominal value, average, and standard deviation (SD) of grid height and width and skin thickness of panels 3 through 6.

Panel 3 Panel 4 Panel 5 Panel 6

Attribute Nominal Average SD Average SD Average SD Average SD

* SD = Standard Deviation

Grid height
Longitudinal 21.37 21.32 0.73 21.3 0.14 21.53 0.13 21.23 0.11
Circumferential 21.37 22.17 0.63 21.2 0.24 21.81 0.27 21.17 0.09
Helical 21.37 21.54 0.13 21.4 0.24 21.61 0.15 21.47 0.19

Grid width
Longitudinal 3.8 3.76 0.09 3.77 0.1 3.78 0.05 3.67 0.09
Circumferential 3 2.98 0.16 3.05 0.15 3.06 0.06 2.86 0.1
Helical 3.8 3.73 0.05 3.9 0.05 3.81 0.05 3.79 0.1

Skin thickness 1.4 1.42 0.06 1.48 0.06 – – – –

**All dimensions are in mm.

239
S. Shroff et al. Thin-Walled Structures 119 (2017) 235–246

Fig. 5. Standard deviation in grid height and width.

Fig. 6. Measured volume fractions in panel 6. a. At grid locations. b. At grid intersections.

bottom edge was fully constrained. A displacement load was applied on evolution of this damage is often reported to be accompanied by a
the top edge to simulate a pure compression loading. In case of the sudden loss in stiffness in the composite laminate. The damage accu-
impact simulation, the potted and constrained edges were modelled as mulates and, eventually, final failure of the composite laminate occurs.
simply supported, reflecting the assumptions made in the tests. Damage progression depends largely upon laminate parameters, and it
Fibre reinforced plastics are typically expected to exhibit a linearly can be captured in detail if a suitable material model is developed
elastic material behaviour up to the point of damage initiation. The coupled with a robust failure criterion and stiffness degradation model.

Fig. 7. Micrographs at various grid locations (with


voids marked with circles and resin rich areas
marked with rectangles). a. Grid intersections. b.
Grid away from intersection.

240
S. Shroff et al. Thin-Walled Structures 119 (2017) 235–246

Table 2
Calculated strength properties at 35% fibre volume.

Property Value

Tensile Strength in 1-direction, 1715 MPa


Compression Strength in 1-direction 72.25 MPa
Tensile Strength in 2-direction 514.5 MPa
Compression Strength in 2-direction 2261.54 MPa
Shear strength in 12-direction 120 MPa
Shear strength in 13-direction 120 MPa

Although composite materials have suitable specific strength and stiff-


ness as compared to metals used in aerospace applications, they lose a
lot of their weight advantage when damaged. A tool drop during
maintenance can cause a barely visible impact damage (BVID) on thin
laminates used as the fuselage skin that can easily escape a visual in-
spection that should call for repair. Hence, it is important to study the
behaviour of such materials under damage events that a visual in-
spection can easily miss.
An analysis model [14] that captures onset of damage as well as its
evolution until final failure using a combination of failure criteria,
Hashin [15], for fibre failure, and Puck [16] for matrix failure, along
with an energy release rate-based approach to predict delamination,
were applied to the finite element model to predict progressive failure
under impact and subsequent compression loads. The failure model is
presented in Fig. 9. The failure criteria are used in combination with a
degradation model such that when failure is predicted, the stiffness of
the element region is reduced by a certain amount. This degradation
model is a modified Lee [17], Hwang and Sun [18] model with varied
degradation for different failure modes and is presented in Fig. 10. The
specific values used in Fig. 10 are the result of a detailed sensitivity
study by Shroff and Kassapoglou in [14].
The general procedure followed in the progressive failure analysis is
as follows:

• At initial state, the stiffness matrix of the model is computed


• A displacement consistent with the selected impact load is applied
on the finite element model
• The strain state of the model is obtained from ABAQUS
• The stiffness matrix and stress state of the model are updated using
the strain update
• The fibre and matrix failure criteria are applied to check for failure.
If failure occurs, the applicable degradation factor from Fig. 10 is
used to update the stiffness matrix Fig. 9. Failure criteria used in the finite element based analysis model.
• The 3D strain energy is computed to check for delamination damage
using Eq. (6). If failure occurs, the applicable degradation factor is
applied to update the stiffness matrix again

Fig. 8. Dimensions and boundary conditions of the compression tested panel.

241
S. Shroff et al. Thin-Walled Structures 119 (2017) 235–246

diameter of 16 mm was dropped with a total mass of 3.23 kg from a


height of 0.47 m and 0.63 m for a low velocity impact of 15 J and 20 J
energy levels at locations 1 and 2 respectively shown in Fig. 11b. The
final dimensions of the impacted panel taking into account the
boundary conditions are also shown.
The six panels produced using vacuum infusion were cleaned and
cut to the final dimensions at all four edges using a diamond wheel
cutter. Each panel was then potted with potting agent Rencast TM CW-
2418/Renr HY5160. The horizontal faces of the potting were milled flat
to ensure uniform compression in the test bench. The geometry of the
finished panel is shown in Fig. 11c. The flat skin side of the panel was
then painted to create a speckle pattern for a DIC while strain gauges
were placed on specific locations on the grid side of the panel on both
the grid and the skin surfaces. The location of the strain gauges is shown
Fig. 10. Degradation model. (Note: Cij are the ply-level stiffnesses in (no summation on in Fig. 11d. For impact, the free edges of the panel were constrained to
indices i, j = 1, 2, 3). imitate simply supported boundary conditions. Therefore, Aluminium
jigs with a “C” cross-section were machined to fit in the potted length of
• If no failure occurs, an increased load is applied and the process is the panel and slide in to support the free edges of the grid stiffened
repeated panel. The jigs were then held in position using clamps.
• Final failure is predicted when the total strain energy of the model During the compression test, strain measurements were taken at
drops signifying no load transfer through the structure 0 kN load level to use as reference. The loading was applied manually in
steps of 10 kN at lower load levels, 5 kN at medium load levels and
Upon simulating an impact on a grid stiffened composite panel, it 2.5 kN at higher load levels, close to the expected failure loads as
was observed that when the impact energy was such that the damage predicted by the FE analysis. At higher load levels, a series of sounds
was contained within a skin pocket, the area delineated by adjacent were heard indicating matrix crushing, limited fibre breakage and de-
grid members, the final compression failure load was within 10% of the lamination in the layers of the skin. The load levels were recorded,
undamaged failure load. This implies that the load in the damaged skin however, the exact location of the sound source was not detected.
can be effectively redistributed to the surrounding grid members with Finally, the panels failed due to grid-skin separation with a very loud
little effect on the overall load carrying capacity. This is a verification of sound. Strain versus load curves with strain results from a sample lo-
the damage tolerance of grid-stiffened structures. The strength reduc- cation on the panel on the grid side (measured with a strain gauge) and
tion was higher if adjacent skin pockets were damaged by impact. Then, on the skin side (measured with the DIC) from the test are shown in
the load transfer from the skin overloaded the grid members adjacent to Fig. 12a. The force versus displacement diagram measured with the
the impacted skin pockets. In either case, the section of the skin con- LVDTs is given in Fig. 12b for panels 1, 2 and 6. This diagram also
taining the impact site was assumed to carry no load. The load was indicates the load levels at which the sounds were heard during the test.
assumed to be redistributed over the grids and the pristine skin section. All these observations have been compared to the finite element pre-
Using the predicted total load carried by the grids and the buckling load dictions of failure instances and show less than 1% difference in the
of the skin pockets from the finite element model, a redistribution ratio load levels at which these failures are predicted to initiate in the finite
could be determined. The progressive damage model was used to pre- element model.
dict the damage initiation loads which, in combination with the re- The impact-tested panels showed very little surface damage under
distribution ratio, could predict the failure load of the single or multiple 15 J impact. More damage with visible fibre slits was evident at 25 J
impacted panel under compression. The final failure load of a panel impact and as shown in Fig. 13. Post-test NDI evaluations were also
with two impacted skin pockets was found to be 7.5% lower than one carried out. The impacted panel was C-scanned using a C-scan built in-
with a single impacted skin pocket. house at the DASML in Delft. The images of the C-scan can be seen in
Sensitivity studies based on grid pocket sizes using an orthogrid Fig. 14a and b. The damage diameter from the C-scan was then com-
stiffened panel under impact [14], showed that using realistic boundary pared to the predicted results from the FE analysis.
conditions, the damage size can be reduced by making the skin pocket
sufficiently small before the damage progresses through the grid, at the 7. Comparisons of analytical predictions to test results
same impact energy. This should be further explored for isogrid and
anisogrid stiffened panels for a robust prediction. Undamaged specimens: The approach described in the analysis sec-
tion was used to predict failure for undamaged specimens. The pre-
6. Tests and post-test evaluation dicted failure sequence, shown in Fig. 15, was as follows (a) Matrix
cracks at or near 95.44 kN (b) Delaminations at or near 103.5 kN (c)
The compression and impact tests were carried out at the Delft Broken fibres at 119.9 kN.
Aerospace Structures and Materials Laboratory (DASML), at the The non-destructive c-scan of panel 6 did not show significant voids
Technical University of Delft, The Netherlands. The compression tests in the skin pockets. Still, in order to better determine material prop-
were performed on a servo-hydraulic compression test bench with a erties and manufacturing consistency, the fibre volume fraction of the
capacity of 600 kN under deflection control. The setup was equipped panel was measured in various locations of the panel. The finite ele-
with Linear Variable Displacement Transducers (LVDTs) to measure ment model was subsequently modified with the updated material
cross head displacements. The strain gauges on the grid side of the properties in the grids and skin for a better comparison of the experi-
panel were connected to a Keithly Data Acquisition System, whereas mental results. A comparison of the out of plane displacement of the
VIC-3D (Correlated, 2013), a Digital Image Correlation (DIC) system section A-A marked on the panel in Fig. 11c from the DIC results with
was setup and calibrated to measure the displacement and strain on the the finite element predictions agreed very well for load levels higher
skin surface during the test at various load levels. The experimental than 50 kN and agreed well with lower load levels. A comparison of the
setup is shown in Fig. 11a. out of plane displacements for section A-A for load levels of 60 kN and
The impact tests were performed on an impact drop tower with a 100 kN show that the behaviour of the panels under the compression
capacity of 1600 J energy level. A spherical steel impactor with a load is very well predicted for both cases with a maximum error of less

242
S. Shroff et al. Thin-Walled Structures 119 (2017) 235–246

Fig. 11. Test specimen configuration. a: Compression test setup on the skin side. b: Final dimensions, boundary conditions and impact locations of test panel. c: Final potted test panel
dimensions. d: Strain gauge locations on grid side.

than 8% for the 100 kN load level. The predicted out of plane dis- to measure the impact damage size for the tested panels. The damage
placement versus panel length curve is compared to the test result in sizes for the various impacts are shown in Fig. 17. The finite element
Fig. 16. predictions are also shown in the figure and are marked by circles. The
To understand the effect of impact, the grid stiffened panel was experimental data is from different grid stiffened panels manufactured
impacted on the skin side of the panel, at the centre of each skin pocket. in separate stages and the skin pocket impacts did not come from the
The energy levels were selected, such that the damage does not go past same global sample. These factors contribute to the scatter at energy
the surrounding grids. Finite element models were created to predict levels of 15 J and 20 J. The FE predictions for the impact damage sizes
impact damage areas with varying levels of impact energy at skin are overestimated for the lower energy levels up to 15 J. The difference
pockets. A phase array ultrasonic system and the c-scan setup were used of the damage areas for higher energy levels is approximately 7%,

Fig. 12. Typical test results from the compression test. a. Strain versus load. b. Load versus displacement.

243
S. Shroff et al. Thin-Walled Structures 119 (2017) 235–246

ultimate failure by grid-skin separation. This failure sequence predicted


by the FE model was reproduced in the tests. The awareness of the
failure sequence can now be incorporated in to design practices to delay
or prevent early failure such that the ultimate failure is also delayed.
For example, the matrix cracks appearing in the skin can be delayed
or prevented by optimizing the skin laminate layup so that the propa-
gation of the matrix cracks leading to delamination are delayed. Even
with an optimized skin laminate, the delaminations at the grid-skin
interface, still remain a big threat to the strength of the panel. The
design must, therefore, consider the grid cross-section, size and shape as
design parameters to prevent the excessive shear stresses and stiffness
differences at the interfaces. The grid orientation and density, as con-
cluded by various previous researchers [19–22], still remain important
Fig. 13. Photograph of 25 J impact damage on panel 4 (skin side).
design parameters.
The grid stiffened design is very damage tolerant as also proven by
whereas at lower energy levels, the difference was found to be between the analysis and tests of skin pocket-impacted panels. The impact da-
35% and 80%. mage does not affect the failure strength of the panel as long as a single
The phased array system used for the measurement of damage sizes impacted pocket is isolated and is located at least two bays away from
was operated manually and therefore, introduces an error, however, another impacted pocket. The proximity of multiple damage sites and
this error cannot be estimated unless better ultrasonic scans are made their effect on the static and fatigue performance of a panel will be an
with calibrated automated systems. The C-scan also showed scatter in important research focus in the near future.
the images, due to the interference of water splashing from the grids Grid stiffened structures are being currently manufactured using
back on to the skin during the scan. Hence, more precise measurements filament winding processes, in combination with autoclave manu-
have to be made in the future with equipment more suited to the stif- facturing [2]. These structures can potentially be produced by an out-
fened panel with high grids. Despite this error, the largest difference of of-autoclave automated tape placement method. Alternatively, dry fibre
80% in the damage area prediction with the FE model, is too large to be placement can also be used to prepare preforms of these structures to be
attributed to just a measurement error. Another reason for this dis- vacuum infused later and cured in an oven [23]. The difficulty in the
crepancy is the fact that infusion was not perfect at the impact loca- first method of tape placement, is the production of the grid intersec-
tions. As a result the skin had lower stiffness there and, under impact, tions. A choice has to be made to allow height build-up, or differentially
would absorb more energy by deflecting than by damage predicted by compress the material so the pressure at intersections is higher, or steer
the model. around and intersection, or even alternately cut fibres to keep the
height constant. Out of autoclave processed prepreg tapes are now
available that can reduce the cost of curing. The difficulty with dry
8. Discussion and conclusions preform fabrication process is the increased cost of dry fibre tows that
must have a layer of binder material to hold the fibres in place and in
The use of progressive damage prediction in a finite element based tension during the placement process.
analysis has led to the conclusion that the failure in the composite panel The analytical predictions discussed in the current paper have
initiates due to matrix cracking in the skin in the vicinity of the grid shown a good agreement with literature and with results of tests con-
interfaces. When these matrix cracks propagate and reach ply inter- ducted by the authors. The final failure load is very accurate, however,
faces, they cause delaminations in the skin. The delaminations then the actual damage sizes differing by as much as 80% but typically
cause loss of stiffness in the region, ultimately overloading the grid-skin within 30% of the test results can still be further improved by im-
interface. Here, the normal and shear stresses exacerbated by the mo- proving the finite element analysis. The current approach used the Puck
ment due to the eccentricity between the grid and the skin, lead to

Fig. 14. Ultrasonic inspection results showing impact damage


and manufacturing defects. a: C-Scan of Panel 2 with multiple
impact damage and manufacturing defects. b: C-Scan of Panel 6
with single impact damage.

244
S. Shroff et al. Thin-Walled Structures 119 (2017) 235–246

Fig. 15. Progression of damage in Panel 6.

Fig. 16. Compression test of undamaged panels compared with the analytical model.

only be done by introducing the manufacturing parameters as a design


input. The current production method also poses limitations in grid
height, density and orientation. Improved automated hardware will
also allow steered fibre paths for optimum performance.
Finally, difficulties with non-destructive inspections to identify the
locations and extent of the damage are an area that must be addressed if
grid-stiffened structures are to break into the mainstream. Once the grid
exceeds a certain minimum height, the C-scan equipment needs to be
improved beyond what is currently commercially available to detect
any damage at all through the thickness and at the grid-skin intersec-
tion.
Fig. 17. Comparison of damage area of impact test with finite element predictions.
References
failure criterion for the matrix based failures. It has been found to
[1] D. Paul, D. Pratt, History of flight vehicle structures 1903–1990, J. Aircr. 41 (5)
perform better than other failure criteria, however, the delamination
(2004) 969–977.
based on fracture energy can be further improved by taking into ac- [2] V.V. Vasiliev, V.A. Barynin, A. Razin, Anisogrid composite lattice structures – de-
count mode mixity to make the prediction even more accurate. velopment and aerospace applications, Compos. Struct. 94 (2012) 1117–1127.
Understanding of the behaviour of composite grid stiffened panels [3] H.J. Chen, S.W. Tsai, Analysis and optimum design of composite grid structures, J.
Compos. Mater. 30 (4) (1996) 503–534.
and low cost manufacturing with consistent quality have improved a lot [4] D.J. Baker, D.R. Ambur, J. Fudge, C. Kassapoglou, Optimal design and damage
in recent years. The automated production process needs improving tolerance verification of an isogrid structure for helicopter application, AIAA/
and further development in terms of accuracy and precision. This can ASME/ASCE/AHS Structures, Structural Dynamics, and Materials Conference, 44,

245
S. Shroff et al. Thin-Walled Structures 119 (2017) 235–246

2003. [14] S. Shroff, C. Kassapoglou, Progressive failure modelling of impacted composite


[5] V.D. Muthyala, Composite Sandwich Structure with Grid Stiffened Core (Thesis), panels under compression, Submitt. to: Compos. Part B (2014).
Louisiana State University, 2007. [15] Z. Hashin, Failure criteria for unidirectional fiber composites, ASME J. Appl. Mech.
[6] T.E. Meink, Composite grid vs. composite sandwich: a comparison based on payload 47 (1980) 329–334.
shroud requirements, IEEE Aerospace Conference, 1, pp. 215–220, 1998. [16] A. Puck, Festigkeitsanalyse von Faser-matrix-Laminaten: modelle fur die Praxis,
[7] S.M. Huybrechts, T.E. Meink, P.M. Wegner, J.M. Ganley, Manufacturing theory for Hanser (1996) (ISBN 9783446181946).
advanced grid stiffened structures, Compos. Part A: Appl. Sci. Manuf. 33 (2) (2002) [17] J.D. Lee, Three dimensional finite element analysis of damage accumulation in
155–161. composite laminate, Compos. Struct. 15 (1982) 335–350.
[8] S.K. Mazumdar, Composites Manufacturing: Materials, Product and Process, [18] W.C. Hwang, C.T. Sun, Failure analysis of laminated composite by using iterative
Engineering, CRC Press, 2012. three dimensional finite element method, Compos. Struct. 33 (1989) 41–47.
[9] F. Chinesta, A. Leygue, B. Bognet, Ch Ghnatios, F. Poulhaon, F. Bordeu, [19] E. Wodesenbet, S. Kidane, S. Pang, Optimization of buckling loads of grid stiffened
A. Barasinski, A. Poitou, S. Chatel, S. Maison-Le-Poec, First steps towards an ad- composite panels, Compos. Struct. (2003).
vanced simulation of composites manufacturing by automated tape placement, Int. [20] N. Jaunky, N. Knight, D. Ambur, Formulation of an improved smeared stiffener
J. Mater. Form. 7 (1) (2014) 81–92. theory for buckling analysis of grid-stiffened composite panels, Compos. Part B:
[10] A. Levy, J. Tierney, D. Heider, P. Lefebure, D. Lang, Simulation and optimization of Eng. 27 (5) (1996) 519–526.
the thermoplastic automated tape placement (ATP) process, Proc. SAMPE (2012). [21] N. Jaunky, N. Knight, D. Ambur, Optimal design of grid-stiffened composite panels
[11] A.T. Nettles, J.R. Jackson, Compression after impact strength of out-of-autoclave using global and local buckling analyses, in: Proceedings of the 37th AIAA/ASME/
processed laminates, J. Reinf. Plast. Compos. 32 (24) (2013) 1887–1894. ASCE/AHS/ASC Structures, Structural Dynamics and Materials Conference and
[12] J.K. Sutter, L. Kenner, L. Pelham, et al., Comparison of autoclave and out-of-auto- Exhibit, pp. 2315–2325, 1976.
clave composites, in: Proceedings of the 42nd International SAMPE symposium and [22] Z. Gürdal, G. Gendron, Optimal design of geodesically stiffened composite cylind-
technical conference (42nd Proceedings), 2010. rical shells, Compos. Eng. 3 (12) (1993) 1131–1147.
[13] L.W. Schemmer Jr., Fundamentals of ultrasonic phased arrays, Mod. Phys. Lett. 22 [23] G. Totaro, F. De Nicola, Recent advance on design and manufacturing of composite
(11) (2008). anisogrid structures for space launchers, Acta Astronaut. 81 (2) (2012) 570–577.

246

Potrebbero piacerti anche