2A Developments in Machining of Stacked

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 26

This article was downloaded by: [Johns Hopkins University]

On: 01 January 2015, At: 07:23


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Machining Science and Technology: An


International Journal
Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/lmst20

Developments in Machining of Stacked


Materials Made of CFRP and Titanium/
Aluminum Alloys
a b
Alokesh Pramanik & Guy Littlefair
a
Department of Mechanical Engineering, Curtin University, Bentley,
Western Australia, Australia
b
Department of Engineering, Deakin University, Waurn Ponds,
Victoria, Australia
Published online: 05 Nov 2014.
Click for updates

To cite this article: Alokesh Pramanik & Guy Littlefair (2014) Developments in Machining of Stacked
Materials Made of CFRP and Titanium/Aluminum Alloys, Machining Science and Technology: An
International Journal, 18:4, 485-508, DOI: 10.1080/10910344.2014.955718

To link to this article: http://dx.doi.org/10.1080/10910344.2014.955718

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the
“Content”) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or
howsoever caused arising directly or indirectly in connection with, in relation to or arising
out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &
Conditions of access and use can be found at http://www.tandfonline.com/page/terms-
and-conditions
Downloaded by [Johns Hopkins University] at 07:23 01 January 2015
Machining Science and Technology, 18:485–508
Copyright 
C 2014 Taylor & Francis Group, LLC

ISSN: 1091-0344 print /1532-2483 online


DOI: 10.1080/10910344.2014.955718

DEVELOPMENTS IN MACHINING OF STACKED MATERIALS MADE


OF CFRP AND TITANIUM/ALUMINUM ALLOYS

Alokesh Pramanik1 and Guy Littlefair2


1
Department of Mechanical Engineering, Curtin University, Bentley, Western Australia,
Downloaded by [Johns Hopkins University] at 07:23 01 January 2015

Australia
2
Department of Engineering, Deakin University, Waurn Ponds, Victoria, Australia

2 The aim of this article is to investigate the drilling of carbon fiber-reinforced plastic (CFRP)
composite/metal stack-ups to have a details picture of the developments in this complex area. The
forces and torque, chip shape, surface finish and geometry, and tool material and tool wear for
drilling composite/metal stack-ups have been analyzed in details in addition to drilling mechanism
of CFRP. The relation between input and output parameters was discussed and the trend of input
parameters for damage free and tight tolerance holes has been investigated based on the literature.
The main findings are (i) heat, built-up edge and chips generated from drilling of metallic layers
damages CFRP surface, (ii) order of material layers affects the drilling outcomes significantly, (iii)
coatings and step-shape on the cutting tool improves the tool performance, (iv) tool materials should
be selected based on the material of metallic layer, (v) chipping, adhesion, abrasion and attrition
are main tool wear mechanisms during machining of CFRP/metal stacks and (vi) application of
coolant improves the machinability.

Keywords aluminum, CFRP, drill tool, titanium, stacks, wear

INTRODUCTION
Composite/metal stack-ups provide an efficient method to increase
bending rigidity without a significant increase in structural weight. Com-
posite means carbon fiber-reinforced plastics (CFRP) and metal indicates
Ti or Al in this article. The composite panels in commercial aircraft
(Airbus A380 or the Boeing 787) are arranged in the form of a sandwich-type
stacking: CFRP/CFRP, CFRP/Aluminum, CFRP/Titanium, etc. (Krishnaraj
et al., 2010). The potential applications for these structures are in trans-
port aircraft components (e.g., control surfaces, engine cowlings, fairings,
and fixed trailing edge wing panels), helicopter blades, optical benches for

Address correspondence to Alokesh Pramanik, Department of Mechanical Engineering, Curtin


University, Bentley, Western Australia 6102, Australia. E-mail: akprama@yahoo.com
Color versions of one or more of the figures in the article can be found online at
www.tandfonline.com/lmst.
486 A. Pramanik and G. Littlefair

space applications, nonferrous ship hulls, etc. Sandwich structures will also
be used prominently in future aerospace applications such as Raytheon’s
Premier I, Lockheed-Martin’s X-33, and future tilt rotorcraft by Textron-
Bell Helicopter/Boeing (Campbell, 2006).
Assembling of composite and metal stack-ups represents a significant
portion of the total manufacturing cost. Assembly operations are labor in-
tensive and involve many steps and can represent as much as 50% of the
total delivered part cost (Askeland, 1989). Mechanical bolting is gener-
ally used to assemble components in aerospace industries where compos-
ites and metals are fasten together to get final parts. It facilitates assem-
bling/dismantling of different parts for repairs and/or maintenance but it
needs to have through holes for mechanical bolting (Shyha et al., 2011).
Downloaded by [Johns Hopkins University] at 07:23 01 January 2015

The number of mechanical fasteners in a typical fighter aircraft might be


in the range of 200,000–300,000, while a commercial airliner or transport
aircraft can have as many as 1,500,000–3,000,000 fasteners, depending on
aircraft size (Campbell, 2006). Thus, holes are drilled through a combina-
tion of materials, such as, CFRP/CFRP, CFRP/Al, CFRP/Ti, for each of these
fasteners.
The multi-material stack-ups need to clamp securely together as inter-
faces are not perfectly bonded during drilling. Then the fasteners are in-
stalled in each hole. Since the joint is the weakest part of any structure, its
design is a critical process and carried out with the utmost care by drilling
good quality holes. Currently in industries, holes in multilayered metal-
lic/composite stacks are produced by drilling each material layer separately
followed by temporary assembly of the workpiece plates for subsequent de-
burring and finishing (Shyha et al., 2011). Sometimes, undersize holes (pilot
holes) are drilled manually during assembly then temporary fasteners are
installed to hold the parts together. The holes are then brought to full size
after all of the pilot holes are drilled. Thus complexities are still existed
in meeting tolerances and increased productivity. Single shot drilling has
the potential to solve the problem; however, the widely different mechan-
ical/physical/machining properties of the stack materials pose significant
challenges (Shyha et al., 2011).
The efficiency of the bolted joints largely depends on the quality of the
drilled holes (Mathew et al., 1999). There are many types of drill motors
and units that can be used to drill structures, but they can be broadly clas-
sified as either hand, power feed and automated drilling units (Campbell,
2006). Many variations of twist drills(Figure 1) are used in drilling metallic
structure. Since drill bit geometries can influence both quality and quan-
tity of holes drilled, different geometries are available. Examples of typical
variations in the standard twist drill include, step drills that drill an under-
sized hole and then the final hole size in one pass, and drill-reamers in
which a hole is drilled and then reamed in the same pass (Zhang et al.,
2008). When drilling aluminum, standard high speed steels give satisfactory
Machining Stacked Materials Made of CFRP and Alloys 487
Downloaded by [Johns Hopkins University] at 07:23 01 January 2015

FIGURE 1 A typical twist drill tool.

drill life. For harder materials, such as, titanium, the cobalt grades of high
speed steel yield longer lives. Carbide drills give even longer life in tita-
nium, but are more prone to chipping on the cutting edges (Zhang et al.,
2008).
The machining parameters for an optimized drilling of composite and
metal materials are not similar. The carbon fiber-reinforced plastics (CFRP),
owing to their anisotropy and abrasive nature of their carbon fiber content,
exhibit totally different drilling results as compared to those of drilling com-
mon metals and other materials (Faraz et al., 2009). In addition, relatively
high sensitivity to heat damage and weakness in the thickness direction make
drilling composites much different than that of metals (Astrom, 1997). The
smoothness of the drilled surface and tool wear are equally important in
drilling composites. However, the smoothness of the drilled surface is pri-
oritized over the tool wear given the difficulty to drill laminate without
producing unacceptable cracks (Dharan and Won, 2006). While standard
twist drills are used for drilling CFRP structure, a number of unique drill
geometries have been developed for composites, several of which are shown
in Figure 2. In addition to these, saw drill, candle stick drill, core drill, step
drill, trepanning drill and orbital drill are also used for drilling composites
488 A. Pramanik and G. Littlefair

FIGURE 2 Composite Drill Configurations (Campbell, 2004). © Elsevier. Reproduced by permission


of Elsevier. Permission to reuse must be obtained from the rightsholder.
Downloaded by [Johns Hopkins University] at 07:23 01 January 2015

(Brinksmeier et al., 2011; Hocheng and Tsao, 2003; Tsao and Hocheng,
2005).
Thus, single shot drilling of the CFRP-Titanium/Aluminum stack-ups
is challenging in terms of the geometry of the drill tools and machining
parameters. There are researches on this topic which resulted considerable
improvements in this area. The literature review described in (Krishnaraj
et al., 2010) missed the main focus on this topic and fail to direct the re-
search community in the right direction. Thus an innovative, systematic and
scientific literature review is still missing. This study clarifies all the available
information systematically and finds the trends in design of single shot drill
tool and machining parameters. The drilling of CFRP alone is very com-
plex. Thus, the problem related to drilling of CFRP is briefly discussed first
then focus is given to the drilling of CFRP-Titanium/Aluminum stack ma-
terial. This will provide a practical understanding and direct the research
community to right direction in the field of single shot drilling of CFRP-
Titanium/Aluminum stack material.

DRILLING CFRP
Every composite material possesses complex deformation behavior,
which leads higher processing cost of these types of materials (Pramanik
et al., 2007, 2008). Challenges in drilling CFRPs could be classified on the
one hand as the excessive tool wear, while on the other hand as workpiece
material-related problems (Enemuoh et al., 2001, Langella et al., 2005). The
latter ones include surface irregularities and defects like material cracking
and delamination (Faraz et al., 2009). Composites are very susceptible to
surface splintering (Figures 3 and 4), particularly if unidirectional material
is present on the surface. Splintering can occur at both the drill entrance
and exit of the hole (Astrom, 1997). As shown in Figure 5, when the drill
enters the top surface, it creates peeling forces on the matrix as it grabs
the top plies. When it exits the hole, it induces punching forces that again
creates peel forces on the bottom surface plies. If top surface splintering
Machining Stacked Materials Made of CFRP and Alloys 489

FIGURE 3 Composite hole splintering (Campbell, 2004). © Elsevier. Reproduced by permission of


Elsevier. Permission to reuse must be obtained from the rightsholder.
Downloaded by [Johns Hopkins University] at 07:23 01 January 2015

FIGURE 4 Delamination and quantification of delamination (Faraz et al., 2009). © Elsevier. Repro-
duced by permission of Elsevier. Permission to reuse must be obtained from the rightsholder.

FIGURE 5 Classification of delamination (Faraz et al., 2009). © Elsevier. Reproduced by permission of


Elsevier. Permission to reuse must be obtained from the rightsholder.
490 A. Pramanik and G. Littlefair

is encountered, it is usually a sign that the feedrate is too fast, while exit
surface splintering indicates that the feed force is too high (Paleen and
Kilwin, 2001).
Mainly three significant damage mechanisms cause delamination: (i)
plate bulge, (ii) spall opening and (iii) spall tearing/twisting (DiPaolo et al.,
1996). A bulge is described as an opening mode of fracture produced by
distributed load acting directly from the chisel edge onto the plate. After the
chisel edge exits the laminate, pieces or spalls from this exit surface of the
laminate start breaking off. This phenomenon occurs via spall opening. The
downward thrust promotes cracking via an opening mode of fracture. This is
produced by a point load at the point of contact of the cutting lip and spall.
The third type of damage is spall tearing and twisting, whereby the spalls
Downloaded by [Johns Hopkins University] at 07:23 01 January 2015

are subjected to a tearing fracture mode as a result of the drill torque and
twisting of the spall due to the combination of the downward-thrust-force
direction and the back-rake angle along the cutting lips.
At a stage, when only the cutting lips machine, the cracks become larger
than the drill radius. When the cutting lips complete their machining, the
flutes act on the spalls with sufficient energy to propagate the crack even fur-
ther. The crack propagation continues until it becomes stabilized (DiPaolo
et al., 1996). The CFRP is sheared continuously and smeared into the gaps
among the fibers at low feeds but the cutting mechanism at higher feeds is
described by compression induced rupture normal to the fibers and shear
fracture along the fiber/matrix interface due to bending (Kim and Ramulu,
2005).
A layer of fabric is often cured on both surfaces of composite parts to
reduce the splintering problem as woven cloth is much less susceptible to
splintering than unidirectional material. In addition, a back-up material,
such as, aluminum or composite, clamped to the backside, reduces back-
side hole splintering (DiPaolo et al., 1996). The generation of heat needs
to be minimized during drilling as the epoxy matrix composites start to
degrade at temperature around 204.4◦ C. Typical drilling parameters are
2000–3000 rpm at feed rates of 0.1–0.127 mm/rev, although this will vary
depending on the drill geometry and the type of equipment used. The de-
sign of the drill tools and the drilling procedures are very dependent on
the materials being drilled. For example, carbon and aramid fibers exhibit
different machining behaviors, and therefore require different drill geome-
tries and parameters. The tool geometry related damages are associated to
the angle between fibers orientation and the cutting edge.
In general, temperature related damages appear as a result of friction
between the drill and the wall of the hole (Abrão et al., 2007). The flat
two-flute and four-flute dagger drills were developed specifically for drilling
stack-ups of carbon/epoxy. The two-flute and four-flute variety normally
run at 2000–3000 rpm (for 6 mm diameter tool) and 18000–20000 rpm,
respectively (Paleen and Kilwin, 2001). The low compressive strength of
Machining Stacked Materials Made of CFRP and Alloys 491

aramid fibers results in fuzzing and fraying during drilling as the fibers move
away rather than being cleanly cut. The drill tools with a “C” type cutting edge
reduce fuzzing and fraying when aramid fiber is used as reinforcement. The
“C” type cutting edge grabs the fibers on the outside of the hole and keeps
the fibers in tension during the cutting process. Typical drilling parameters
for aramid fiber composites are 5000 rpm (for 6 mm diameter tool) and a
feed rate of 0.025 mm/rev (Bolt and Chanani, 1987).
Although standard high speed steel (HSS) drills work well in glass and
aramid composites, the extremely abrasive nature of carbon fibers requires
carbide drills to obtain an adequate drill life. For example, a HSS drill may
only be capable of drilling one or two acceptable holes in carbon/epoxy,
while a carbide drill of the same geometry can easily generate 50 or more ac-
Downloaded by [Johns Hopkins University] at 07:23 01 January 2015

ceptable holes. For drilling carbon/epoxy in rigid automated drilling equip-


ment, polycrystalline diamond (PCD) drills have exhibited outstanding pro-
ductivity improvements (Bolt and Chanani, 1987). Although PCD drills are
very expensive, the number of holes obtained per drill and the fewer changes
required make them cost effective (Pramanik et al., 2009). It should be noted
that PCD drills cannot be used with freehand or non-rigid setups; the tip of
the tool will immediately chip and break if any vibration or chatter is present
during drilling (Bolt and Chanani, 1987).

DRILLING STACK MATERIAL


Drilling and fastening of hybrid materials in one-shot operation reduces
time of assembling structures. The most common problem faced in one-
shot drilling and riveting of stack is the curl up of long chips on the body of
the tool. The drilling of CFRP is manageable in the stack if the aluminum
or titanium layers are below the CFRP. But the hot and continuous chips
produced during drilling considerably damage the CFRP hole (Zitoune et al.,
2012). The quality parameters, such as, geometric accuracy, cylindricity,
surface texture, and formation of metal burrs, are significantly deteriorated
due to the dissimilar machining characteristics of stack materials (Kim and
Ramulu, 2004). When drilling through composite-to-metal stack-ups, the
drill geometry is usually controlled more by the metal than the composite
and standard twist drill geometries are often used.

Force and Torque


There are significant changes in force generation while drill tools moves
from one layer to next layer of stack materials due to difference in specific
cutting pressures generated between the drill and the workpiece materials.
This specific cutting pressure depends of material of the layer and drill tool
geometry. This sudden increase affects the performance of the drill and
quality of the hole in CFRP (Zitoune et al., 2010).
492 A. Pramanik and G. Littlefair

In 4.2 mm CFRP/ 3 mm Al stacks, the magnitude of thrust force and


torque during drilling of Al is double at lower feed rate (0.05 mm/rev) but
at higher feed (0.1 and 0.15 mm/rev) it is approximately three times higher
than that of CFRP. The force and torque for aluminum can be much higher
with the increase of drill tool diameter. It is noted that 8 mm drill generates
higher thrust force and torque than that of 4 and 6 mm tools during drilling
of CFRP/Al stacks. This is because of the steep increase in cross sectional
area of the chip and greater chisel edge length at higher diameter. For
8 mm of diameter the chisel edge length is around 1.602 mm and for 4 mm
of diameter it is 0.6 mm (Zitoune et al., 2010). The thrust force due to the
chisel edge is 40% of the total thrust force at lower feed rate but it can be
60% when the feed rate is higher (Won and Dharan, 2002). Thus a drill
Downloaded by [Johns Hopkins University] at 07:23 01 January 2015

diameter of 6 mm or less is preferable to drill multi-material stacks (Zitoune


et al., 2010).
In drilling 4.2 mm CFRP/ 3 mm Al stacks, thrust force rises steadily up
to drilling 30 holes then it is stable until 60 holes after that thrust force rises
steadily again at speed 2020 rpm and feed 0.1 mm/rev while a 6 mm plain
carbide (K20) drill is used (Zitoune et al., 2010). The increase of force with
the number of holes drilled indicates the progression of tool wear. Higher
wear means lower sharpness of the drill tool which causes tendency to plough
rather than cut. A coated (2.32 µm thick CrAlN /a-Si3 N4 -Tripple Alwin) tool
experiences lower tool wear compare to uncoated one. This result in lower
thrust force and less increase of thrust force with number of holes for a
coated tool. For coated tool, thrust force increases to 72% (115–198 N) after
drilling 70 holes; on the other hand this increase is around 92% (from 142 N
to 278 N) for uncoated tool (Zitoune et al., 2012).
The thrust force during drilling of composites with uncoated tool is
20–25% larger than those recorded during drilling with a coated tool. During
drilling the aluminum layer this difference can be reached to 47%. This
indicates that the coating on tools largely reduced (i) the friction between
the body of the drill and the machined surface, (ii) the friction between the
chips and the flutes of the cutting tool (rake face) and (iii) tool wear by
protecting the tool from being worn out (Zitoune et al., 2012). The spindle
speed has little influence on the thrust force when coated tool is used.
A slight decrease in thrust force in the aluminum (around 5%) is noted
during drilling with uncoated tool at a constant feed rate of 0.15 mm/rev.
The reduction of the thrust force is around 10% while drilling CFRP with
the uncoated tool. This is due to thermal softening of aluminum and CFRP
with the increase of the spindle speed. Thus, the temperature generation
in case of uncoated tool is higher than that of a coated tool. It seems that
coating reduces the friction as well as improves the thermal conductivity of
the tool (Zitoune et al., 2012). The thrust force and torque during drilling
of CFRP/Al stack increase with the increase of feed rate. This is mainly due
to higher amount of the fibers/material are cut with time as the feed rate
Machining Stacked Materials Made of CFRP and Alloys 493

is increased while drilling CFRP. In addition higher feed reduces effective


clearance angle of the drill which generates rubbing of tool against the
CFRP/Al stack resulting in higher thrust and torque (Velayudham et al.,
2005).
During drilling of 7.54 mm CFRP/ 6.73 mm Ti stacks, the drilling forces
of the WC and PCD drills gradually increase in both CFRP and Ti due to tool
wear. The drilling forces for WC drill at the higher spindle speed increased
much faster than those at the lower spindle speed. The drilling forces do
not increase that significantly with hole number for the PCD drill (Park
et al., 2011). Figure 6 represents a typical force and torque profiles with
time along the depth of stack with carbide tool. These profiles are not steady
which indicate that tool wear, built-up edge and chip clogging are affecting
Downloaded by [Johns Hopkins University] at 07:23 01 January 2015

the drilling process.

Chip Breakability
Chip breakability can be defined as the number of chips in a certain
weight of chip (Zitoune et al., 2010). In the machining operation, small well
broken chips are desirable (Nomani et al., 2013). Curled chips, entangled
on the body of the drill, cause two main problems. It makes the automatic
drilling and riveting impossible, and increases the probability of damaging
the hole entry of the composite as well as that of the wall (Zitoune et al.,
2012). Larger chips produce a consistently rougher surface finish (Nomani
et al., 2013) and counter boring. Cutting parameters influence the shape and
size of the chips of metallic alloys and CFRP significantly. The low feed rate
(0.05 mm/rev) and speeds: 1050–2750 rpm (for 6mm diameter tool) pro-
duce continuous chips and do not influence the shape and size of the chips.

FIGURE 6 Force and torque profiles with time along the depth of stack at tenth hole with carbide tool,
Feed: 0.0762 mm/rev (CFRP), 0.0508 mm/rev (Ti), Speed: 6000 rpm (CFRP), 800 rpm (Ti) (Park et al.,
2010). © Elsevier. Reproduced by permission of Elsevier. Permission to reuse must be obtained from
the rightsholder.
494 A. Pramanik and G. Littlefair

With the increase of feed rates (from 0.1 mm/rev to 0.15 mm/rev)
chips are broken and leads higher thrust force and surface roughness of
the sandwich structure (Zitoune et al., 2012). During drilling 10 mm Al/
10 mm CFRP/ 10 mm TiAl6V4 or 20 mm CFRP/ 10 mm Al stack, chips
removal through the chip grooves and the adhesion of titanium chips on
the cutting edges erode and delaminate CFRP-layer which lead damage of
hole surface and decreased process reliability. For both of the materials
chips from the metal layers are long at speed: 10, 20 m/min, feed: 0.15 mm
and tool diameter 16 mm (Brinksmeier and Janssen, 2002). Figure 7 shows
typical chip removal process during drilling and long chips generated from
titanium layer. Thus the surface damage of the upper layers of stack is likely
when the chips from the lower layers are longer as shown in Figure 7.
Downloaded by [Johns Hopkins University] at 07:23 01 January 2015

The diameter of drill and feed rate affect the chip breakability signif-
icantly as the cross sectional area of chip increases with the increase of
diameter and feed rate whereas effect of spindle speed seems to be smaller
(Zitoune et al., 2010). Figure 8 shows the shapes of the Al chips during
drilling of 4.25 mm CFRP/3 mm aluminum stack at different machining
conditions. It clearly indicates that higher speed, higher feed and low or high
depths of cut generate highly expected short chips. For 4.25 mm CFRP/3 mm
aluminum stack, the shape and size of the chips depend also on the cutting
tool. Drilling at a feed rate of 0.1 – 0.15 mm/rev and a spindle speed of
2020 rpm (for 6mm diameter tool) with nano-composites (2.32 µm thick
CrAlN /a-Si3 N4 -Tripple Alwin) coated tool (6 mm micro grain carbide K20
with 132◦ point angle) gives broken chips and better results compared to
uncoated drills (Zitoune et al., 2012).
Titanium chips are long at low speeds, and became shorter with the
increase of feed. CFRP chips are long at low feeds but at higher feeds dust-
like chips are generated. Short aluminum chips are generated at higher
speed, feed and depth of cut (Krishnaraj et al., 2010). Thus, it is really
challenging to optimize the drilling the composite-metal stacks by balancing

FIGURE 7 (a) Chip removal problems (Brinksmeier and Janssen, 2002) and (b) long chip when drilling
TiA16V4 (Shyha et al., 2011). © Elsevier. Reproduced by permission of Elsevier. Permission to reuse
must be obtained from the rightsholder.
Machining Stacked Materials Made of CFRP and Alloys 495
Downloaded by [Johns Hopkins University] at 07:23 01 January 2015

FIGURE 8 Chip shapes and size of CFRP/Al composites: (a) feed versus drill diameter (b) speed versus
feed (Denkena et al., 2008). © Elsevier. Reproduced by permission of Elsevier. Permission to reuse must
be obtained from the rightsholder.
496 A. Pramanik and G. Littlefair

the machining parameters. It seems that design of cutting tool and setting
different machining condition at different layer is highly essential.

Surface and Geometric Condition


In case of CFRP-Al/Ti stack material the elastic modulus of different
materials causes different elastic deformations and therefore varying depth
of cut and tolerances along the depth of hole. In addition to the chip
removing through the hole, build-up cutting edges for metals and tool wear
affect the hole quality (Brinksmeier and Janssen, 2002). A large variation of
the surface roughness in CFRP is noted because of protruding fiber tips. The
hooking of the fibers to the stylus may cause additional errors. The surface
Downloaded by [Johns Hopkins University] at 07:23 01 January 2015

roughness of CFRP depends mainly on the stylus path with respect to fiber
direction as the directions of fibers may change from layer to layer. Thus
it depends highly on fiber and matrix properties, fiber orientation, cutting
condition and direction, and measurement direction (Zhang, 2009; Zitoune
et al., 2012). The increase of feed rate leads to a significant increase in the
value of the roughness which is independent of tool surface (Pramanik et al.,
2009). For example, the mean roughness increases from 0.43 µm to 0.94 µm
when the feed rate increases from 0.05 mm/rev to 0.15 mm/rev (Zitoune
et al., 2012).
A similar tendency is found when measuring the roughness in the CFRP
holes though the roughness of CFRP holes are significantly higher than
those of aluminum. As mentioned earlier, this is attributed to the heteroge-
neous nature of composite materials and also to the effects of carbon fiber
orientation relative to the direction of cut (Zhang, 2009). A conventional
drill tool (16 mm drill after 6 mm pre-drill) gives deviations in a range of
nearly 80 µm of the diameters on a 10 mm Al/10 mm CFRP/10 mm Ti
stack. The diameter values in the CFRP layer are distinctly lower because
of lower elastic modulus. The use of stepped tool geometry (step of 15.4
to 16 mm with diamond coating) improves the tolerances. In case of step
drill, mechanical loads are reduced by distribution on several cutting edges
and reduced depth of cut. These cause minor elastic deformation and lower
tolerances.
The chip transportation causes additional friction at the clearance angle
of the tool circumference and the hole surface. Thus, CFRP layer is eroded
by hot and sharp-edged metal chips and leads to bad surface qualities and
functional problems in subsequent assembly processes. The depth of erosion
can be reduced by using stepped tool where the secondary step of the tool
acts like a reaming tool and leads to improved chip transport (Brinksmeier
and Janssen, 2002). The quality of the holes increases due to application
helical milling of CFRP-Ti stacks (Denkena et al., 2008).
As noted earlier, fibers may hooking to the stylus and can cause addi-
tional errors during measuring surface roughness. Therefore, application
Machining Stacked Materials Made of CFRP and Alloys 497

of non-contact methods for measuring the surface roughness for composite


materials can be considered. There are several methods for non-contact mea-
surement of surface roughness, such as, non-contact optical profiler (NOP),
scanning tunnelling microscopy (STM) and atomic force microscopy (AFM)
(Poon and Bhushan, 1995). For STM, the surface to be measured needs to
be electrically conductive. Thus NOP and AFM might be used to measure
roughness of CFRP. However, it is common practice to use stylus profilmeter
for measuring surface roughness of CFRP with great care (Bagci and Işık,
2006; Ramulu et al., 1993). So far there is no report of using methods other
than stylus profilometer (contact method) for measuring surface roughness
of CFRP.
In CFRP/Al stacks, the circularity of CFRP increases with the increase of
Downloaded by [Johns Hopkins University] at 07:23 01 January 2015

feedrate. It was found that the hole diameter of CFRP shrinks due to relax-
ation of the (i) elastic stresses during machining (due to difference in elastics
stresses of epoxy and carbon fiber) and (ii) cutting edge radius of carbide
drill. This increases the friction between the drill and the surface of the hole
while drilling aluminum. Moreover, similar to titanium the aluminum chips
affect the quality of the hole of CFRP during evacuation (Zitoune et al.,
2010). In CFRP/Al stacks, better surface roughness of CFRP was found at
low feed rates (f = 0.05 mm/rev) for all spindle speeds and tool diameters
though longer chips (Fig. 7) at lower feed can damage the surface locally.
The surface roughness increases with increase in feed rate for all diameters.
The effect of the spindle speed on surface roughness is small. The surface
roughness and circularity of aluminum is better compared to that of CFRP
(Zitoune et al., 2010).
Different tools, such as, CVD diamond coated, uncoated carbide, C7
coated, produced almost similar quality holes during drilling 10 mm tita-
nium/10 mm CFRP/10 mm aluminum stacks (Shyha et al., 2011). Figure 9
shows picture of drilled surface of titanium and aluminum in 10 mm tita-
nium/10 mm CFRP/10 mm aluminum stack. Chips adhesion to the ma-
chined Ti surface and workpiece smearing noted during drilling of tita-
nium/CFRP/aluminum stacks. The quality of the surface is not similar at
exit and entry of the cutting tool. These require a finishing process follow-
ing drilling. Burr height is generally below 500 µm but it increases up to
1 mm, particularly in the Al section when spray mist is used. The supporting
layers of Al and Ti significantly reduce CFRP delamination. But the sharp
exit burrs of titanium cause minor damage around hole edges of the CFRP.
The picture of typical burrs in exit and entry of titanium and aluminum
layers are shown in Figure 10. The drilling process strain hardened up to
200 µm from the machined surfaces of the Ti and Al materials (Shyha et al.,
2011).
For drilling Gr/Bi–Ti stacks, speed affects the cylindricity and titanium
exit burr heights, and feed affects average diameter and surface roughness.
At higher speed drill bit wears suddenly and HSS-Co drills produce high
498 A. Pramanik and G. Littlefair
Downloaded by [Johns Hopkins University] at 07:23 01 January 2015

FIGURE 9 Typical machined surface in 10 mm titanium/10 mm CFRP/10 mm aluminum stacks (a)


Ti-middle of hole (b) Al-middle of hole (c) Ti-entry of hole and (d) Al-entry of hole (with tool wear)
(Shyha et al., 2011). © Elsevier. Reproduced by permission of Elsevier. Permission to reuse must be
obtained from the rightsholder.

FIGURE 10 Burr morphology at entry and exit of holes drilled in (a) Ti and (b) Al layers (Shyha et al.,
2011). © Elsevier. Reproduced by permission of Elsevier. Permission to reuse must be obtained from
the rightsholder.
Machining Stacked Materials Made of CFRP and Alloys 499

values for cylindricity, exit Ti burr height and surface roughness. The higher
surface roughness occurs due to fiber pull-out at 90◦ as well as −45◦ plies
against cutting direction in drilling (Kim and Ramulu, 2004).
The nano-composite coated drill remarkably reduces (more than 40%)
the surface roughness of the holes in aluminum and composite. Nano-coated
(2.32 µm thick CrAlN /a-Si3 N4 -Tripple Alwin) drills reduce the thrust force
by 47% and 20–25% during drilling aluminum and composite respectively
compare to uncoated drills. The surface roughness (Ra ≤ 3 µm) produced
by nano-composite coated (2.32 µm thick CrAlN /a-Si3 N4 -Tripple Alwin) tool
in CFRP is 30% less compared to that produced by an uncoated tool (Zitoune
et al., 2012).
At low spindle speeds built-up edge deteriorates drilling performance
Downloaded by [Johns Hopkins University] at 07:23 01 January 2015

significantly when the aluminum or Ti is stacked at the top of CFRP. In


addition, increase in cutting temperature or burr will be a source of damage
to the CFRP as the drill passes through the Ti or Aluminum first. The con-
tinuous chips passing through the CFRP deteriorate the quality of the hole
when the aluminum or Ti is stacked at the bottom of CFRP (Krishnaraj et al.,
2010; Vijayaraghavan and Dornfeld, 2005). In this case back counterboring
can occur in addition to crack and delamination as shown in Figure 11,
as the metal chips (i.e., aluminum or titanium) travel up the flutes, they
tend to erode the softer matrix material, causing eroded and oversized holes
(Parker, 2001).
Back counterboring can be reduced by (1) eliminating all gaps, (2) using
a drill geometry that produces small chips, (3) changing speeds and feeds,
(4) providing better clamp-up, (5) reaming the hole to final diameter after
drilling, or (6) by peck drilling (Parker, 2001). Peck drilling (Figure 12) is a
process in which the drill bit is periodically withdrawn to clear the chips from
the flutes. Peck drilling is used almost exclusively when drilling composite-
to-titanium stack-ups, due to the back counterboring potential of the hard
titanium chips.

FIGURE 11 Back counterboring.


500 A. Pramanik and G. Littlefair

FIGURE 12 Peck Drilling.

The process also greatly reduces the heat build-up that can rapidly occur
when drilling titanium (Ramulu et al., 2001; Zhang et al., 2008). A typical
peck drilling cycle for a 5.8 mm diameter hole through carbon/epoxy-to-
Downloaded by [Johns Hopkins University] at 07:23 01 January 2015

titanium would be a speed of 550 rpm with a feed rate of 0.05–0.1 mm/rev
and 30–60 pecks per inch of thickness (Bolt and Chanani, 1987). However,
peck drilling is a more commonly performed for deep hole drilling opera-
tions (Smid, 2003) where depth of the hole is greater than the three times
of the tool diameter in a certain material (Hurst, 2006; Nomani et al., 2013).

Tools and Tool Wear


Generally different cutting tool geometry and cutting conditions are
selected for CFRP, aluminum and titanium alloys. Maximum tool life is
achieved when CFRP is machined at a slower speed and faster feed. Faster
feed corresponds to less cutting distance travelled for the same depth of drill
as the cutting edge takes its spiral path down through the hole. This leads to
less tool wear. However, lower feed is usually used to keep delamination and
breakout low for CFRP. Lower speed and feed are chosen for Ti compared
to CFRP because low thermal conductivity, high hardness and complex chip
formation process of Ti alloys (Pramanik, 2013, Pramanik et al., 2013a).
Each material can be drilled in its optimum machining condition by
changing the speed and feed layer-by-layer according to the type of materials.
This is common practice in machining stack materials, which leads to a
longer tool life, better surface finish and more accurate dimensions (Park
et al., 2011). However, experimental investigations shows that the cutting
conditions applied to metals can also be applied to composite materials
(Brinksmeier and Janssen, 2002). Thus conventional twist drills can be used
to drill CFRP if those tools can resist the abrasion of the carbon fibers (Konig
et al., 1985; Park et al., 1995; Rahman et al., 1999).
Theoretically stack materials require different tools that fit the at-
tributes of a composite and aluminum or titanium. The friction contact
at the rake and the flank faces when machining titanium alloys (in alu-
minum/CFRP/titanium stack) leads to unfavorable thermal and mechani-
cal loads on the cutting edges which increase tool wear. Adapted step drills
improve diameter tolerances, surface quality, and reduced tool wear for this
Machining Stacked Materials Made of CFRP and Alloys 501

kind of materials (Brinksmeier and Janssen, 2002). Sharp and high hot hard-
ness tool materials are required for machining stack materials. Tool with low
hardness (high-speed cobalt: HSS-Co) wears rapidly when drilling Gr/Bi–Ti
stacks. Flank and crater wear are developed in these drills.
Severe tool wear is noted on the helical cutting edges of the low hardness
drills as well. For the harder tools, such as carbide drills, flank wear is minor
for drilling same number of holes. It this case least amount of wear occurs
at the tip of the point angle (Kim and Ramulu, 2004). The use of step
drills and the use of minimum quantity lubrication improve the process
characteristics when drilling multi-layer materials. Application of minimum
quantity lubrication reduces the build-up edges at the cutting edge and flank.
These are responsible for increased tolerances and bad surface qualities
Downloaded by [Johns Hopkins University] at 07:23 01 January 2015

(Brinksmeier and Janssen, 2002; Kelly and Cotterell, 2002; Marques et al.,
2009).
In drilling 4.2 mm CFRP/ 3 mm Al stacks the wear of a plain carbide
(K20) 6-mm drill increases steadily (up to 30 holes) then it becomes stable
(30–60 holes) and after that the wear starts to increase again (Zitoune et al.,
2010). The increase in tool wear results in a change in the material removal
mechanism. The wear of drill tool reduces the cutting capability of tool and
results in buckled fibers (Kim and Ramulu, 2007). The buckled fibers are
forced out from matrix material uncut and further contribute to tool wear
and surface roughness (Zitoune et al., 2010). The wear tests have shown
that, from 30 to 60 holes stable cutting force is found, this region could be
attributed to normal wear region. This range is ideal for drilling of CFRP/Al
stack (Zitoune et al., 2010).
The optimum drilling conditions for desired hole quality and cost for
drilling of Gr/Bi–Ti are found to be a combination of low speed and low
feed for carbide drills and, low speed and high feed for HSS-Co drills (Kim
et al., 2005). During drilling titanium alloys, the friction contact at the rake
and the flank faces generate unfavorable thermal and mechanical loads
which increase tool wear. Because of this, the tool life during machining
aluminum/CFRP/titanium is shorter. The tool geometry and cutting pa-
rameters control the wear behavior of tools. The reduced speed significantly
reduces the adhesion of titanium chips on the cutting edges (Brinksmeier
and Janssen, 2002).
In case of carbon fiber-reinforced plastics (CFRP) stacked on top of
titanium (Ti) the PCD tool performed much better than WC tool. The rate
of tool wear is much higher in a WC tool. More adhesion of Ti is observed
on the WC drills than on the PCD drills. In case of a WC drill no significant
micro-chipping occurs on the cutting edges but on the cutting edges near the
drill margin at the higher spindle speed. The wear is generally smooth and
uniform along the cutting edges of PCD drills. Though the edge wear and
flank wears are similar at the lower spindle speed, the flank wear dominates
at the higher spindle speed in drilling Ti layer due to the higher cutting
502 A. Pramanik and G. Littlefair

TABLE 1 Recommended Drilling Condition and Expected Tool Life for 6.35 mm
Drill (modified from Garrick [2007])

Stack Material CFRP/CFRP CFRP/Al CFRP/Ti

Tool Material PCD PCD WC


Layer 1
Speed (rpm) 18000 18000 4500
Feed (mm/min) 4572 4572 457
Layer 2
Speed (rpm) 18000 10000 700
Feed (mm/min) 4572 1016 51
Life (holes drilled) 1500 1000 45
Downloaded by [Johns Hopkins University] at 07:23 01 January 2015

temperature. The cutting edges near the drill margin of PCD tool develops
micro-chipping. This tool shows less Ti adhesion and higher wear resistance
in drilling of CFRP. In CFRP/Ti stack drilling, the tool wear mechanisms for
CFRP and Ti are different. Hard carbon fibers in CFRP abrades the cutting
edge and causes edge wear while Ti wears flank land in addition to carbide
grain pull-outs when removing attached Ti (Park et al., 2010).
Boron aluminum magnesium coating reduces the titanium adhesion
and tool wear at high cutting speed compared to uncoated tools during
drilling 7.54 mm CFRP/ 6.73 mm titanium stack. This coating wears out
after drilling reasonable numbers of (approximately 20) holes and, the ad-
hesion and wear are increased with further drilling of holes mainly due to
the coating delamination. The delamination may be delayed by improving
the surface quality of the coating. The dominant tool wear mechanism is
edge wear by hard carbon fibers in CFRP (Park et al., 2012). Increased tool
life and improved hole quality can also be achieved by PCD veined drill
during drilling CFRP/Ti stacks (Garrick, 2007). Since temperature is a large
factor in machining titanium, best results could be achieved when using
coolant. Table 1 gives an idea about tool life and cutting conditions of PCD
and WC for different combination of stack materials. The recommended
conditions for drilling a 7.7-mm CFRP/6.15-mm titanium stack using a 6.35-
mm PCD veined drill are given in Table 2.

TABLE 2 Operating Parameters of 6.35 mm Diameter PCD Veined Drill in


CFRP/Ti Stack (modified from Garrick [2007])

Stack material 7.7 mm CFRP/6.15 mm titanium


CFRP
RPM 6000
Feed 1520 mm/min
Ti
RPM 700
Feed 30 mm/min
Peck Depth 1.016 mm
Machining Stacked Materials Made of CFRP and Alloys 503

When the cutting tool material is softer, e.g., carbide (Grade K20 un-
coated), the tool wear is mainly caused by CFRP, which impact all the cut-
ting edges. The build-up layer on the tool during drilling metallic layer is not
strong enough to protect the cutting edge during drilling CFRP. In addition,
the build-up layer may start chemical wear of the tool. The wear generated
by metallic (Ti6Al4V ) layer is mostly confined to cutting edge chipping on
the drill corner (Poutord et al., 2013). For the harder cutting tools, e.g.,
PCD, chemical reaction takes place if there is a Ti layer though PCD can
handle the abrasion of hard carbon fibers in CFRP. The expected tool life of
a 6 mm PCD drill is much more than 50 holes in 3.1 mm CFRP/4.1 mm Ti
(Gao and Zhang, 2011). When drilling carbon/epoxy-to-aluminum, a speed
of 2000–3000 rpm with a feed rate of 0.025–0.05 mm/rev might be used,
Downloaded by [Johns Hopkins University] at 07:23 01 January 2015

while a stack-up of carbon/epoxy-to-titanium would require a slower speed


(e.g., 300–400 rpm) and a higher feed rate (0.1–0.127 mm/rev). Titanium
(Ti-6Al-4V ) is also very sensitive to heat build-up (hence the lower speed)
and tends to rapidly work harden if light cuts are used (hence the higher
feed rate) (Campbell, 2006).

Improved Drilling Techniques


Brinksmeier et al. (2011) performed orbital drilling to drill 16 mm
hole by 11 mm diameter cemented carbide tool on CFRP/aluminum/
CFRP/titanium stacks. In orbital drilling, the tool spins on its own axis as
well as rotates eccentrically around a principal axis. In this method the drill
tool is feed rather than pushed through the material (Margolis et al., 1999).
The thermal and mechanical loads generated in this case were significantly
lower compare to that of conventional drilling under same cutting speed and
material removal rate. However, similar to conventional drilling, the average
temperature and feed force increase from drilling the aluminum, over CFRP,
to the titanium layer in orbital drilling. The orbital drilling reduces the aver-
age temperature and load in CFRP and titanium by half (Brinksmeier et al.,
2011).
In the case of conventional drilling, the friction between the tool clear-
ance face and the borehole surface is much higher than that of orbital
drilling process because of the continuous contact. In orbital drilling the
cutting edges go into contact with the borehole intermittently.
The tool loses the contact to the borehole surface after certain engage-
ment angle (Brinksmeier et al., 2011). The effects related to chip evacuation
damage due to chip flow and delamination are minimized in orbital drilling.
In orbital drilling less tool inventory is needed since one tool can be used for
different diameter holes (Margolis and Piliavin, 1999; Ni, 2007). Denkena
et al. (2008) also used orbital drilling to drill a 10-mm diameter hole on
TiAl6V4/CIRP stack by a 8-mm diameter TiAlN coated solid carbide end
mill tool. It was noted that bore hole diameter in the CFRP layer is always
504 A. Pramanik and G. Littlefair

higher than in the titanium layer which is similar to conventional drilling.


An increase in the axial feed per tooth increases feed force. These reduce
the diameters of holes in the CFRP as well as titanium layer. Similarly, the
diameters of holes in both layers increases with the increase of the tangential
feed per tooth (Denkena et al., 2008).
Similar to orbital drilling, the force vibratory drilling might be useful for
making holes in metal/composite stack materials. Though there is no study
on application of this method on metal/composite stack materials, it is found
very effective in drilling of CFRP. Vibratory drilling is a branch of vibration
assisted machining, has some basic difference with conventional drilling. It
is a pulsed intermittent cutting process where the thrust force is reduced by
vibratory effect in metals and the quality of the holes is improved (Arul et al.,
Downloaded by [Johns Hopkins University] at 07:23 01 January 2015

2006). Thrust force is considered as the cause of delamination in polymeric


composites. Thus the vibratory drilling has the capacity to limit the thrust
force below which the delamination can be minimized and/or eliminated.
Based on the above discussion it can be said that vibratory drilling is a
promising technique for making holes in metal/composite stack materials
as this method improves the drilling of metals as well as CFRP (Qixin et al.,
1994; Takeyama and Kato, 1991).

Application of Coolant
The machinability of different materials can be improved by applying
different types of coolants. Coolants introduce lubrication effect as well as
cool the machining zone. The effectiveness of cooling method depends on
its ability to penetrate the coolant into the machining zone. Properly directed
coolants at high pressure show increased effectiveness. The performance of
coolants during drilling of stacked materials may not be similar to that of
single layered materials. Shyha et al. (2011) tested performance of different
cooling methods on Ti/CFRP/Al stacks at speed 20–40 m/min and feed
0.05 mm/rev. It was noted that the flood coolant produced undersized
holes in all the layers where maximum errors were 14, 20 and 15 µm in Ti,
CFRP and Al layers, respectively.
On the other hand, oversized holes were generated when spray mist
was used as coolant. In this case the maximum errors were 6, 120 and 28
µm in Ti, CFRP and Al layers respectively. In addition, better roundness and
cylindricity were achieved in flood cooling. The maximum deviations in hole
roundness were up to 78, 39 and 53 µm in Ti, CFRP and Al layers, and the
cylindricity error over the entire stack varied between 23 and 120 µm when
flood cooling was applied. This increased up to an average of 170 µm for
the spray mist cooling. Thus discrepancies in dimensions of holes for three
different materials are generated which are due to the different mechanical
properties of the stack elements and in particular their elastic modulus and
thermal properties (Brinksmeier and Janssen, 2002).
Machining Stacked Materials Made of CFRP and Alloys 505

From the above information it is clear that the flood cooling performs
much better than that of spray mist. This indicates that the mist spray is
unable to cool the drilling process effectively which leads to higher tempera-
ture in machining zone. Higher temperature reduces elasticity of materials,
increases tool diameter and affects larger amount of materials in the radial
direction of hole. Therefore, it is likely that the hole diameters are larger
than expected. Though, flood cooling performs better than mist cooling, the
chip clogging within the drill flutes causes premature tool failure frequently
in these cases (Soo et al., 2010). This indicates that the application of high-
pressure coolant (∼70 bars, through spindle) may facilitate chip evacuation,
increased tool life and better cooling effect. However, too much coolant
applied during machining goes to environment which is very harmful.
Downloaded by [Johns Hopkins University] at 07:23 01 January 2015

Therefore, measures have been taken to reduce the amount of coolant


without compromising the effective cooling of machining zone for minimiz-
ing the adverse effects on environment. This is known as minimum quantity
lubrication (MQL). The application of MQL enhances the tolerance and
surface quality. Brinksmeier and Janssen (2002) achieved diameter toler-
ances of 17 µm along the entire hole by using MQL with fatty alcohol.
The use of MQL facilitates chip removal by reducing the friction and ad-
hesion of aluminum debris on the tool grooves (Brinksmeier and Janssen,
2002).
Having said this, it is very difficult to compare different cooling methods
side by side for machining composite-metal stacked materials as the different
researches used different machining conditions as well as necessary infor-
mation is missing in some cases. However, from the preceding discussion, it
can be concluded that application of MQL at high pressure through cutting
tool is the best option for drilling metal-composite stacked materials.

CONCLUSION
The investigation in drilling composite-metal stacks is still in the primary
stage. The composite-metal can be stacked in verities of combinations. Thus,
results obtained by drilling must have huge variation. However, there are
several trends of drilling outcome of these stacks materials, which can be
summarized in the following points.

(a) The presence of metallic layers reduces the exit and entry delami-
nation in CFRP. However, heat generation, built-up edge and chip
evacuation during drilling metallic layers deteriorate the surface of
CFRP.
(b) Order of materials has significant influence on material surface finish
and tool wear. If the metallic layer is on the top, then chip evacuation
506 A. Pramanik and G. Littlefair

does not affect the CFRP surface but heat and built-up edge affect the
CFRP significantly.
(c) Coating and step drill significantly improves the tool performance,
surface finish and geometric tolerance. PCD tools also generate better
surface finish and geometric tolerance. The selection of the tool
material depends on the type of material layers used.
(d) Chipping, adhesion, abrasion and attrition are main tool wear mech-
anism during machining of CFRP/metal stacks. In fact, all the tool
wear mechanisms involved in drilling CFRP, Al and Ti are responsible
for drilling stack materials.
(e) High pressure or internal cooling improves tool life, surface finish,
Downloaded by [Johns Hopkins University] at 07:23 01 January 2015

chip evacuation and tolerance.


(f) The defect generation during drilling of these materials is not sub-
stantially eliminated even after taking different types of preventive
measures.

REFERENCES
Abrão, A.M.; Faria, P.E.; Rubio, J.C.; Reis, P.; Davim, J.P. (2007) Drilling of fiber reinforced plastics: A
review. Journal of Materials Processing Technology, 186: 1–7.
Arul, S.; Vijayaraghavan, L.; Malhotra, S.K.; Krishnamurthy, R. (2006). The effect of vibratory drilling on
hole quality in polymeric composites. International Journal of Machine Tools and Manufacture, 46(3):
252–259.
Askeland, D.R. (1989) Mechanical testing and properties, In The Science and Engineering of Materials, 2nd
edition. Chapman and Hall, London, 145–181.
Astrom, B.T. (1997) Manufacturing of Polymer Composites. Chapman & Hall, New York.
Bagci, E.; Işık, B. (2006). Investigation of surface roughness in turning unidirectional GFRP composites by
using RS methodology and ANN. International Journal of Advanced Manufacturing Technology, 31(1–2):
10–17.
Bolt, J.A.; Chanani, J.P. (1987) Solid tool machining and drilling. In Engineered Materials Handbook Com-
posites, vol. 1, T.J. Reinhart (Ed.), University of Michigan, Ann Arbor, MI, 667–672.
Brinksmeier, E.; Fangmann, S.; Rentsch, R. (2011) Drilling of composites and resulting surface integrity.
CIRP Annals-Manufacturing Technology, 60(1): 57–60.
Brinksmeier, E.; Janssen, R. (2002) Drilling of multi-layer composite materials consisting of carbon fiber
reinforced plastics (CFRP), titanium and aluminum alloys. CIRP Annals-Manufacturing Technology,
51(1): 87–90.
Campbell, F.C. (2004) Assembly. In Manufacturing Processes for Advanced Composites. Elsevier Ltd., Oxford,
UK, 440–469.
Campbell, F.C. (2006) Structural assembly. In Manufacturing Technology for Aerospace Structural Materials,
1st ed. Elsevier Ltd., Oxford, UK.
De Garmo, E.P.; Black, J.T.; Kohser, R.A. (2011) DeGarmo’s Materials and Processes in Manufacturing. John
Wiley & Sons, New York.
Denkena, B.; Boehnke, D.; Dege, J.H. (2008) Helical milling of CFRP–titanium layer compounds. CIRP
Journal of Manufacturing Science and Technology, 1(2): 64–69.
Dharan, C.K.H.; Won, M.S. (2006) Machining parameters for an intelligent machining system for com-
posite laminates. International Journal of Machine Tools and Manufacture, 40: 415–426.
DiPaolo, G.; Kapoor, S.G.; DeVor, R.E. (1996) An experimental investigation of the crack growth phe-
nomenon for drilling of fiber-reinforced composite materials. Transactions of the ASME Journal of
Enginreeing for Industry, 118: 104–110.
Machining Stacked Materials Made of CFRP and Alloys 507

Enemuoh, E.; El-Gizawy, A.S.; Okafor, A.C. (2001) An approach for development of damage-free drilling
of carbon fiber reinforced thermosets. International Journal of Machine Tools and Manufacture, 41(12):
1795–1814.
Faraz, A.; Biermann, D.; Weinert, K. (2009) Cutting edge rounding: An innovative tool wear criterion
in drilling CFRP composite laminates. International Journal of Machine Tools and Manufacture, 49(15):
1185–1196.
Gao, H.; Zhang, X. (2011) Drilling of composite and titanium stacks with alternating machining param-
eters by PCD drill. International Journal of Machining and Machinability of Materials, 10(4): 280–292.
Garrick, R. (2007) Drilling advanced aircraft structures with PCD (poly-crystalline diamond) drills. SAE
Technical Paper, DOI 10.4271/2007-01-3893.
Hocheng, H.; Tsao, C.C. (2003) Comprehensive analysis of delamination in drilling of composite mate-
rials with various drill bits. Journal of Materials Processing Technology, 140: 335–339.
Hurst, B. (2006) The Journeyman’s Guide to CNC Machines. Lulu.com, Raleigh, North Carolina.
Kelly, J.F.; Cotterell, M.G. (2002) Minimal lubrication machining of aluminum alloys. Journal of Material
Processing Technology, 120: 327–334.
Downloaded by [Johns Hopkins University] at 07:23 01 January 2015

Kim, D.; Ramulu, M. (2004) Drilling process optimization for graphite/bismaleimide titanium alloy
stacks. Composite Structures, 63(1): 101–114.
Kim, D.; Ramulu, M. (2005) Machinability of titanium/graphite hybrid composites in drilling. Transactions
of the NAMRI/SME, 33: 445–452.
Kim, D.; Ramulu, M. (2007) Study on the drilling of titanium/graphite hybrid composites. ASME Journal
of Engineering Materials and Technology, 129: 390–396.
Konig, W.; Wulf, C.; Grass, P.; Willerschied, H. (1985) Machining of fiber reinforced plastics. CIRP
Annals-Manufacturing Technology, 34(2): 537–548.
Krishnaraj, V.; Zitoune, R.; Collombet, F. (2010) Comprehensive review on drilling of multi material
stacks. Journal of Machining and Forming Technologies, 2(3–4): 1–32.
Langella, A.; Nele, L.; Maio, A. (2005) A torque and thrust prediction model for drilling of composite
materials. Composites: Part A, 36: 83–93.
Margolis, B.; Piliavin, J. (1999). ‘Stacking’ in major league baseball: A multivariate analysis. Sociology of
Sport Journal, 16(1); 16–34.
Marques, A.T.; Luı́s, M.D.; António, M.G.; João, F.S.; João, M.R.S.T. (2009) Delamination analysis of
carbon fiber reinforced laminates: evaluation of a special step drill. Composites Science and Technology,
69(14): 2376–2382.
Mathew, J.; Ramakrishnan, N.; Naik, N.K. (1999) Investigations into the effect of geometry of a trepan-
ning tool on thrust and torque during drilling of GFRP composites. Journal of Materials Processing
Technology, 91(1): 1–11.
Ni, W. (2007). Orbital drilling of aerospace materials. Heat Treatment, 2012, 3–12.
Nomani, J.; Pramanik, A.; Hilditch, T.; Littlefair, G. (2013) Machinability study of first generation duplex
(2205), second generation duplex (2507) and austenite stainless steel during drilling process. Wear,
304: 20–28.
Paleen, M.J.; Kilwin, J.J. (2001) Hole drilling in polymer-matrix composites, in ASM Handbook of Composites.
ASM International, Material Park, OH, 21: 646–650.
Park, K.H.; Beal, A.; Kim, D.D.W.; Kwon, P.; Lantrip, J. (2011) Tool wear in drilling of composite/titanium
stacks using carbide and polycrystalline diamond tools. Wear, 271(11): 2826–2835.
Park, K.Y.; Choi, J.H.; Lee, D.G. (1995) Delamination-free and high efficiency drilling of carbon fiber
reinforced plastics. Journal of Composite Materials, 29(15): 1988–2002.
Park, K.H.; Kwon, P.Y.; Castro, G.; Kim, D.; Lantrip, J. (2010) Preliminary study on tool wear in drilling
of composite/titanium stacks with carbide and PCD tools. Transaction of NAMRII/SME, 38: 283–290.
Park, K.H.; Kwon, P.Y.; Kim, D. (2012) Wear characteristic on BAM coated carbide tool in drilling
of composite/titanium stack. International Journal of Precision Engineering and Manufacturing, 13(7):
1073–1076.
Parker, R.T. (2001) Mechanical fastener selection. In ASM Handbook Composites. ASM International,
Material Park, OH, 21: 651–658.
Poon, C.Y.; Bhushan, B. (1995). Comparison of surface roughness measurements by stylus profiler, AFM
and non-contact optical profiler. Wear, 190(1): 76–88.
Poutord, A.; Rossi, F.; Poulachon, G.; M’Saoubi, R.; Abrivard, G. (2013) Local approach of wear in drilling
Ti6Al4V/CFRP for stack modelling. Procedia CIRP, 8: 315–320.
508 A. Pramanik and G. Littlefair

Pramanik, A. (2013) Problems and solutions in machining of titanium alloys. International Journal of
Advanced Manufacturing Technology, 70(5–8): 919–928.
Pramanik, A.; Islam, M.N.; Basak, A.; Littlefair, G. (2013) Machining and tool wear mechanisms during
machining titanium alloys. Advanced Materials Research, 651: 338–343.
Pramanik, A.; Neo, K.S.; Rahman, M.; Li, X.P.; Sawa, M.; Maeda, Y. (2009) Ultraprecision turning of
electroless nickel: effects of crystal orientation and origin of diamond tools. The International Journal
of Advanced Manufacturing Technology, 43(7–8): 681–689.
Pramanik, A.; Zhang, L.C.; Arsecularatne, J.A. (2007) Micro-indentation of metal matrix composites–an
FEM analysis. Key Engineering Materials, 340: 341.
Pramanik, A.; Zhang, L.C.; Arsecularatne, J.A. (2008) Deformation mechanisms of MMCs under inden-
tation. Composites Science and Technology, 68(6): 1304–1312.
Qixin, Z.; Shiyu, S.; Jianwei, L.; Youbin, F.; Chengzian, M.; Xifu, T. (1994). Study on ultrasonic vibra-
tion drilling in carbon fiber reinforced polymers. Chinese Journal of Mechanical Engineering (English
Edition)(China), 7(1): 72–77.
Rahman, M.; Ramakrishna, S.; Prakash, J.R.S. (1999) Machinability study of carbon fiber reinforced
Downloaded by [Johns Hopkins University] at 07:23 01 January 2015

composite. Journal of Materials Processing Technology, 89–90: 292–297.


Ramulu, M.; Branson, T.; Kim, D. (2001) A study on the drilling of composite and titanium stacks.
Composite Structures, 54: 67–77.
Ramulu, M.; Wern, C.W.; Garbini, J.L. (1993). Effect of fiber direction on surface roughness measure-
ments of machined graphite/epoxy composite. Composites Manufacturing, 4(1): 39–51.
Shyha, I.S.; Soo, S.L.; Aspinwall, D.K.; Bradley, S.; Perry, R.; Harden, P.; Dawson, S. (2011) Hole quality
assessment following drilling of metallic-composite stacks. International Journal of Machine Tools and
Manufacture, 51(7): 569–578.
Smid, P. (2003) CNC Programming Handbook: A Comprehensive Guide to Practical CNC Programming. Industrial
Press Inc., Norwalk, CT.
Soo, S.L.; Aspinwall, D.K.; Shyha, I.; Bradley, S.; Dawson, S.; Pretorius, C.J. (2010). Drilling of tita-
nium/CFRP/aluminum stacks. Key Engineering Materials, 447: 624–633.
Takeyama, H.; Kato, S. (1991). Burrless drilling by means of ultrasonic vibration. CIRP Annals-
Manufacturing Technology, 40(1): 83–86.
Tsao, C.C.; Hocheng, H. (2005) Effect of eccentricity of twist drill and candle stick drill on delamination in
drilling composite materials. International Journal of Machine Tools and Manufacture, 45(2): 125–130.
Velayudham, A.; Krishnamurthy, R.; Soundarapandian, T. (2005) Evaluation of drilling characteristics
of high volume fraction fiberglass reinforced polymeric composite. International Journal of Machine
Tools and Manufacture, 45(4): 399–406.
Vijayaraghavan, A.; Dornfeld, D.A. (2005) Challenges in Modeling Machining of Multilayer Materials. Lab-
oratory for Manufacturing and Sustainability, University of California, Berkeley. Retrieved from:
http://escholarship.org/uc/item/60k6×64r. Last accessed 5 May 2013.
Won, M.S.; Dharan, C.K.H. (2002) Chisel edge and pilot hole effects in drilling composite laminates.
Journal of Manufacturing Science and Engineering, 124(2): 242–247.
Zhang, L.C. (2009) Cutting composites: A discussion on mechanics modelling. Journal of Materials Pro-
cessing Technology, 209(9): 4548–4552.
Zhang, P.F.; Churi, N.J.; Pei, Z.J.; Treadwell, C. (2008) Mechanical drilling processes for titanium alloys:
a literature review. Machining Science and Technology, 12(4): 417–444.
Zitoune, R.; Krishnaraj, V.; Almabouacif, B.S.; Collombet, F.; Sima, M.; Jolin, A. (2012) Influence of
machining parameters and new nano-coated tool on drilling performance of CFRP/Aluminum
sandwich. Composites: Part B, 43: 1480–1488.
Zitoune, R.; Krishnaraj, V.; Collombet F. (2010) Study of drilling of composite material and aluminum
stack. Composite Structures, 92: 1246–1255.

You might also like