Sei sulla pagina 1di 75

COMPOSITI AVANZATI

INTRODUZIONE AI MATERIALI FIBROSI


Parte I: ASPETTI GENERALI

Giacomo Frulla. Tel 0115646842–e- mail : giacomo.frulla@polito.it


Tecnologia delle fibre
Nella ricerca di materiali con basso valore di densità, l’attenzione si è rivolta verso quegli elementi della
tavola periodica che occupano i posti più bassi. Dopo idrogeno ed elio si trovano in successione Litio,
Berillio, Boro, Carbonio.
Con l’esclusione dei primi due estremamente costosi , quelli su cui si concentrò l’attenzione per impieghi in
campo aeronautico, furono il Boro e il Carbonio .
Caratteristica di entrambi : valori elevati del modulo elastico e di resistenza alla rottura a trazione, basso
valore della densità buone proprietà rispetto alla fatica e al creep, resistenza alla corrosione , disponibilità
costo e lavorabilità.

Negli Stati Uniti vennero concentrate le risorse sullo sviluppo delle fibre di boro con buoni risultati. In
Giappone ed in Inghilterra invece si concentrarono le risorse sulle fibre di carbonio prodotte mediante
pirolisi del poli-acrilo-nitrile (PAN).

FIBRE DI CARBONIO
Il materiale di partenza che deve essere usato per realizzare la fibra deve possedere le seguenti
proprietà:
-Temperatura di fusione maggiore della temperatura di decomposizione
- Contenuto di carbonio il più alto possibile
- Scarsa presenza di ossigeno nella molecola (per evitare la riduzione della resistenza a trazione).
Tra i polimeri disponibili in commercio alcuni vennero scartati pur avendo una percentuale di carbonio
elevata (Polistirolo 92%, Polietilene 86%, Polipropilene 86%, Poliestere 80%) relativamente al primo
punto delle caratteristiche precedenti .
Furono quindi utilizzati il Poliacrilonitrile (PAN) la cellulosa (rayon) , il Nylon (6,6), il Bitume.
Reazione
La reazione primaria di pirolisi avviene in un intervallo di temperatura
compreso tra 423 e 513 Kelvin (150-240 C) in presenza di ossigeno
(decomposizione primaria lenta e decomposizione secondaria veloce) ,
con la successione indicata nella figura precedente.
L’ultimo schema presenta la formazione di unità peridinica. Per successivi
passaggi si giunge alla polipiridina.

Si ottiene quindi un anello stabile , alla fine della reazione primaria,


composto da 5 parti in carbonio e il rimanente in azoto e idrogeno.
Il materiale così ottenuto ha notevole resistenza al calore, ma scarso
interesse strutturale.
Le caratteristiche meccaniche furono incrementate mediante
carbonizzazione in atmosfera inerte fino a temperature di 2773 K (2500
C), trasformando la molecola in struttura simile a grafite . (espulsione
dell’azoto e dell’idrogeno rimanenti) .

L’incremento ottimale delle


caratteristiche meccaniche però fu
ottenuto mediante applicazione alla
struttura precedente di una trazione
durante la fase iniziale di pirolisi
(200C)
Effetto della trazione sulle caratteristiche meccaniche della grafite.
Si evidenzia inoltre come le fibre di grafite abbiano un coefficiente di
dilatazione termica negativo . (bloccaggio della fibra durante la
lavorazione per imporre la voluta trazione)
FIBRE DI BORO

Vengono prodotte mediante decomposizione termica di un composto del


boro su un filamento di tungsteno .
Il metodo più diffuso consiste nel far passare un filamento di tungsteno
(0.013mm diametro) attraverso una camera di reazione contenente una
miscela di tricloruro di boro e di idrogeno. Il filamento viene riscaldato a
temperature comprese tra 900C e 1400C permettendo la decomposizione
del composto e la deposizione del boro sul filamento.
La fibra risulta quindi spessa ( 0.14-0.20 mm ) con una superficie
irregolare che viene rifinita (eliminazione dei vuoti e delle cricche )
mediante il controllo della velocità di produzione (150 m per ora) . Elevati
costi di produzione.
FIBRE DI ARAMIDE
Note come Kevlar, (brevetto Du Pont) . Viene prodotta da una soluzione di p-
fenileneterftlmide originando una struttura molecolare cristallina come in figura .
Le sue caratteristiche a trazione sono tra le più elevate rapportato alla densità.
Può essere utilizzato fino a 300C conservando il 75% delle caratteristiche
meccaniche .
Presnta però un valore molto basso di resistenza a compressione.
MATRICI IN RESINA

Mentre la fibra ha il compito di sopportare i carichi,. La matrice viene utilizzata


per distribuire i carichi tra le fibre, mantenerle allineate nella posizione voluta e
proteggere dagli agenti esterni garantendo una buona adesione tra gli strati.
Le resine più diffuse sono quelle termoindurenti capaci di sopportare temperature
elevate (max 175C) e le resine poliimidiche (max 315C).

Epossidiche: maggiormente utilizzate. Sono caratterizzate da un gruppo


epossidico (atomo di ossigeno che lega due atomi di carbonio adiacenti della
catena polimerica) che forma legami tridimensionali.
Sistema tipico basato su diglicidil-etere e bisfenolo A .
Come agenti polimerizzanti vengono usate le Ammine aromatiche.
Precauzioni ambientali.
Esistono anche resine polyestere e vinlyestere con differenti caratteristiche.
Effetto del micro-crack e
della presenza di acqua
sulla resina
La resina deve essere capace di
deformarsi con la fibra pertanto in base
agli allungamenti massimi delle fibre
considerate, deve essere scelta la resina
adatta.
Tipici risultati per compositi fibra-epoxy
Tipici risultati per compositi fibra-epoxy
Tipici risultati per compositi fibra-resina
Utilizzo del composito.
Tipologie di tessuti e laminati
Plain
Twill

Basket
Satin
Tipologie di tessuti e laminati
COMPOSITI AVANZATI
INTRODUZIONE AI MATERIALI FIBROSI
Parte II: ASPETTI GENERALI

Giacomo Frulla. Tel 0115646842–e- mail : giacomo.frulla@polito.it


DEFECTS IN COMPOSITES
1. Porosity
2. Prepreg gaps
3. Contamination (solvent, so1id, prepreg backing sheet)
4. Fibre alignment
5. Lay-up order
6. State of cure of the matrix
7. Fibre / resin ratio variations
8. Prepreg joints
9. Interply delaminations
10 Skin to core de-bonding
11. Resin micro-cracking
12. Damaged honeycomb core (interfaced with a composite)
13. Misplaced potting compound (interfaced with a composite)
14. Edge member/shear connection debond (interfaced with a composite)
DEFECTS IN COMPOSITES

In-service phenomena Defects sustained


Vibration/fatigue/creep Fibre fracture
Impact/shock/high rate loading Fibre de-bond
Lightning strike Matrix cracking, hole, local
overheating
Environmental cycling (heat, humidity, etc.) Delaminations/de-bonds
Bacterial/chemical degradation Delaminations, de-bonds,
loosening of mechanical properties
Galvanic corrosion Sandwich core crushing,
local damages
Poor maintenance (e.g. tool drop) and repair Sandwich skin delamination,
sandwich core crushing
Fig. : Example of manufacturing defect: backing paper embedded between plies

Defect in
adhesion.
POROSITY
Porosity is also known as resin voids and void content. It usually occurs as a
result of the application of pressure at too late a stage during the cure cycle,
which causes entrapment of air or vapour. This results in porosity throughout the
lay-up and is not usually confined to a restricted area. Contamination by dust
particles or other foreign objects increases the possibility of void nucleation.
Resin rich zones are also preferred sites for the formation of voids. A high
population of voids within the lay-up causes some reduction in the load transfer
capabilities of the lay-up, therefore affecting the resin dominated properties. Voids
also act as reservoirs for moisture retention and can have a marked effect on the
moisture Initiated changes in the lay-ups overall properties.
Porosity can occur in filament wound components when departures from correct
procedures occur. These include insufficient resin applied to the yarn, over-rapid
winding and incorrect curing. The presence of these voids can cause changes in
the mechanical and environmental properties of the composite.
POROSITY and voids
Prepreg Gaps
A sheet of unidirectional prepreg is produced by rolling and flattening several bundles of fibres
which are then impregnated. Either damage or inadequate monitoring of this process can
result in gaps in a fibre sheet, or less seriously, a disturbance of the fibre direction. These gaps
in as-delivered prepreg should be identified by quality control procedures, both by the
producer and the user, and are, therefore, not a major problem in practice.

Contamination
Contamination can occur in the form of dust (see Porosity), grease or solvent, or solids which
become incorporated into the laminate. This problem is normally controlled by careful
cleanliness procedures. Two types of contamination demand particular care. The first is sharp
solid debris such as metal swarf which may, if incorporated into a laminate, cause a risk of
significant fibre cutting and local damage. The second is the accidental inclusion of small
pieces of protective prepreg backing sheet Into a laminate. These can be left when prepreg is
cut to shape with subsequent removal of the backing sheet. Where backing sheet is present
between plies, no bonding occurs with a consequent drastic local change in laminate
properties.
Fibre Alignment
During prepreg manufacture and handling, fibres can become detached from the
prepreg and form either into whorls (during prepreg manu-facture) or misoriented
fibres (during lay-up or in the prepreg). These affect the local properties of the
laminate and introduce stress raisers and resin rich areas which constitute
preferential void forma-tion sites. Whorls are coils of fibre that are formed 1n the
prepreg sheet. They occur at the end of production prepreg runs and are rarely
encountered. Misoriented fibres are usually caused by the essential prepreg "tack" or
stickness causing fibres to be dragged during lay-up operations.

Lay-up Order
The number of plies and their relative orientations are determined by the design
strength and stiffness of a laminate. The same is true for the number and winding
angle of layers in a filament wound item. It is possible for a particular layer to be
either laid or wound at the incorrect angle to the other layers or even to be omitted
altogether. If either of these occur, the properties of the laminate will not be those
specified in the design.
Fibre and lay-up defects
State of Cure
This heading incorporates both the state of the resin determined by cure cycle and the state of
the resin as affected by variations in resin chemistry. It is included because both effects can
have signi-ficance for the final properties of the composite. The cure cycle consideration is
rare as a cause of faulty laminate production because procedures can be readily established
and auto-matically monitored. These procedures are usually the result of extensive
development using destructive and mechanical test programs.
Variations of resin chemistry within a single prepreg sheet can be assumed to be very small;
however, significant variations between batches have been found for some types of resins.
Prepreg properties should be within the tolerances laid down by the prepreg supplier and note
of these tolerances should be made when designing with the pre-preg. Although the chemistry
of a resin changes throughout storages as the "life" of the prepreg is used, the most rapid
changes occur during the cure cycle. Consequently to ensure even chemistry throughout’s
fabrication, it is necessary to ensure even curing in terms of tempera-ture distribution,
pressure etc. An important note here is that there is no practical non-destructive means of
checking the state of cure. Furthermore, even destructive quality control samples taken from
peripheral test areas cannot ensure uniformity across a laminated sheet.

Incorrect Cure
Prepreg Joints
Two types of prepreg joint can occur in any given layer. These are end-to-end joints (joint
perpendicular to the fibre direction) and side-by-side joints (parallel to the fibre direction). The
first case, end-to-end joints, are not recommended except possibly in the most uncritically
stressed area of a specimen. There is a sharp discontinuity of stress transfer at the joint; the
high mechanical properties of the fibre being lost.
Joints parallel to the fibre direction are much less critical to the integrity of the fabrication. They
should be arranged so that the two pieces of prepreg butt exactly together; an overlap causes
an extra ply to be formed locally with resin rich fillets; a gap results in a local lack of fibres and
a resin rich area.

Local Fibre/Resin Ratio Variations


The manufacture of prepreg utilises tows of fibres which are compounded with the resin. In
some cases an uneven distribution of fibres through-out the resin can result. With a sample
size of greater than 1 gram, this variation is usually no greater than ±2 or 3 % of the nominal
value.
Interply Delaminations
This class of defect is defined to be the failure, for any reason, of two plies within a laminate to bond
together. The most common cause of this is the inclusion of protective prepreg backing sheet during lay-
up - causing a complete delamination. Other causes can be entrapped air bubbles formed during lay-up
over complex curved surfaces.
Delaminations are particularly deleterious to a laminate as the whole basis of the strength and stiffness of
the composite structure is modi-fied, particularly with respect to compressive loads. The presence of
delaminations severely reduces the resistance of the laminate to buckling failure. Furthermore the fatigue
strength of the laminate is likely to be significantly affected as the delamination represents a crack-like
fatigue initiator.
If a delamination occurs in one face of a CFRP-faced honeycomb struc-ture, the interlaminar shear
strength is locally reduced to zero and this introduces a degree of asymmetry which affects the stability of
the whole structure.

Delaminations
and disbond

Fig. : Disbonds
Delaminations and craks

Matrix cracks

Resin Micro-cracks
These can be caused by processing during the cure cycle and further thermal cycling. Resin rich zones
are preferred sites for the nucleation of cracks within the resin but they can be created throughout the
whole cured laminate. Due to the nature of CFRP composite a differ-ential stress is inherent in the
material and it is believed that micro-cracking is the mechanism by which this stress is relieved. As the
resin is a plastic the rate of loading will affect the strain at which the resin breaks, therefore rapid
temperature changes are more likely to cause micro-cracks than slower temperature changes. It has also
been found that resin starvation can give rise to micro-cracking in the direction of the fibres. Whilst resin
micro-cracking is essentially a defect in the matrix, it may have either beneficial or adverse effects
depending on the require-ments for the composite. There are instances where the presence of micro-
cracks provides stress relief in the composite, thus providing a beneficial effect.
Cracks and disbonds

Skin to Core Debonding


This is a particularly deleterious defect which affects composite skinned/honeycomb panels as the
mechanical properties of the panel are seriously degraded in the region of the disbond, particularly with
respect to buckling failure. This defect can be caused by contamination of, or damage to, the film adhesive
used to join the face laminate to the core, contamination of the face laminate itself, or contamination of the
bond area of the core. In addition, mechanical damage resulting in either crushing of the core cell walls or
local damage to the cell walls can result in a debond between the face laminate and core. If a face skin is
bonded to a crushed core such that an indentation is formed in the skin, a shifting of the neutral axis will
occur with consequent changes in stability and bending stiffness.
Skin to Core Disbonding
Damaged Honeycomb Core
There are two types of damage which can occur in the honeycomb core. The first is crushing damage in
which the cell walls are depressed such that there is a depression in the honeycomb surface. The result of
such a depression is at best a thickening of the bond line between the skins and the core and, at worst, a
total disbond in which skin, adhesive and core never make contact. The second type of damage occurs
when the core material is formed over a curved mandrel. The curvature sometimes becomes concentrated
along a line of cells instead of being distributed over the surface. This causes a discontinuity in the surface
which again causes bond line thickening or possibly a disbond.

Misplaced Potting Compound


Where fixings have to be made to a honeycomb sandwich panel, the crushing strength of the panel is
locally Increased by filling the honeycomb cells with a potting compound. This is normally done before
integration of skins and core and it is possible that the potted region can be wrongly positioned such that,
in the extreme, the fixation misses the strengthened area altogether. It is therefore necessary to ascertain
the position of the potting before fixation takes place.
Edge Member Debond
There are two types of edge member: the structural and the non-structural. Non-structural edge members
are for cosmetic and handling protection purposes and are used on free edges only. Commonly known as
'edge-closeout’, this consists usually of vented tape or foamed plastic. They have little effect on the
performance of the composite. The tape edge member can reduce surface quality and may have a
detri-mental effect on a paint finish. The foamed plastic avoids both these problems but does incur a
weight penalty. Structural edge members are designed to join composite honeycomb sandwich
components to other parts of the structure. The joint between the honeycomb sandwich and the edge
member may have to transfer the loads from bolts, rivets or adhesive joints and distribute them
through-out the face skins. The shear components of these loads are transferred from the edge member
to the honeycomb core by means of an adhesive bond lying parallel to the core cell walls. This bond is
usually effected by means of a foaming adhesive. Defects in this bond invariably involve large voids (or to
a lesser extent distributed porosity) in the adhesive layer. The layer can be several millimeters in thickness
and several hundred millimeters in lateral extent. There is thus the possibility for large voids to occur. The
presence of these voids affects the load transfer and stress distribution properties of the composite.
Environment
IMPACT
Principles for achieving damage tolerant composite structures
In order to make composites cost effective, allowable damage limits must be set at their maximum
allowable size while still meeting regulatory ultimate load requirements. To achieve this goal, test data and
analytical methods encompassing the complete range of potential damage sizes and types are required.
Damage tolerance assessments involve the application of known damage threats to the structure during
its typical service usage and demonstration that this damage will not alter the safe operation of the
structure. Damage tolerance builds upon fail-safe analysis by including determinations of damage growth
characteristics and by establishing damage detection methods and inspection plans. A no-growth
approach has been used to demonstrate compliance with damage tolerance requirements for primary
structures
Barely visible impact damage (BVID) is defined as damage that establishes the strength design values
to be used in analysis demonstrating compliance with the regulatory ultimate load requirements. The
extent of such damage is established prior to the design phase. The term visible is used since the primary
inspection method in current use involves visual observation.
Allowable damage limits (ADL), defined as damage that reduces the residual strength to the regulatory
ultimate load requirements, are determined to support maintenance documents. Given that the structure’s
strength with BVID damage will result in positive margins at design ultimate load, the corresponding ADL
will be larger than the BVID. Characteristics describing the detectability of the ADL as well as the type and
extent of the damage are documented to support maintenance programs.

Maximum design damage (MDD) is defined as damage that establishes the strength design values to be used
in analysis damage tolerance requirements. The extent of such damage is established prior to the design phase.
Critical damage thresholds (CDT) are defined as damages that reduce the residual strength to the regulatory
requirements. Given that the structure’s strength with MDD size damage will result in positive margins at design
limit load (DLL), the corresponding CDT will be larger than the MDD. Characteristics describing the detectability
of the CDT as well as the type and extent of damage are documented to support the establishment of required
inspection methods and intervals. Using the selected inspection technique, realistic damages smaller than the
corresponding CDT are shown to be detectable with high probability before any growth causes it to exceed the
CDT.
Readily detectable damage (RDD) is defined as damage that can be detected within a small number of load
cycles. For damage that is not readily detectable, the structure should be evaluated for all possible damage
growth mechanisms. The maximum extent of damage that is considered readily detectable, but which is not
immediately obvious, should be established.
Damages larger than the maximum readily detectable damage are considered to be immediately obvious.
Established parameters:
Established parameters:

It can be seen that only a small band of the residual strength curve is defined. Since
residual strength of in-service damage sizes are not available, often BVID sizes are
used for ADL sizes in structural repair manuals. This are conservative sizes and can
result in increasing maintenance costs.
The following are recommended approaches for developing data to support certification and to
allow for reduced maintenance costs of composite structures:
- the residual strength curve for each significant type of potential damage on each principal
structural element should be determined by analysis and/or test;
-characteristics describing the inspectability of the CDT as well as the type and extent of the
damage should be documented to support maintenance planning activities;
-for readily detectable damage, the magnitude of the threats that should be considered, should
include impact damage by ground vehicles and ground handing equipment, impacts, etc. Service
experience has shown that damage associated with such events may persist before the damage
is detected and the structure repaired. The extent of damage that should be considered must be
established by taking into account susceptibility to each type of accident.
-Structural damage design should be coupled with development of the eventual maintenance
plan in order to reduce damage occurrences and costs. Test validation and analyses should
address design ultimate strength, damage growth, residual strength and maintenance issues for
composite structures. Independent studies of design ultimate load or limit load strength without
data and analyses at intermediate load levels will not provide a balanced design that supports
cost effective maintenance, when maintenance is possible.
REPAIR
Damaged composite structures are definitely repairable, shown clearly in the before and after images
below.
However, there are challenges:
Hidden damage issues, including manufacturing defects. (for example, a low velocity impact, which
normally wouldn’t cause much damage may cause a sandwich structure to disbond between the skin and
core due to poor adhesion during manufacture. If this disbond is the only damage, there may be no visible
trace of it from the surface.)
Unexpected damage sources. (for example, an aircraft vertical tail part may be designed to withstand
hailstone impact but not able to resist damage from being dropped during shipping or removal for
inspection)
“Best” repair techniques are heavily dependent on details of the structure. In other words, because
composites excel at being tailored to meet very specific needs, there are few “universal” materials and
methods that can be used to achieve successful results. Composite repair specifics really have to be
determined on a case-by-case basis.
COMPOSITI AVANZATI
INTRODUZIONE AI MATERIALI FIBROSI
Parte III: Riparazioni

Giacomo Frulla. Tel 0115646842–e- mail : giacomo.frulla@polito.it


Basic Repair Process

The very basic fundamentals of composite repair include the following steps:
Inspect to assess damage (extent and degree)
Remove damaged material
Treat contaminated material
Prepare repair area
Complete composite repair
Inspect repair for quality assurance (e.g. delaminations, inclusions,
proper cure, etc.)
Restore surface finish
Damage assessment

Considerations,
practicalities and
parameters of composite
repair:
Types of Repair
Basic types of composite repair include the following:
Cosmetic A superficial, non-structural filler is used to restore a surface to keep fluids out until a
more permanent repair is made. This type of repair will not regain any strength and is used only
where strength is unimportant. Due to high shrinkage, cosmetic repairs may start to crack after
a relatively short time in service.
Resin Injection This type of repair can be effective in limited instances, where the delamination
is restricted to one ply. However, not much strength is regained, and the primary benefit is that it
is quick and cheap. At best, this type of repair can hope to slow the spread of delamination and
is generally considered a temporary measure. :

Structural Mechanically–fastened Doubler


Full structural repairs using bolted doublers can be
used in heavily loaded solid laminates. This is
often the only practical means of repairing such
structures. However, such repairs are not
aerodynamically smooth, and may cause
“signature” problems in structures where low-
observability by radar is required. They also leave
the original damage and simply attempt to transfer
loads around the damage. Finally, they can create
stress concentrations at their corners and edges.
Structural Bonded External Doubler
Bonded external doublers are often used to perform repairs to lightly loaded thin laminate
structures. This type of repair is especially common using wet lay-up materials. They may be
room-temp or high-temperature cured, depending on the matrix resin system used. These repairs
can regain a significant portion of the original strength of the structure—or even full strength—
although with a significant stiffness and weight penalty in many cases. This type of repair is
generally easy , relatively quick and does not require the highly developed skills of flush structural
repairs.

Structural Flush Repair


This repair restores full structural
properties by forming a joint between the
prepared repair area and the repair patch.
The repair patch is made by replacing
each ply of the composite laminate that
has been removed from the damage
area. The size of the repair patch should
fit exactly the area prepared for repair,
except for a final cosmetic or sanding
layer, which is often slightly larger to allow
for sanding down to achieve a smooth
and/or cosmetic surface.
Bolted or Bonded
Which is better? This depends on the definition of “better” for each specific case. For example,
“better” for some repairs may be lighter, while for others it may be heavier but longer lasting. In some
cases cheaper is “better”, while in others producing the repair quickly is the goal.
Bonding usually excels with very thin structures—bonding two thin pieces of aluminum is much stronger
than bolting them—and with dissimilar structures—one can glue a piece of paper to a steel plate which
isn’t feasible with mechanical fasteners like bolts or screws. Other advantages to bonding include:
Bolted joint and repair strengths increase as the substrates get thicker. This isn’t true for bonding,
except for very flexible materials. Other advantages of bolting include:
Doesn’t require meticulous surface preparation
Easy to inspect for quality Easily disassembled

Bonding minimizes corrosion. Adhesives also make good sealants. No fastener holes to weaken
structure. No point stress concentrations. Smooth surface finish.
Therefore, a basic rule for achieving the highest strength composite repair is to use bonding for
thin laminates and bolting for thick laminates. Thickness here refers to the original laminate or
skins, excluding any core materials. The crossover point varies tremendously with the specific
details of each case, but in general 1/64inch (0.4mm) is considered thin and 1/2inch (12.5mm) is
somewhat thick. Note that parts with an aerodynamic or cosmetic surface usually require a flush
repair, which almost necessitates bonding. However, when bonding is not possible, a bolted
temporary or permanent repair may be necessary.
Scarfing
After completing initial damage removal, the area around the repair must be prepared. The
corners of the repair hole must be rounded off and the hole itself should be tapered to provide the
best load transfer when the repair patch is bonded in. Scarfing, or taper sanding, is usually
achieved using a compressed-air powered high-speed grinder. This is a gentle process, which
prepares the damaged area for application of a repair patch. It is imperative to follow all repair
manual guidelines, and significant skill and practice on the part of the repair technician is
mandatory. Note: If the damaged area exceeds allowable repair limits in applicable repair
manuals, then specific engineering support is required in order to proceed with the repair.

Scarfing : Ply Determination


For scarfing, it is important to know the number of plies in the composite laminate. The ply
orientations are not needed for preparing the damaged area but will be needed for cutting repair
materials and fabricating the repair patch.
Scarf Angles
A crude rule of thumb for the amount of material to remove during scarfing is to taper sand
approximately 1/2 inch (12.5mm) of area per ply of composite laminate. Typical scarf distances are
from 20 to 120 times the thickness of the laminate being scarfed.
The material removed during scarfing is often referred to by angles, rather than distance per ply. The
flatter the scarf (more area per ply), the larger the adhesive bond, and the lower the load on the bond.
The steeper the scarf, the less undamaged material is removed. Lightly loaded structures may be able
to tolerate a smaller, steeper scarf. Heavily loaded structures usually require a larger, more gentle
scarf, such as 20:1 up to as much as 100:1.
Scarfing : Scarfing vs. Stepping
Stepping is an alternate method for removing material in preparation for applying a repair patch. In
stepping, the overall angle is achieved by removing a precise area of material per ply of composite
laminate. Both methods work. Most consider scarfing to be easier, and it is generally considered to be
better. Stepping leaves abrupt edges and butt joints in each repaired ply. It is also hard to do without
cutting through and damaging the underlying plies.
The Repair Patch
The edges of the repair patch should be tapered and all plies should have rounded corners. The repair
patch is attempting to replace the damaged area in the composite laminate exactly, restoring it as much
as possible to original. Thus, the number of plies and orientations of each ply must match, layer for
layer, that of the original structure.
So, in an exact ply-by-ply replacement, will the repaired structure be as strong as the original? No.
Depending on many details, the repaired structure is typically about 60-80% as strong as the original
undamaged structure.

Is it possible to make a repaired structure as strong as the original? Yes. However, extra repair plies
must be added to compensate for the loss of strength caused by the repair. This means the repair will
not be perfectly flush, and also that the repaired structure will be stiffer than the original.
Is the extra stiffness a problem? If the original structure is not stiffness critical but is primarily loaded in
straight tension or compression, then a stiffer repair will most likely be fine. However, if the structure
flexes significantly under load, the “stiff spots” caused by a full-strength repair can cause failure at the
edges of the repair. Some repairs may therefore need to be deliberately under-strength, in order to
match the stiffness of that original area in the structure.
Scarfed Repair
Below is a review of the general outline for steps in performing composite repair, but with
detail added specifically for a scarfed repair with flush bonded repair patch.
1. Inspect for extent of damage: Visual , Tap, Ultrasonic, and/or X-rays
2. Get best access possible, both sides if feasible.
3. Remove all damaged and delaminated material. Circular or rounded corners
4. Grind away scarf angle taper: Smooth, flat ground surface
5. Determine ply orientations and materials of original structure.
6. Replace plies: Adhesive layer first , One or more filler plies, orientation not important ,
Repair plies - match orientations with original structure ;Extra plies - usually orientation
matches original outer ply ;Often an outer adhesive layer
7. Vacuum-bag and cure repair plies as required
8. Trim to net edge dimensions after cure, if necessary.
9. Inspect repair for delaminations, inclusions, proper cure documentation, etc.
10. Sand and finish as required. Do not sand into fibers of repair plies.

Potrebbero piacerti anche