Вы находитесь на странице: 1из 27

Asymmetric Single Point Incremental Forming of Sheet Metal

J. Jeswiet1 (1), F. Micari2 (1), G. Hirt3, A. Bramley4 (1), J. Duflou5 (2), J. Allwood6
1

Queens University, Kingston, ON, Canada


2
University of Palermo, Palermo, Italy
3
University of Aachen, Germany
4
University of Bath, United Kingdom
5
Katholieke Universiteit Leuven, Belgium
6
University of Cambridge, Cambridge, England
Abstract
The use of computers in manufacturing has enabled the development of several new sheet metal forming
processes, which are based upon older technologies. This paper describes modifications that have been
made to traditional forming methods such as conventional spinning and shear forming, forming processes in
which deformation is localized. Recent advances have enabled this localized deformation to be accurately
controlled and studied. Current developments have been focused on forming asymmetric parts using CNC
technology, without the need for costly dies. Asymmetric Incremental Sheet Forming has the potential to
revolutionize sheet metal forming, making it accessible to all levels of manufacturing. This paper describes
the genesis and current state-of-the-art of Asymmetric Incremental Sheet Forming.

Keywords:
Forming, Rapid Prototyping, Sheet Metal

1 INTRODUCTION
New methods of forming sheet metal are now at a stage
where it is possible to make either custom manufactured
parts or to manufacture small batch production quantities,
with very short turn around times from design to
manufacture. Schmoeckel [1] predicted in 1991 that with
the increase in automation metal forming equipment would
become more flexible. That has happened in this case.
The idea of incrementally forming sheet metal with a single
point tool, called dieless forming, was patented by Leszak
[2] well before it was technically feasible. There have been
many studies, which have lead to the present situation [3
8]. Today, there are new processes whereby sheet metal
is plastically deformed at a local point enabling truly
flexible production of complex sheet metal parts. This can
be done in either small batch lots with short lead times, or
in production of usable rapid prototypes within one day.
The new processes are attractive because manufacturing
sheet metal can be accomplished by any facility having a
three-axis CNC mill.
Inspiration for the emerging processes is usually found in
traditional forming methods. These conventional processes
are typically constrained as far as achievable part
geometry is concerned and require dedicated tooling and
dies. CNC hardware and software have reached a mature
state of development enabling the development of new
sheet metal forming processes. The new forming methods
give the possibility to create flexible forming facilities,
without dies, capable of producing complex shaped
surfaces, while applying generic tooling. The ultimate goal
is dieless forming.
This paper will show how a product progresses from the
design stage, to developing toolpaths for metal forming, to
the final production process.
Spinning is the forerunner of asymmetric incremental sheet
metal forming. Hagan and Jeswiet [9] sketched a state-ofthe-art for a number of the newly emerging sheet metal
forming processes, all having a genesis in spinning, and all
having a potential for rapid prototyping applications with
sheet metal.

A REVIEW OF SYMMETRIC INCREMENTAL SHEET


METAL FORMING PROCESSES
An in-depth discussion of spinning can be found in Brown
[10, 11]. It can be broken down into the following three
subgroups: 1) conventional spinning; 2) shear forming; 3)
flow forming. Conventional spinning and shear forming will
be discussed briefly.
2.1 Spinning and Shear Forming.
Conventional spinning is performed through a series of
sweeping strokes with a forming tool [12]. Figure 1
illustrates the movement of material from the original blank
to the final conical shape using several passes. In shear
forming, a roller tool is used and motion is programmed to
follow the profile of the part surface [13]. In the ideal shear
forming process no radial displacement of material occurs
[14, 15]. Three types of computer-aided control are used in
forming the sheet metal. CNC and numerical control (NC)
systems and control type, programmable numerical control
(PNC), which uses playback technology [16].
The final thickness of a shear formed part can be
calculated using the sine law [13], where ti is the initial
thickness, tf is the final thickness, and is the spinning
angle; see Figure 2. An element is shown for the wall and
flange of a shear formed cone. There is no displacement
of material in the flange region thus the wall deforms under
pure shear and is defined as the draw angle.
The plastic deformation forming limits in spinning can be
determined by calculating spinnability [18, 19]. Kegg [18]
defined spinnability as the maximum reduction of
thickness, t, that a material, with initial thickness ti can
achieve; see Figure 2.
2.2 Symmetric Incremental Forming
Spinning without a mandrel showed dies are unnecessary
with a single point forming tool. The work by Kitazawa [22,
23, 24, 25] showed controlled deformation is possible with
a single point tool. Also, the techniques developed by
Kitazawa [23] were a useful step toward multiple pass
forming used in asymmetric incremental forming.

Formed cone

Formed Cone (one pass)

Rotating
Mandrel

Rotating
Mandrel

Initial Blank

Initial Blank

Shear forming

Conventional
spinning

tf

ti
ti

sine law: t f = t i sin

before deformation

Figure 1. Conventional spinning and shear forming of a


cone [9].

ellipsoid
mandrel

spinnability:
t t
%R = i
ti

Figure 2. The spinning sine law [14] and the shear


spinnability test designed by Kegg [18].
3 ASYMMETRIC INCREMENTAL SHEET FORMING
Several new metal forming techniques have been
developed in the last few years due to advances in: 1)
computer controlled machining; 2) symmetric single point
forming (spinning); and 3) the development of toolpath
postprocessors in CAD software packages. One significant
outcome of this technology is the ability to form
asymmetric shapes at low cost, without expensive dies.
The asymmetric sheet metal incremental forming
techniques discussed here can be divided into different
categories. First is a method initially developed by Powell
and Andrew [26], which was subsequently called the
backward bulge method by Matsubara [27]. Bambach et al.
are also active with this application [28]. Both symmetric
and asymmetric shapes can be created using this process.
See Figures 3 (c) and (d). The next class includes work by
Jeswiet [29], Kim [30], Leach [31], and Felici [32], all of
whom have studied the application of incremental CNC
forming technology to asymmetric shapes. See Figure 3
(a). In this process, the blank remains stationary and
forming occurs during CNC control of the tool in a CNC
mill. The foregoing are discussed in the following sections.
3.1 Definition, making asymmetric shapes from sheet
metal without dies
Asymmetric incremental sheet forming (AISF) can be
interpreted in different ways. Hence, a definition with
Figures is included here, so that the process described
cannot be confused with other incremental forming
processes. Asymmetric Incremental Sheet Forming (AISF)
is a process which:
is a sheet metal forming process,
has a solid, small-sized forming tool,
does not have large, dedicated dies,
has a forming tool which is in continuous contact with
sheet metal,

has a tool that moves under control, in three


dimensional space
can produce asymmetric sheet metal shapes
It is the last characteristic that separates symmetric
spinning from AISF. AISF processes are purely a
consequence of the introduction of CNC mills and CAD
software with toolpath postprocessors. The idea was first
introduced in a patent in 1967 [2], but the foregoing tools
were not available at that time. The term dieless, as
applied to this process, was first used in that patent.
Asymmetric sheet metal parts can be made with either 1) a
machine specifically designed for the process, or 2) a
three-axis CNC mill, which most manufacturing facilities
possess. Although machines have been designed
specifically for this process AISF of sheet can be carried
out by anyone having access to a three-axis CNC mill and
off-the-shelf software, which generates machine toolpaths.
Figure 3 shows the different configurations that are
included in the group of asymmetric incremental sheet
forming (ASIF) techniques. It can be seen that three of the
processes meet the above criteria. The fourth, Figure 3(d),
uses a full die and does not meet the dieless criteria.
However, it is related directly to the other processes and it
is being used successfully to make rapid prototype shapes.
Hence it is included in the list of dieless forming
techniques.
The configurations shown in Figures 3(a) and (b) have a
stationary sheet metal blankholder. The configurations
shown in Figures 3(c) and (d) have a sheet blankholder
that moves along a vertical axis, as the forming tool
deforms the sheet, hence the vertical motion arrows
shown. The configuration shown in Figures 3(c) and (d)
are usually associated with machines dedicated to the reference 1
incremental forming process, and those in Figures 3(a)
and (b) are usually found in a CNC mill application.
AISF includes two specific types of incremental forming:
SPIF, single point incremental forming and TPIF, two point
incremental forming.
stationary blank holder
sheet

forming tool

faceplate

(a) Single Point


Incremental Forming;
SPIF
partial die

counter tool
(b) Incremental Forming
with counter tools

blankholder motion

(c) Two point


incremental forming
(partial die); TPIF

full die

(d) Two point


incremental forming
(full die); TPIF

Figure 3: Process principles of AISF. Four variations are


shown [28].
3.2 Common Types of Asymmetric Incremental Sheet
Forming
Asymmetric incremental sheet forming has four basic
elements: 1) a sheet metal blank, 2) a blankholder, 3) a
single point forming tool, 4) CNC motion. These basic

elements are illustrated in Figure 4; F is the metal forming


force, v is the tool feed and is the spindle rpm.
The two common types of AISF are Two Point Incremental
Forming (TPIF), and Single Point Incremental Forming
(SPIF). These are discussed in the chronological order in
which they appeared, historically.
In both cases there is a single forming tool whose motion
is usually described in terms of Cartesian coordinates, with
tool motion in the horizontal sheet plane labeled as the xaxis and y-axis, and the vertical z-axis being the direction
in which deformation occurs.

of a series of incremental contours. Each toolpath profile


consists of a contour at a constant depth, and the
subsequent contours, in the z-axis direction, are at levels
offset down by a z increment. Sharp edges may be
created in this process, depending upon the forming tool
diameter used, since the tool follows the convex surface of
the part during forming. Contours for a simple cone are
shown in Figure 6. To form this shape the tool would
follow the top contour, and then move incrementally
downward along the z-axis, until finished.
Shear forming theory can be applied. Experimental work
with cones formed at various wall angles [27] shows
measured thickness matches thickness calculated with the
sine law. Because the flange material remains
undeformed, the wall thickness can be easily calculated by
the sine law, assuming a constant volume.
Forming tool

blank
Bushing

Figure 4. The basic elements needed for asymmetric


Incremental Sheet Forming (AISF).
Asymmetric Two Point Incremental Forming
TPIF was introduced by Powell and Andrew [26] and used
Matsubara [27] to meet the need for quick, inexpensive
production of low volume asymmetric sheet metal parts.
With the TPIF process, the metal blank moves vertically on
bearings, which move on blankholder posts, along the zaxis, as the forming tool pushes into the sheet metal.
This process has two points where the sheet metal is
pressed, simultaneously, hence it is called Two Point
Incremental Forming (TPIF) to differentiate it from Single
Point Incremental Forming (SPIF) which has just one point
at which force is applied. See Figures 3(c) and 3(d). The
point where plastic deformation occurs is directly under the
forming tool. When it is used in a CNC mill, it is mounted in
the spindle. The forming tool pushes down on the sheet
metal, causing plastic deformation at a point, while tracing
a path, which is the outline of the shape being
manufactured. In TPIF, the other point is a static post that
creates an upward counter force on the sheet. One tool
presses into the sheet and the other acts as a partial die.
Because of the partial die, TPIF is not truly dieless,
although it is often called that.
The TPIF equipment, shown in Figure 5, consists of an
apparatus, which clamps the sheet metal (blankholder),
and allows for downward movement with toolpath
increments in the z-axis direction. The centre of the blank
is supported with a stationary post (a partial die) and a
clamped perimeter (blankholder) that moves down on
bushings as deformation of the sheet progresses. To
prevent twisting of the shape about the partial die, and also
provide a back pressure on the strip, there is a support
plate underneath the blank. The partial die can be replaced
by a shape that acts as a mould (full die) over which the
sheet is formed incrementally by the single point. Both
configurations are illustrated with details in Figure 3.
The forming tool is a steel rod with a smooth,
hemispherical tip, for instance a 12 mm diameter. The
motion of this tool is controlled through CNC programming

Support Post

Figure 5. Backward bulge forming apparatus [27].

y
x
z

Figure 6. Toolpath contours for a cone [27].


Asymmetric Single Point Incremental Forming
The asymmetric single point incremental forming (SPIF)
research performed by Jeswiet [29], Leach [31] and Fratini
[32] has shown the SPIF forming method can be
performed on a standard three-axis CNC mill. This
includes application of CAD/CAM software to plan the
process toolpath allows for easy fabrication of complex
parts. There is one major difference with respect to the
apparatus shown in Figure 5 and that of Jeswiet [29] and
Leach [31]. It is the lack of a partial or full die. In SPIF the
back surface of the sheet being deformed is a free,
unsupported surface. This creates different strain and
stress patterns in the sheet as it is deformed, in
comparison to TPIF.
Chronologically, SPIF followed TPIF. It became apparent
that any CNC milling machine, when used in conjunction
with toolpath planning software, could be used to make
sheet metal parts incrementally [29, 31, 32], as shown in
Figure 3(a). It can be seen that this configuration is truly
dieless forming as envisioned by Leszak [2]. It has been
shown that shear forming can also be applied to this
process [33].
Process Advantages and Limitations
The advantages and disadvantages of SPIF are as follows
[34].
Advantages:

Useable parts can be formed directly from CAD data


with a minimum of specialized tooling. These can be
spindle speed , in terms of feed rate ,
either Rapid Prototypes or small volume production
tool radius r, and wall angle
runs.
The process does not require either positive or
Invoking the cosine law where:
negative dies hence it is dieless. However it does need
c 2 = a 2 + b 2 2ab o cos
a backing plate to create a clear change of angle at the
setting c = dmax ; a = b = r; = 2
sheet surface.
forming
Changes in part design sizes can be easily and quickly
d max = r 2(1 cos )
tool
accommodated, giving a high degree of flexibility.
1
Making metal Rapid Prototypes is normally difficult, but
d =r
(1 cos )
easy with this process.

2
The small plastic zone and incremental nature of the
r
d
1
d =r
(1 cos(2 ) ; d = max
process contributes to increased formability, making it
2
2
easier to deform low formability sheet.

=
spindle
speed;

=
feed
rate
A conventional CNC milling machine or lathe can be
used for this process.
v
v
=
=
The size of the part is limited only by the size of the

d
1
dmax
machine. Forces do not increase because the contact refer 23()
r (1 cos 2 )
2
zone and incremental step size remain small.
The surface finish of the part can be improved.
Figure 7. Tool geometry and spindle speeds.
The operation is quiet and relatively noise free.
Disadvantages:
To keep friction heat minimal the tool must roll over the
The major disadvantage is the forming time is much
surface of the work piece as it is formed. This result
longer than competitive processes such as deep
requires that the distance traveled along the work piece
drawing.
(i.e. the feed rate) be equal to the average circumference
As a result the process is limited to small size batch
of the tool in contact with the material multiplied by the
production.
spindle speed. The following equation, derived in Figure 7,
The forming of right angles cannot be done in one step,
describes this mathematically. Spindle speed and feed rate
but requires a multi-step process.
are represented by and respectively and the
springback occurs, however algorithms are being
hemispherical tool radius is r.
developed to deal with this problem.
v
(1)
3.3 Forming Tool Spindle speeds
=
1
One major difference between the different sheet
r (1 cos 2 )
incremental forming processes, described by Hagan [9]
2
and other users of the process [29, 30, 31, 32], is the way
the tool moves while deforming the sheet. In the case of
Using Friction Heat
SPIF the following have been done: 1) move the spindle
without rotation, 2) move the spindle with the spindle
Increased spindle rotational speed is used sometimes to
rotating, at different spindle rotating speeds.
increase Formability [35]. The Formability increase is due
to both a local heating of the sheet and, what is more, a
In the second case the spindle rotates so that the forming
positive reduction of friction effects at the tool-sheet
tool rolls over the sheet surface. Controlling this variable
interface.
controls the heating of the sheet during deformation.
Spindle, Free Rotation
Reduced Friction Heating Speeds
In a case study of manufacturing a solar oven cavity, the
The forming tool has a hemispherical shape, which is
spindle could rotate freely in a CNC mill [36]. This allowed
pressed into the material to cause deformation as shown in
the friction at the tool/workpiece to cause the tool to rotate
Figure 9.
at a speed that automatically matched the spindle surface
The most obvious source of heating is friction. As the tool
rotation speed. This method is also used by a machines
travels over the surface of the work piece it is also spinning
specially built for Incremental Forming [37 38, 39, 40].
at a certain number of revolutions per minute. If the tool is
3.4 Equipment used for Incremental Forming
stopped it will slide along the surface of the material. In all
cases heating will occur due to sliding friction. If the tool is
The total package needed to incrementally form sheet
rotated at a high speed, the tool surface will slide over the
metal consists of a forming tool and the machinery that
work piece much more often and there will be excessive
moves the forming tool in a controlled manner. These are
heating. The relative motion of the surface of the tool, to
discussed in the following.
the surface of the work piece, is directly proportional to the
The main element is the single point forming tool. Solid
heat generated by sliding friction.
hemispherical tools are usually used when plastically
If the relative motion between the tool surface and
deforming sheet metal incrementally. A wide variety of
workpiece is small during forming (i.e. all friction is rolling
solid tools is used, however, other types of tooling, such as
friction, and not sliding friction) the heating is minimized.
water jets, are being investigated and these are reviewed.
Tools are designed and made by the users, they are not
For the draw angle, , there will be a point where the sheet
yet part of an assortment made available in the market.
is tangent to the hemisphere. This is the location of the
maximum diameter of contact (dmax). From then on the
Solid Forming tools
work piece is in contact with the tool down to the very
A solid hemispherical head is generally used for
bottom of the sphere, at which point the diameter of
asymmetric single point incremental forming; see Figure 8.
contact is zero. This is an assumption. The average
This assures a continuous point contact between sheet
diameter of contact is therefore half dmax, see Figure 7.
and forming tool; see Figure 7. At very steep wall angles it

can become necessary to use a smaller tool shank than


the sphere diameter. Contact between shank and sheet
metal is avoided this way. This must be taken into account
while generating the toolpath. Once a tool shape is
established, usually a specific radius with a hemispherical
ball-head, tool materials must be chosen.
In most
instances, the ball-head tools are made out of tool steel,
which is suitable for most applications. To reduce friction,
and to increase tool lifetime, the tool can be coated with or
even be made out of cemented carbide; see Figure 8. For
some tasks plastic tools are necessary to avoid chemical
reactions with the sheet material and thus increase the
surface quality. Wear of the tool can then become an
important consideration. In addition, lubrication helps
reduce the wear.
Next the diameter of the ball-head must be chosen. A wide
range of tool diameters is used, starting at small diameters
of 6 mm and going up to large tool diameters of 100 mm
for the manufacturing of large parts. These require much
more power because of the much larger angle of contact
involved. The diameter used depends upon the smallest
concave radius required in the part. It also has an
influence upon the surface quality and/or the
manufacturing time. Furthermore small tools can reach
their loading limit while forming materials like stainless
steel or titanium. The most commonly used diameters are
12 mm and 12.5 mm [27, 28, 29, 30, 31, 32].
As indicated in section 3.2, special measures can be taken
to reduce the friction between the tool and strip. Methods
of controlling the relative velocity are discussed in the
section on speeds used in forming. In addition to concern
about friction heating, there is a concern about the surface
roughness of the deformed sheet metal, which can be
influenced by both the tool size and the friction. Where
surface quality is a concern, the relative velocity can be
controlled or a tool designed specifically for the deep
rolling of surfaces can be used; see Figure 9. This tool
consists of a ball supported by pressure fluid that can
rotate freely in all directions and thus decreases friction on
the sheet.
Incremental forming and related flexible forming methods
and their generic tooling
Several other tool configurations are being investigated for
their ability to plastically deform sheet metal at a local
area. Some of these are based upon techniques that have
been in existence for sometime, such as shot peening, and
others are newer using much different deformation media,
such as water jet forming and laser forming.
In the case of shot peening, the sheet metal is hit
repeatedly with a large number of small balls made of
materials such as cast steel, glass, or ceramic. The size
can vary from 0.125 mm to 5 mm in diameter. This process
is used, traditionally, to create compressive stresses on
the surface of bulk products. However, while shot peening
has been used as a forming method since the 1950s [41,
42, 43], recent work by Kopp [44] has demonstrated the
possibility of creating both convex and concave shapes by
simultaneously using double sided shot peening.

Figure 8: Cemented carbide tools with 6, 10, 30 mm,


and a plastic tool (right) [37].

Sleeve, z-axis
Socket (fixed)
Flange
(rotating collet)

Pressure fluid
ca. 320 bar

Forming
tool
aperture

CBN-ball
13 mm

Figure 9. A universal tool head [37].


This approach results in improved plastic deformation,
while a well-chosen balance between the kinetic energy
supplied from both sides allows a specified curvature of
the work piece. Both convex and concave shapes can be
achieved in this way. This effectively opens perspectives
for a flexible forming tool for larger, single or double-curved
surfaces, characterized by large radii of curvature. The
potential for expanding the process window to generic part
shapes, including strongly curved surfaces, has not yet
been thoroughly researched. Therefore it has not been
selected as an explicit focus point for this paper.
Water jet forming is a relatively new process and also has
potential as a SPIF tool. Water jet cutting has been used,
successfully, in many cutting situations since its
development in the early 1980s. An attempt to apply water
jet technology to incremental forming was carried out by
Iseki [45], who explored the possibility to incrementally
bulge a sheet metal. Subsequently Jurisevic et al. [46]
analyzed the modifications that are necessary in order to
utilize standard water jet cutting systems for incremental
forming. They found water pressures up to 50 MPa are
needed and with a higher water volume flow (up to 50
l/min) in comparison to cutting applications. The above
mentioned researchers carried out their experiments on
0.5mm thick AA6082 sheets. Water jet forming is still at a
very preliminary research stage and requires a wide
research effort to investigate its possibilities and
limitations.
Laser forming is a new technology developed in the early
1990s [47, 48, 49]. It also shows promise as a forming
tool in the asymmetric incremental forming of sheet metal.
In this technique a local area is heated repetitively, with a
two phase heating and cooling process along a selected
path. The sheet bends in a preferential direction, usually
depending upon the resident stresses in the sheet.
Research on laser forming has dealt with the analysis of
process precision in order to obtain higher accuracy as
required by the industrial applications [50]. Recently,
closed loop control systems have been developed to
improve the effectiveness of laser forming processes in the
achievement of even complex 3-D shapes [51]. The
investigation of laser forming to make complex industrial
parts requires a further research.
Male el al. [52] followed up on the laser forming techniques
by using a plasma arc in place of a laser. The main
reasoning given is less expense. Some shapes have been
formed successfully, however, this is still in the embryonic
stage.
Blank holders and platforms
Rigid blankholders are used to clamp the sheet. For SPIF,
this is a rigid apparatus consisting of a metal plate on four
posts, to which the sheet can be clamped. To achieve a
high accuracy and to avoid an undesired draw-in of the
sheet, different faceplates can be mounted on the
apparatus.

Moveable blank holders are used to enable asymmetric


TPIF with either partial or dedicated dies. In two point
incremental forming, a vertically moveable blank holder is
essential. In this case the manufacturing starts at the
centre of the desired geometry. As the tool descends,
gradually going down in small increments, the blank holder
also descends as the forming process progresses; see
Figures 3, 5 and 10.
Forming Machinery
In general all CNC-controlled three-axis CNC machines
are suitable to perform ASIF. High speeds, large working
volumes and sufficient stiffness are favourable. Milling
machines are available in different designs, which differ in
working volume, maximum feed rate, maximum load,
stiffness and cost prices.
The following includes machines, which can be used for
AISF. In most cases they can be used for other machining
processes, hence they are multi-purpose. Only three-axis
mills have been used to date. One manufacturer makes a
specially designed machine that is dedicated to
incremental forming only [38]. Hence, it is less flexible
according to the criteria set out by Schmoeckel [1].
A list of the types of machines available to do incremental
forming is:
CNC milling machines;
Purpose built machines;
Robots;
Stewart platforms and Hexapods.
Common, applicable, shop, CNC milling machines that can
be used [53] are:
Gantry milling machines; large working volume, high
speed drives, high forces, expensive.
Gateway milling machine; large working volume, high
feed rates, high forces and stiffness, expensive.
Bedplate type milling machines; large working volumes
available, width is limited, low cost price, lower stiffness.
Console milling machine; available in almost every
workshop, low cost price, relatively small working
volume. This is the only type of three-axis CNC mill used
to date.
A dedicated, single use, forming machine is commercially
available. It has high feed rates, medium sized working
volumes, and is equipped with a controlled movable blank
holder. See Figure 11. It is based upon the technology
developed by Matsubara [26] and Amino [36] including a
patent by Aoyama [38].

Figure 10: Upgraded conventional milling machine with


moveable blank holder [36].

Figure 11: A purpose built AISF machine [37].

Figure 12. One-off design by Allwood [40].


The concept of forming sheet metal incrementally with a
single point has spawned other designs such as the
purpose built, one-off design by Allwood [40] in Figure 12.
A whole other class of potentially usable machines is
available. Some can be used for re-entrant shapes. These
are being tried currently as follows.
Industrial robots have a large working volume, fast drives,
low stiffness, and very low maximum allowable forces.
Several institutes are trying to apply robots to incremental
forming [54, 55, 56]. This method of forming is in the
embryonic stage and looks promising. A special case of a
robot application is, that in place of a continuously moving
rigid tool, there is incremental forming by hammering. In
this case, the tip of the forming tool has a fast oscillating
movement beating the sheet into the desired form [55].
This device is especially suited for the application on an
industrial robot because it partially compensates the
lacking of stiffness of the robot arm.
A Stewart Platform [57] offers infinite degrees of freedom.
None are being used, but the potential is great, especially
if compared to five-axis milling machines.
Special tooling is also being developed. For instance, to
reduce the tooling support effects in two point incremental
forming, Iseki [58] has developed a set of universal roller
punches on the opposite side of the universal forming tool
to support the geometry to be formed. This multi-roller tool
means more flexibility but has the disadvantage of
requiring the numerical control of several tools.
It has been shown that there is a wide diversity in the type
of machine that can be used to achieve asymmetric Single
Point Incremental forming. A manufacturer can choose
between dedicated machines or flexible machines as
envisaged by Schmoeckel [1].

4 MATERIAL PROCESS PARAMETERS


Material process parameters and toolpath generation are
at the heart of AISF. One of the first steps in a design is to
create a CAD drawing of the part. For AISF processes the
drawing may be designed in-house or be imported from
third parties and modified as required. Fortunately, CAD
drawings are usually made with a commercial package
that allows the user to create a file in a neutral data format.
This allows the CAD file to be exported easily into a CAM
package. The procedure followed in making a part, from
receiving a CAD model to generating a toolpath, is shown
in Figure 14.
The material parameter, draw angle , defined in Figure 2
is important in AISF. The largest design draw angle in the
part, d, must be less than the material parameter max, the
maximum value of the draw angle, which is a material
characteristic. An engineering designer can see from the
CAD drawing if max has been reached and by making
judicious choices when embedding the part in the sheet
surface, d can often be made smaller than max. Once a
part has been embedded in the surface of the sheet metal
the comparison between d and max will show if the part
can be made in one pass, two passes or several passes,
after which toolpaths can be generated.
4.1 Formability Criteria for Incremental Forming
The AISF process is characterized by increased material
formability. Much higher strains than normally observed in
deep drawing are observed in AISF, for instance strains
can be well over 3. Recent studies have shown that the
forming limit curve, which describes the formability in a
minor-major strain plane, may be expressed as a straight
line with a negative slope [32, 66, 67]. Bi-axial strains, in
the deformed sheet, are measurable only during particular
conditions, such as corners or plane intersections, with
most strains localized close to the major strain axis, max.
For this reason, material formability can be represented as
an artificial index, which is the draw angle , as shown in
Figures 2 and 15.
AISF process mechanics are fully characterized by a
small, localized plastic zone, which is limited to the small
area between the tool and workpiece. Visioplastic
evaluation shows the deformation mode is very close to
plane strain deformation [60] and is almost pure stretching
in the forming area [35]. Much higher strains are possible
due to the presence of hydrostatic pressure caused by the
elastic deformation of the area surrounding the confined
plastic deformation zone [61].
Using a combination of max for a material at a specific
thickness, and forming limit diagrams, an engineering
designer can decide if a part can be made in one pass
without tearing, or if a two pass or multiple pass sequence
should be used. These are discussed in the following.
4.2 Maximum Draw Angle, max: a formability
benchmark
It is recognized that the material behaviour and maximum
formability in AISF can be described by the maximum
value of the draw angle max [32, 33, 60, 61, 62]. As
increases the thickness reduction reaches a minimum
value where fracture occurs as a consequence. This is
related to the both the sine law and spinnability relation
shown in Figure 2. Knowing the parameter max for a
material at a specific thickness, a designer can take the
first step in deciding if a sheet metal part can be made in
one pass without tearing, or if a two pass or multiple pass
sequence should be used; see Figure 14.
Many parameters influence the process and have to be
considered to fully understand the process mechanics and

formability results. For this reason, experimental


investigations have been developed, taking into account a
benchmark product. If the details of max, and formability
are unknown then the cone shown in Figure 15 is used to
conduct tests [35]. The benchmark cone base diameter,
Do, is 72 mm with a height, H, equal to 40 mm. Tests are
carried out at varying degrees of draw angle of the cone
up to failure, which is defined as tearing in the specimen
wall.
Table 1 shows examples of material initial thickness and
corresponding values of max that have been observed in
AISF. The rule for spinning, in Figure 2, and the calculation
for spinnability [18], both indicate that material thickness
has an impact upon the formability of the material, and by
simple geometry, see Figure 2, it can be seen that the

spinning angle and draw angle are related directly.


CAD file
arrives

Definitions:
d = largest draw angle in a CAD
max = the maximum draw angle for
a material
Embed shape in
the sheet metal
surface; use
minimum angles.

Check
forming limit
diagrams for
formability.

Check for
the largest
d ;
if d > max

List of materials
available and
their properties

Second check
of forming limit
diagrams for
formability.

Check
where the
maximum
d occurs
Choose a material.
Find the maximum
draw angle, max
for that material.

Compare the maximum drawing angle


to the largest forming angle expected:
d > max ; d < max.

d < max.

d > max.

Generate toolpath
Use two pass
method

Use one pass


method

Figure 14. Procedure followed for making a part.

D0

10 mm

Figure 15: Cone for testing forming parameters [35].

max

Material
AA 1050-O

FLDo

67.5 2.305

to , mm
1.21

Filice [31]

1.0

Micari [63]

AA 6114T4

60

Al 3003 O

78.1

2.1

Jeswiet [59]

Al 3003 O

72.1

1.3

Jeswiet [59]

Al 3003 O

71

1.21

Jeswiet [59]

Al 3003 O

67

0.93

Jeswiet [59]

Al 5754 O

62

1.02

Jeswiet [59]

Al 5182 O

63

0.93

Jeswiet [59]

AA 6111T4P

53

0.93

Jeswiet [59]

DC04, mild steel

65

1.2

1.0

Hirt [65]

DDQ

70

2.718

1.0

Micari [63]

HSS

65

1.924

1.0

Micari [63]

Copper

65

1.808

1.0

Micari [63]

Brass

40

0.701

1.0

Micari [63]

0.841

3.0

Table 1: A list of materials with initial thickness and


maximum draw angles. FLDo is the maximum major strain
at a minor strain of zero.

to = 1.5mm
= 40, t1 = 1.15mm
= 60, t2 = 0.75mm

to= 1.5mm

Figure 16: Asymmetric incremental forming of 1.5 mm thick


DC04, into a pyramid with varying slope sides [60].
Using a pyramid shape, see Figure 16, Hirt [60] found the
same result for material limits as Fratini [67]; that thinner
sections occur with steeper sides and increasing . The
process limit fracture is approximately 65o for AA 1100
aluminum and DC04 mild steel.
Jeswiet et al [66] conducted a parametric study, aimed at
determining the relation between material formability
(measured to the maximum safe value of max) to sheet
thickness. Testing two aluminium alloys gave:
AA 3003O

max = 8.5to + 60.7

(2)

AA 5754O
max = 3.3to + 58.3
(3)
The upper and lower limits to the linear function are
defined by spinnabillity [18] and the sine law. Although the
these linear functions have limitations, they are useful to
the designer in giving choices for combinations of
thickness and draw angle for single pass forming.

4.3 FLDs and forming limits


Forming limit diagrams are traditionally one of the tools
used to decide if a material of a particular thickness can be
formed by a deep drawing process. This has been applied
to AISF.
Work on AISF forming limit diagrams has been performed
by Felice and Micari [32, 59], Hirt et al. [61] and Young and
Jeswiet [62]. Micari used an agreed upon shape [59] to
develop FLDs while Hirt et al. [64] have used a pyramid
with varying slopes and Young and Jeswiet [62] have used
a grouping of five shapes for one FLD. These are all
presented in the following.
Traditional Forming Limit Diagrams
The single point incremental forming of a cone, shown in
Figures 15, 17 and 41, illustrates the use of FLDs. The
sequence of steps followed by the tool are shown. For the
case shown, the z incremental steps are labeled 5, 9, 13,
etc. and the x steps are labeled 4, 8, 12, etc. The
hypotenuse of x and z is at the drawing angle .
The four parameters usually of interest are: thickness of
the sheet, size of the step down, z, speed of deformation
and size of the forming tool.
Step down: The size of the step down, z, (pitch) has a
significant influence upon formability and surface
roughness. With increasing z the blank undergoes
heavier deformation conditions. Micari [59] conducted step
down tests on AA 1050-O, 1 mm thick sheets with a cone
configuration as shown in Figure 15. It was found that
sheet formability decreases as a direct consequence of
increasing the tool pitch z. The results are summarized
by means of FLDo in Figure 18. A boundary area, which
corresponds to the better choice for technological
parameters, is highlighted. This has also been observed by
Hirt et al. [60] and Hagan and Jeswiet [63].
Increasing spindle angular speed (spindle rpm) can
increase formability [59]. The formability increase is due to
both a local heating of the sheet and, what is more, a
positive reduction of friction effects at the tool-sheet
interface. There is a negative aspect in that the forming
tool wears very quickly, plus lubricants tend to burn
thereby creating safety and environmental problems. If the
material, AA 1050-0 in this case, has a relatively lower
formability, then a more formable material such as AA
8008-0 can be tried if lower spindle rpm is needed [36].
Forming tool diameter: An important role is played by the
forming tool diameter where a small radius concentrates
the strain at the zone of deformation in the sheet under the
forming tool, while a larger radius tends to distribute the
strains over a more extended area. As the forming tool
radius increases the process becomes more similar to
traditional stamping, thereby reducing formability limits.
Micari [59] found decreasing tool size increased the
forming limits; see Figure 18. Results found by Hirt [60],
see Figure 19, show that as the tool diameter decreases
from 30 mm to 6 mm, much higher strains and
deformations can be achieved.
It has been shown that sheet thickness has an effect upon
the maximum draw angle. Hirt et al. [68] showed this to be
true for forming limits with a study on DC04 taking into
account three different sheet thickness while fixing the
other experiment parameters. Similar trends were found by
Kim and Park [70] with AA1050O sheets. Their tests were
carried out on very thin sheets (lower than 0.3 mm) with
FLDo reductions from 1.2 to 0.92 for a sheet thickness
decrease of 70 percent.

Figure 17: Single Point Incremental Forming of a cone.

3.5

max, FLDo

Upper limit

2.5
2
1.5

Lower limit

1
0.5
0
0.00

1.00

2.00

3.00

4.00

Depth of step, z, mm
Figure 18: FLDo for different step sizes for AA 1050-0, with
upper and lower bounds, with a 12 mm diameter tool [59].

2.0
1.8
1.6

6 mm
tool

Major strain

1.4
1.2
1.0
0.8

10 mm
tool
30 mm
tool

1
2
4

x 2
x

1 x

Kim & Park [70] focused their attention on the influence of


anisotropy on formability. For this purpose, a set of
measurements of the major and the minor strains were
carried out both along the rolling direction (RD) and the
transverse one (TD). The tests were developed for
pyramid specimens with a varying tool diameter. The
material was the aluminium alloy 1050-O, with E = 70GPa,
= 33MPa, R0 = 0.51, R45 = 0.75, R90 = 0.48. They
concluded that formability along the transverse direction is
greater when small diameter tools are utilized, while along
the rolling direction it is larger with large diameter tools.
In order to fully understand the increase in formability with
AISF, a simple FEM was developed by Micari et al. [59]
and Bambach et al. [69]. They found that with decreasing
step size, z, the strain increments imposed at each loop
decrease and any point is overlapped other active loops,
while strains increase with increasing wall angle. Also,
from a stress point of view, a negative mean stress
distribution is observed under the tool and in the nearby
elements; in this way, the tool action postpones ductile
fractures during the process, until the tool is out of contact
with the sheet. Finally, at decreasing z the stress value
along the wall decreases too, so that a higher deformation
can be imposed without tears occurring.
Nontraditional Forming Limit Diagrams
Forming limit diagrams usually have the dashed V-shape
shown as FLC in conventional forming in Figure 20.
However, extensive research has shown that not only are
much higher strains achieved in this process, but the
Forming Limit Curve (FLC) in SPIF has a negative slope
as shown in Figure 20 [32, 62]. In one case, the maximum
forming strains [32] were actually much higher because the
AA 1050-0 sheet used in the experiments had been etched
with a grid causing stress concentrations and hence
premature failure to occur.
Young and Jeswiet [62] used more than one shape to
develop a composite FLD for 1.21 mm thick AA 3003-0.
They used five different shapes with varying angles, and
convex and concave curves. The shapes were: a dome, a
cone, a hyperbola, a pyramid and a shape with both
compressive and tensile stresses. These all have their own
FLD, and when all placed in one, common FLD, they
represent most combinations of shape found in a part. An
example of their work is shown Figure 21. The boundary
for safe forming, without any tearing, is shown by the
dashed red line. It can be seen that both Filice [32] and
Young [62] have the same slope for the boundary and that
very high strains were achieved.
It can be seen that the non-traditional FLDs being
developed [32, 62] provide the designer with an additional
method of judging if a design can be made in one pass
with AISF.
FLC incremental forming

1.0
0.9

0.6
0.2
0

Scattering band for


mild steel (t=1.5 mm)
-0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4
Minor strain

Figure 19: FLD for to = 1.5 mm DC04; influence of forming


tool size upon forming limits [60]. Graph points x1 to x4
correspond to positions on the sheet marked by x.

Major strain, max

0.8

0.4

0.7
0.6
0.5
0.4
0.3
0.2
0.1

FLC conventional forming


min

-0.2 -0.1 0

0.1 0.2 0.3 0.4 0.5 0.6

Figure 20: FLC for 1.21 mm thick AA 1050-0 forming of


pyramids [32]. Both traditional and non-traditional curves.

max
2.5

min = max + 3

2,5
2

1,5

1.5
1

0.5
-1.0

-0.5

major strain, max

0,5
0.5

1.0

0
min
Figure 21: Composite FLD for five parts formed by
asymmetric single point incremental forming of 1.21 mm
thick AA 3003-0 [62]. Each colour a shape and each point
represents a successfully formed shape.
4.4 Formability and material properties
Materials all have different formabilities and Table 1 in
section 4.2 shows how materials have a different
maximum draw angle max. In most sheet metal operations,
formability is limited by a local necking instability. Once a
neck is initiated the high hydrostatic stress that develops
within the neck leads to rapid void nucleation, growth and
coalescence [60, 72]. However, in certain forming
operations, such as bending and stretch flanging, necking
is suppressed and formability is controlled by the evolution
of void damage and shear band instability. If the strain is
evenly distributed in the metal, as it is in SPIF, very large
strains can be achieved due to the foregoing void damage
and shear band instability.
A comprehensive analysis was instigated to understand
the correlations between material formability and other
material properties [59]. The experimental campaign was
developed with a set of different blank materials, typically
utilized in the automotive and other sheet forming
industries. The materials used and their main
characteristics are summarized in the Table 2.
As before, the formability results were formalized again
through the FLDo values, while the testing parameters
were fixed for all the experiments. An FLD of all the
materials in table 2 is shown in Figure 22.
An FLD does not give information on which material
parameters predominate. Therefore a statistical analysis
was carried out to determine the influence of the material
properties in table 2. Five variables (K, n, Rn, UTS, A%)
were chosen as inputs, each one with five replications with

-0,5

minor strain, min

-0,25
0
0,25
0,5
Copper
HSS
AA6114
Brass
DDQ
AA 1050 - O
Figure 22: An FLD of the material parameters shown in
Table 2 [59].
FLDo fixed as the output. The intention of the analysis was
to determine how each input variable influenced the output
parameter. The experimental results were analysed by a
response surface statistical model where the qualitative
influence of the predictors (pure linear terms and
interactions among inputs, i.e. hybrid quadratic terms) on
FLDo were reported. The results indicate the highest
influence on material formability for AISF processes is the
strain hardening coefficient, n. In addition, there is a high
influence of a combined strength coefficient and strain
hardening coefficient. Strength coefficient and elongation
also have an influence upon the formability. The statistical
analysis leads to a six dimension hyper-surface:
FLDo = 8.64 - 36.2n - 0.00798K + 0.373Rn - 0.104A% +
0.0301Kn + 0.607n A% [59].
(4)
Jeswiet et al. [66] also found that both strength coefficient
and strain hardening coefficient had a major influence and
that a more accurate model of the stress-strain condition
could be obtained with a Voce model [71]. Verification of
this can be found in work by Lievers et al. [72] who
investigated three different automotive aluminum alloys in
developing a model of stress strain characteristics:
AA5182-0, AA5754-0 and AA6111-T4P. Their model
shows the Hollomon-Ludwig [73] relationship fits the
stress-strain curve at low values of strain, but at higher
strains the plastic section of the curve was more accurate
when fit to a Voce curve of the form:
(5)
M = s s y eq

where

( )

q = Mp

M is the flow stress matrix and MP is the effective plastic


strain. Youngs modulus, E, and Poissons ratio were taken
to be 70 GPa and 0.3 respectively. s is the saturation
stress, y is the yield stress, and both and are curve fit
parameters. Material descriptions and fit parameters for
the three materials are given in table 3.

Table 2: Materials and properties used to test correlations


with formability by asymmetric incremental forming [59].

Specific materials and material properties


Common sheet materials are employed in AISF with the
most common being aluminium and steel. Aluminum and
its alloys are used most frequently for AISF, mainly
because of the reduced forming forces. The majority of the
experiments with aluminium (AA 1100, AA 3003 and AA
1050) have been to investigate the formability and material
properties of the material. Fratini et al. [67] and

t
mm

MPa

MPa

AA5182-0

0.93

0.88

118

379

7.35

0.890

AA5754-0

1.02

0.76

93

298

7.48

0.844

AA6111-T4P

0.93

0.70

164

455

5.42

0.837

5050?m
m
DC04, initial state

505 m
?m
o

DCO4, = 80 , z = 1.75mm

Table 3: Material descriptions and fit parameters for the


Voce curve for the three alloys [72].
Jeswiet et al. [66] focused on the investigation of material
models for two common sheet materials, AA 1050 and AA
3003.They found appropriate material models for
Aluminum, as follows:
1050-0, to = 1.0 mm, = 111

0.14

3003-0, to = 1.21 mm,

= 152

0.213

[67]

(6)

[76, 77]

(7)

Due to the higher forming forces required for the


processing of steel materials, the sheet thickness
spectrum investigated so far is smaller. Nevertheless, steel
materials are used for the manufacturing of incrementally
formed parts since they have an important role in the field
of conventional sheet forming. Hence, there is a focus on
the use of mild steel and stainless steels. Among the mild
steel materials, DC04 is by far the most commonly used
for incremental forming due to its high formability and its
numerous applications in the field of sheet metal
components. Amino et al [76] manufactured a fender and
the inner of a car seat out of 0.8 mm DC04 sheets.
Changes in a materials microstructure induced by the
incremental forming process have also been investigated
[58]. The results of this investigation are displayed in
Figure 23. This was done with common steel materials,
DC04 and 1.4301. In both cases, the grains are
significantly elongated due to the high strain. Besides mild
steel, various kinds of stainless steel have been employed
for incremental forming, but mainly for research purposes.
Amino et al [76] have manufactured a stainless steel
incubator bed and an exhaust part and Hirt et al. [78] have
made various stainless steel demonstrator parts.
Generally stainless steels are significantly more difficult
than mild steels or aluminum when incrementally formed.
This is mainly due to their high tensile strength, the high
strain hardening coefficient and their distinct inclination to
elastic spring back. Many other materials have been
employed to perform experiments on incremental sheet
forming. These include brass, copper, silver, gold, titanium
and platinum. Because of its frequent application in the
field of aircraft manufacturing, titanium occupies an
important position among these. Some demonstrator parts
have been formed successfully out of RT12S (Ti grade 1)
[79], which is also suitable for deep-drawing.
4.5
Surface roughness
Surface roughness is a major concern in a final product. In
AISF the major factor in determining surface roughness is
the incremental step size, z [88, 89, 90]. Two examples
are shown in Figure 24.
In one in-depth study of the effect of pitch, z, on surface
roughness [88], the objective was to observe which
surface roughness indicator was more useful. Figure 25
shows the 3D Figures for different pitch sizes. Figure 26
shows 2D profiles for three pitch sizes with the tool profile
superimposed. At a pitch of 0.13 mm, when the pitch is 1
percent of the tool radius the ridging observed in Figure 25
virtually disappeared. It was also found that Rz, mean peak
to valley height was a more useful measure.

100
m
5 ?m
100?
m
1.4301, initial state

505 m
?m
o

1.4301, = 80 , z = 1.75mm

Figure 23: Change in steel microstructure with large angle


of deformation in asymmetric single point incremental
forming [59].
Junk et al also studied the effect of pitch z, draw angle ,
and tool radius upon surface roughness [89]. They found
that pitch size, tool radius, draw angle and surface
roughness were interrelated as shown in Figure 27 and
that with the right combination surface roughness due to
pitch and draw angle can be eliminated.
Spindle rpm has an effect upon maximum profile height
[88] and an equation, which relates maximum profile height
created by incremental forming to pitch size z, can be
developed. It has also been observed that surface
roughness is higher in the case of non-rotating tools [88].
By decreasing the relative motion between tool and
workpiece the surface roughness can be reduced. This will
also reduce the incidence of spalling due to repeated
stress over the same surface.
There is an unwanted effect of which the designer should
be aware. At high draw angles there is an orange peel
effect [33, 61]. The size of the effect is influenced by the
incremental step size, x and y, and the draw angle, .
This effect occurs on free surfaces with very large plastic
strains and is the result of texture and microstructural
effects [75].
4.6 Forces in Incremental Forming
Potential users of incremental sheet forming processes are
often concerned about the forces that are generated,
especially if a CNC mill is the being used. Groups at KU
Leuven and Queens University have designed special
sensors specifically for this purpose, with the results being
published recently [80, 81].
In one case, when measuring forces in SPIF, a force
sensor design was based upon friction measurement work
done with cantilever beams by Nyahumwa [82]. The
cantilever sensor design shown in Figure 28 is a spindle
mounted cantilever beam with strain gauged Wheatstone
bridges. Each bridge is designed to measure one of three
orthogonal forces: two bending directions, Fr and Ft,
and one axial direction, Fa. An added attraction of this
sensor is it can be used for friction measurement studies.
The forces measured with the sensor shown in Figure 28
are for 1.21 mm thick AA 30030, being formed from flat
sheet into a truncated pyramid shape. The force
measurement results are shown in Figure 29. Snapshots
are taken of the forces at percent completion of the part.
These snapshots show how the forces vary at different
points in the process. It can be seen the maximum forces
encountered are around Fa = 450 N in the axial direction.

x2
z2

h2
x1
material
ridge

z1

h1

Figure 24: An illustration of how the ridges occur due to


pitch size, z [89].

z = 1.27mm

z = 0.76mm

z = 0.51mm

z = 1.02mm

Figure 25: 3D Surface roughness for four pitch sizes, z.


Profiles were obtained by white light interferometry and
are 3.6 mm x 4.6 mm [88]. The tool diameter is 12.5 mm.

1.78 mm z
1.27 mm
0.51 mm

Profile height, mm
0.03
0.02
0.01

tool profile

-2.5 -2

0
-1.5 -1 -0.5 0
0.5 1
1.5
Distance Along Surface (mm)

2.5

Figure 26: 2D surface roughness profiles for three pitch


sizes [88].
RZ [m]
21,3

RZ [m]

19,1

20
15
10
5
0

11,5

as received
roughness
3,3
2,5
1,9
1,1
1,7
0,9

0.1

10,1
4,5
4,1
4,8
3,9

0.2
0.5
Pitch z, mm

10,4

20
15
10
5

7,3

47
65
m
47
5m
65
mm
m
1.0
15
,m
s
u
i
Wall angle
rad
ol
o
T

Figure 27: The influence of pitch z, draw angle , and tool


size upon surface roughness Rz; 1.5 mm AA 1100 [89].

Most of the forming energy goes into pushing down shown


by , Fa., the axial force, and that Fb, the resultant of Ft and
Fr, is much lower than Fa. See Figure 28 for definitions.
Forces were also measured in TPIF with a gauged partial
die using a Poisson gauge configuration [80]. In forming
cones, measurements were made at different draw angles,
. The forces were averaged for each complete circuit
around the partial die (called a profile #). Only peak values
are of interest, therefore not all profile data points are
o
shown, for instance the points for a draw angle of = 60
are not all shown. The peak values for draw angles of =
30o, 45o and 60o are shown in Figure 30. The maximum
values of force, experienced at a draw angle, , can simply
be found by dividing by the cosine of the draw angle, .
The maximum forces for both processes, measured by
different sensors, can be seen to be close in magnitude.
Aside from finding the maximum values of force, one
o
noticeable result is the large peak force at = 60 . This
peak value is thought to indicate the sheet is approaching
the maximum value of force that the sheet can sustain,
similar to the ultimate tensile strength, uts, observed in
engineering stress-strain diagrams. Hence this could be
used as an indicator of the approach of failure at the
maximum draw angle, max.
In the third case, Duflou et al. [81] conducted a series of
experiments with a six-component force dynamometer
mounted under a single point incremental forming rig and
derived relations between forces and slope angle, tool
diameter and incremental step size. The measurements
were performed on the reference conical part described in
Figure 15. Figures 31 and 32 provide an overview of the
results for 1.21 mm thick AA 3003-0. Table 4 gives a
summary of the results. In Table 4, Fp is the maximum
force reached during the forming process for a given part,
and Fs refers to the average force measured after a stable
level is reached (excluding the transient behaviour interval
and the peak force).
It was concluded [81] that if the vertical step size z, tool
diameter or wall angle , are increased, the total forces on
the forming tool, and thus the machine tool, also increase.
For these three parameters, the vertical step size has the
least significant impact and can therefore be increased
without penalty, to reduce part production times.
An increase in the tool diameter has a substantial impact
on the magnitude of force required to form a given part.
Although it can improve surface finish and reduce
production time by allowing larger vertical step sizes
without affecting surface quality, large increases in tool
diameter result in much higher forces and could become a
limiting factor. When the drawing angle increases the
forming forces are much greater.
Forces, that occur in deforming 1.21 mm thick AA 3003-0
sheet, have been measured successfully, with three
separate sensor designs and for two separate types of
AISF: SPIF and TPIF. Useful design/manufacturing
information is:
peak forces can be observed in the area where failure
occurs at maximum draw angles, max,
increasing the vertical step size z increases forces,
larger tool diameters increases forming forces.
Wall Thinning
The sine law, which is used in spinning [10, 13, 86],
assumes a uniform wall thinning. This is true in low strain
situations as shown by Matsubara [27], and Hagan [77].

Force Vector Sum, N

600
500
400

20
30
40
50
60

300
200
100
0

750
500
1000 1250
Time, s
Figure 31: Total force curves for conical parts with different
drawing angles for a 0.5 mm vertical step size, 10 mm tool
diameter, feed rate of 2000 mm/min [81].

Figure 28: Details of the Single Point Incremental Forming


Sensor [80].

Fa: axial force on sensor


Fr: radial, bending force on sensor
Ft: tangential, bending force on sensor

500
450
400
350
300
250
200
150
100
50
0
- 50

Ft
10

20

80

30
40 50
60
70
Percentage Complete

Figure 29: Forces measured in forming a pyramid at


o
= 40 in SPIF of 1.21 mm thick AA 3003-0 [80].

Peak forces
observed

N, ( lbf)
712, (160)
623, (140)
534, (120)

445 N

445, (100)
356, (80)

forming tool

60o
b

45o

289 N

(60)

cone 30
cone 45

(40)
(20)
0

F45o F60

30o 45o 60o

596 N

30o

cone 60
a

250

500

750 1000 1250


Time, s

Process
parameter

F(N)

R2

Tool diameter
(in mm)

Fp=12.761 +434.5
Fs=12.812 +410.8

Part wall angle


(in degrees)

Fp=6.793 +210.4
0.9826
2
Fs=-0.2284 +22.753 0.9919
47.0

Vertical step size


z (in mm)

Fp=151.58 z +466.9
Fs=148.13 z +447.17

0.9809
0.9884

0.9655
0.9725

Table 4. Experiment determined relation with total peak


force Fp and average force Fs [81].
static tool post

Force

D
of irec
to tio
ol n

Forces measured in
TPIF of AA 3003-0

10.0
12.7
15.0
20.0
25.0

Figure 32: Force curves for parts formed with tools from a
10 mm to 25 mm diameter [81].

Ft

Fr

250

800
700
600
500
400
300
200
100
0

Fa

Fb

Fr

tool
path

Force, N

Forces measured in forming a pyramid.

Force Vector Sum, N

10 20 30 40 50 60
# of Profile from top to bottom

70 80

Figure 30: Force measurement data, measured with a


gauged, static tool post, for three draw angles
o
o
o
= 30 , 45 and 60 , for 1.21 mm thick AA 3003-0 [80].

However, nonuniform wall thinning has been observed [35,


60, 77, 83]. Figure 33 is a typical example of wall thickness
measurements for 30 and 70 degree cones in SPIF [83].
The initial sheet thickness was found to be 1.2069 mm +/0.0180 mm, The sheet thins until it stabilizes at an
approximate thickness of 1.00 mm, and remains at this
value for the remainder of the cones surface until it thins
again slightly just below the apex of the cone. The uniform
portion of the cones wall profile is thinner than the
predicted sine law value of 1.10 mm, indicating that an
over-spinning condition is occurring [18]. The initial
deformation is due to bending with subsequent deformation
being due to shear.
The thinning zone shown as an inset in Figure 33 is a
property of the process rather than of the particular
geometry. Thinning is dramatic at high draw angles, and is
a precursor of failure just after max. The sine law does not
predict this band, and if the theory is solely followed,
unexpected failures will occur in this location.

1.40
1.20

sine law, 30o

Thickness, mm

1.00
0.80
0.25 mm
0.37 mm

0.60
0.40

sine law, 70o

0.20
0.00
0

10

20 30

40

50 60

70 80 90

Distance Along Surface, mm

Figure 33: Single pass wall thickness profile for 30 and 70


degree cones. AA 3003-0, to = 1.21 mm [83].

forming
tool
offset

4.8
Toolpath Generation
Once all the material characteristics are known a toolpath
can be planned. There are several choices available for
the forming toolpath shown in the following.

1.21 mm

bending
applied force

0.65 mm

Figure 34: Detail of bending observed during first pass of a


two-pass asymmetric incremental forming process [83].

4.7 Multiple
pass
asymmetric
single
point
incremental forming
To remedy thinning shown in the discrete zone in Figure
33, and to avoid failures, material available elsewhere in
the part can be used to stop both thinning and failure. In
their work on SPIF Kim and Yang [85] used a pre-form, or
two-pass forming as a means of equalizing the strain
distribution across the surface of a part. Jeswiet and
Hagan [29], and Hirt et al [61] also used the same
technique. This involves the creation of high strains in a
pre-form at areas where low strains are present in the final
geometry, and low strains at areas where high strains are
present on the final geometry [83]. Pre-forms can then be
designed to include a combination of wall angles and
artificially large offsets from the backing plate.
The hypothesis behind the steep wall angles and large
offsets is that this encourages the undeformed sheet in the
flange area to bend downwards over the backing plate and
into the part. This presents a thick surface that is already
inclined at a portion of the final angle desired in the
finished geometry. When shear forming theory for preformed blanks is consulted, a two-pass sine law is
employed [83, 84] as follows:

tf = tp

Sin ( f )
Sin ( p )

flange, then the wall thickness of a 70 degree cone would


increase to 0.72 mm. The traditional sine law gives a
prediction of 0.41 mm for the same 70 degree cone made
in a single shear-forming pass.
It is important to note that this increase in thickness will
only occur in the area of the part that was affected by the
bending of the flange. Unlike shear forming with a pressed
pre-form of constant thickness, the pre-form of the first
pass will vary in thickness from the thick bent flange area,
to the remainder of the part that experiences normal
thinning in the process; see Figures 33 and 35. Hence
two-pass forming serves to alleviate the thinning band
observed at high draw angles. The wall thickness of a
multistage formed pyramid [61] shows thinning at the initial
bend has been moved away from the initial bend; see
Figure 35. This technique was put to good use in the SPIF
of a rapid prototype for a car headlight reflector [29]. Four
steps were needed to create the reflector surface; see
Figure 36.

(7)

where: tf is the final wall thickness, tp is the wall thickness


of the pre-form, f is the final wall angle, p is the wall
angle of the pre-form.
For example, if a pre-form could be designed to bend the
flange inwards to a 55 degree angle while maintaining a
sheet thickness which is the same as the undeformed

Contour milling toolpath is a finishing pass, typically


defined by fixed z increments between consecutive
discrete contours. This is also the most common technique
used. The disadvantage is it leaves marks at the transition
point between layers and creates force peaks. Surface
quality depends on tool radius, step size [29, 31, 32], and
slope angle as well as lubrication system and spindle
speed.
A spiraling toolpath is continuous with incremental descent
of the tool distributed over the complete contour of the part
[32, 68]. The advantage is that no marks occur at step
down.
Multiple toolpath strategies include creating intermediate
forms that are defined within the cavity of the final surface
and are typically characterized by limited slope angles and
curvature. This is comparable to a roughing step in milling,
followed by a finishing pass that can be a conventional
contour milling or spiral toolpath or a strategically chosen
toolpath aiming at stretching out the cavity bottom and
increasing part slope angles without causing excessive
strains in the steeply sloped areas [83]. Both Figures 35
and 36 are examples.
Contour toolpath generation
Because it is the most common method used, this method
is discussed in detail. First, the flat plane of the sheet is
defined as the x-y plane before being deformed; see
Figure 37. This is an artificial horizontal plane that acts as
the original sheet reference when forming along the z-axis;
the z-axis is equal to zero at this point. All portions of the
required geometry must be at or below this plane. In the
case of SPIF the forming tool moves from the outside
edge, point a, toward the centre. In the case of TPIF this is
reversed and the tool moves from the centre out.
To make SPIF possible in one pass, a part should be first
oriented so that steep walls (which means a draw angle
equal to 65 degrees or greater) are reduced by rotating the
part around the x and y axes. Then the steep walls will
have a shallower angle relative to the z-axis, where
possible. This initial manipulation greatly increases the
ease with which the part can be formed successfully.
Once the workplane is set, a modification to the required
geometry is the addition of false surfaces to create an

Top surface
Top radius RTop
Flange
Preform area

sheet thickness t, mm

Blank
Bottom radius RB
1.6
1.4
1.2
1.0
0.8
o
sine
law;
Sine
lawf = 45
0.6
o
1100-0
Material: Al99.5
sine law; f = 81
0.4 Thickness:
to=1.5mm
s0=1.5mm
RR=15mm
0.2 Tool:
TT=15mm
Angle: =81
= 81o
0.0
0
20
40
60 80 100 120 140 160
Rtop
Top surface

Blank
Rblank
Preform area
x- coordinate [mm]

Figure 38: Positioning of an irregular surface with support


walls [29].

Flange

Figure 35: Thickness measurement for a multistage


o
formed pyramid; up to = 81 [61].

Figure 36: An example of using four stages to form a rapid


prototype of a reflector surface for an automobile [29].
direction of forming tool, from
outside edge to the centre
a
forming tool
x-y plane of flat sheet

unbroken perimeter from the edges of the part up to the


workplane. These surfaces support the part while it is
being formed, and help to ensure that the desired
geometry is obtained. The reflective lens shown in Figure
38 is an irregular shape that requires support surfaces. As
these new surfaces will be removed once the part has
been formed, their positioning and geometry is not critical,
though the maximum draw angles of the process still
apply. The support surfaces are shown as straight lines in
Figure 38. In some cases this is called lofting. Wall
angles of 40 to 55 degrees are recommended in the
support surfaces where possible. This eliminates the
chance of tears occurring in these sections during forming.
The completed perimeter of the workplane is the outline
that should be cut in the steel backing plate for the forming
backplate. An offset of 1 mm is helpful to account for any
minor misalignment of the forming rig during set-up on the
milling machine traverse table.
Once the CAD file has been modified, and embedded in
the sheet surface, the CNC toolpaths must be created for
the milling machine to follow. Any available CAM software
package can be used with equal success. Due to the
range of packages available and the rate at which new
software develops, only commands for a specific case will
be discussed here.
There are several choices for CNC control of the forming
tool as shown in Figure 39. Commands vary with software
package. Using of off-the-shelf CAM modules can be
problematic because usually they are optimized for milling
processes [34]. However the following strategies will show
how this can be overcome.

1st contour

Figure 37: The asymmetric incremental forming of a


reflective lens surface from sheet metal [29]. Green lines
are the embedded shape.

Figure 39: Suitable toolpaths for SPIF [34].

Figure 40: Twisting observed in cones with TPIF [100].


The first step for the CAM portion is to check the CAD file
visually for potential errors using a graphical interface, and
then the toolpaths are created. Toolpaths may be set for
either a roughing pass, or for a finishing pass. The
difference between the two is that with the roughing
process, commands are created to mill out all of the
material as if the part is being milled from a solid block,
while the finishing process only takes a fine, final cut
around the surface of the geometry. CAM software is not
designed for the incremental forming process, but by using
the finishing pass commands, it will generate the proper
CNC instructions when set to cut with an appropriately
sized ball end mill.
With the contour finishing pass toolpath selected, the
remaining necessary machining parameters may be
tailored to the forming process. These include feed rates,
step-down, roll over all surfaces, filtering, and transitioning.
Feed Rates: The feed rates for the forming process are
much higher than typical machining feeds. As the tool is
hemispherical, there is no concern about the amount of
material cut per tooth per revolution, a critical factor in
determining feed rates in conventional milling.
Step-down: This factor controls the z-axis increments. In
some cases an adjusted step-down for various wall angles
has been used to maintain a constant traverse distance
over the metal surface when manufacturing a range of
pyramids and cones with different wall angles [59, 60].
This is more difficult with more complicated geometry
containing different wall angles at different locations.
Usually a constant diameter forming tool, with a
standardized step-down is used. The step-down controls
the surface finish. Keeping the same step-down with steep
and shallow wall angles, can give a wavy surface on very
shallow geometries. To correct this, a shallow command
can be selected to automatically reduce the step-down
increments in these areas for a better finish.
Roll over all surfaces: This setting is very important in the
forming process. When a finish toolpath and a ball end
mill are used together, the computer automatically begins
the initial z-axis height at a level significantly below the
metal sheets surface. This height is determined as the
required z-axis position to contact the side of the ball with
the steep wall geometry. Unfortunately, the steep wall
geometry that the computer expects does not exist with
forming, and the tip of the tool makes contact with the
metal sheet long before the side of the ball contacts the
theoretical steep wall. The overly large deformation on the
first tracing of the parts perimeter can lead to a premature
failure of the metal sheet, so to avoid this, the roll over all
surfaces feature is selected. This will have different
command names, depending upon the software. This
feature adjusts the initial toolpath so that the z-axis height
begins with the tip of the ball in contact with the flat sheet
metal blank.
Filtering: The filtering command simply replaces the fine
splines created by a software package with a series of
arcs. Doing this significantly reduces the length of the
CNC instruction file, and can allow very large metal
forming toolpaths to fit on a standard computer floppy disk.

This method of data transmission is far superior to the


alternative RS-232 serial cable compatible with the milling
machines, and permits the use of a faster feed rate than
would be possible with the serial cable. Care must be
taken when using the filtering command to prevent the
size of the replacement arcs from becoming coarse
enough to affect the surface finish of the final part.
Transitioning: The machining technique used in some
cases is known as 2 D. The CNC code can be broken
up into a z-axis traverse and then a simultaneous x and yaxis traverse. Problems arise when the transition is made
between various z-axis levels unless care is taken to avoid
them. The standard z-axis transition is a straight x-y axis
traverse, then a plunge z-axis traverse. It is simple to
program and does not affect the outcome of conventional
machining processes. Unfortunately, an undesirable trait
can occur when this transition is used in metal forming.
When the z-axis plunge is performed at the same relative
x-y co-ordinates for each pass of the forming process, the
sheet metal can work harden to a point where this
transition results in a visible flaw along the finished part, or
even worse, it may lead to a fracture along this band of
overly worked material. To correct this, a ramp transition
can be selected, which breaks the large x-y axis and zaxis traverse up into a series of smaller traverses that
occur incrementally along a specific length of the forming
path. Subsequent transitions begin where the pervious
one ended along the forming path, so in this fashion, the
transitions are spaced equally around the perimeter of the
part.
Twisting of the sheet metal part, after forming, has been
observed. This is due to the forming process and the path
chosen as shown by Jadhav [100]; see Figure 40. The
solution to this is to choose appropriate paths as indicated
in Figure 39.
Finally, the difference in toolpath planning between SPIF
and TPIF is the tool moves from the outside inwards in
SPIF and in TPIF it moves from the centre to the outside.
Otherwise the planning for step increment, draw angle and
tool size are generally the same.
The end result is shown in Figure 37 where the inside of a
cone with forming tool can be seen. The sequence of
steps followed by the tool are also shown. The steps in
Figure 17 correspond to those shown in Figure 37. For the
case shown, the z steps are labeled 5, 9, 13, etc. and the
x steps are labeled 4, 8, 12, etc. The hypotenuse of x
and z is at the drawing angle .
Spiral Toolpath
A spiral trajectory is another way of controlling the surface
roughness and also eliminating the mark left on the
product surface by the previous method. The difference is
that the path is given a pitch. Figure 38 shows the toolpath
used by Filice [32]. The details of producing a toolpath, as
discussed in the previous section, also apply to a spiral
toolpath. However, CAM programs do not generally permit
the use of spiral trajectories. Filice [32] solved the problem
by using a simplified approach and did not utilize any
particular CAM program. Instead they developed a spiral
path inside EXCEL, thereby obtaining a large set of points
that carefully defined the trajectory and gave it to the CNC
milling machine. A couple of consecutive points were
linked through linear interpolation. However, for complex,
industrial parts it is necessary to use CAM programs that
do not generally permit the use of spiral trajectories.
Lubrication is one additional factor that must be taken into
account. In all cases cited some form of lubrication is
used. Bramley [103] conducted a set of tests to determine
how much lubrication changes the surface roughness. The

upshot of the tests was the type of lubricant did not appear
to be a factor, but lubrication is necessary to obtain a
smooth surface. Table 5 shows results from some tests.
Shape

lubrication

Spindle
rpm
1000

Ra
mm
9.42

Rt
mm
74.73

sphere

none

pyramid

grease

1000

1.24

14.43

sphere

oil

20

0.538

5.09

pyramid

grease

20

0.564

5.43

Table 5: The effect of speed and lubrication on surface


roughness [103].

Figure 41: Forming of a cone, showing the forming tool


inside the cone and the outside surface of the cone. The
steps shown are in sequential order and are for
incremental, unidirectional steps [104].

Figure 42: The spiral toolpath used by Filice [32].


4.9 Process Mechanics and FEA
The knowledge of process mechanics is not only important
with respect to the process limits, but the final properties of
parts made by the new process are also of major
importance with respect to details such as service life. As
with most processes, a model can help explain/understand
certain things such as the state of stress, strains and
forming limits. In principle it is possible to model the AISF
process, though at the moment experiments will usually be
faster in forming small parts. There is a trade-off for large,
complicated parts; it would not be economical to run many
tests to verify if a shape can be made.

AISF is characterized by a cyclic, local plastic deformation


of the sheet metal and has several peculiarities. It has
been investigated mainly by simplified analytical
deformation models and by full scale finite element
analysis (FEA). The most prominent analytical model is
the sine law [86], which is described in Figure 2.
For single-step forming strategies that adhere to the zlevel type toolpath generation and for low draw angles ,
the sine law is in very good agreement with measured
sheet thickness values, thus providing (in combination with
the FLD for a given material) a quick estimation of the
feasibility of designs.
Kim [94] used the sine law to calculate the thickness
distribution of a demonstrator part by means of a triangular
mesh. In addition Iseki [95] proposed a plane strain
deformation model, which assumes that the sheet metal in
contact with the tool stretches uniformly.
Unfortunately, analytical models are limited to the
approximate prediction of strains. For further studies, the
finite element method has been used. Micari [59], Hirt [68],
Bambach [69], Ambrogio [91], He [107, 108] and Henrard
[109] have been active in developing FEA models. Explicit
FEA models developed by Hirt [68] give good results with
respect to thinning and strains, however they are less
accurate when predicting geometry and springback.
Combining aspects such as alternating contact loci, large
strain plasticity, high local field gradients and complex tool
kinematics, make an FEA of the process a challenging
task. Since asymmetric incremental sheet forming is a
slow process with a time scale of minutes or hours,
simulations with standard FEM systems and workstations
can be expected to take much longer than the actual
process. Two basic approaches are possible for the
formulation and numerical solution of the governing
equations of metal forming: implicit and explicit methods.
There are advantages/disadvantages to either method.
For instance, with explicit methods, the analysis of the
springback phase often has a large computation time while
with implicit codes, the springback phase can usually be
performed in a few increments. Therefore, explicit finite
element codes for sheet forming often use an implicit code
to perform the springback analysis.
Bambach et al. [93] did a benchmark analysis of AISF with
implicit and explicit finite element models for a symmetric
cup. The FEA used both ABAQUS/Standard and
ABAQUS/Explicit. In both cases, the sheet is meshed with
2304 shell elements with nine through-the-thickness
integration points. DC04 sheet is modeled as an elastoplastic material with isotropic hardening. The friction is
assumed to be 0.05 between tool and sheet, and 0.15
between sheet and backing plate, which supports the part
during forming. A fictitious time of 5.31 s is used for the
process. This corresponds to the duration of the process in
reality at full tool speed, i.e. if acceleration and
deceleration are neglected. The result for the implicit FEA
with respect to sheet thinning is given in Figure 43.
For evaluation of the finite element calculations, the sheet
thickness and the geometry are compared to experimental
data along a radial section in positive x-direction for both
the implicit and explicit FE models (Figure 44). To check
the accuracy of the calculated geometry, the maximum
normal distance dmax and the average normal distance dav
between the experimental data and the FE results are
determined. The prediction of sheet thickness is judged by
means of the maximum deviation dth,max between the
experimental data and the FEA.
The implicit analysis provided very good results when
compared to experimental data, however the time is large

Thinning [%]
60
50
40
30
20
10

Figure 43: Sheet thinning for a symmetric cup [93].

5
experiment
explicit FEA
implicit FEA

0
z [mm]

-5
-10
-15
-20
-25

10

20

30
40
x [mm]

50

60

70

Fig. 44: Comparison of calculated and measured part


geometry [93].
FEA

Dt [s]

dmax
mm

dav dth,max CPU


mm [%] time [h]

implicit

0.004

1.09 0.59

11.7

7.32

explicit

0.0001 1.82 1.19

15.6

0.58

Table 6: Comparison of implicit and explicit FEA with


experimental data [92].
sheet
thickness
[mm]
1.50
1.44
1.38
1.32
1.27
1.21
1.15
1.09
1.03
0.98
0.92
0.86
0.80

Figure 45: Calculated thickness distribution for the


pyramidal frustum [93]. The tool outline is shown.

3
2
1
0
-1
-2
-3

pl,

even for a small benchmark part and short toolpath, so


that an implicit analysis cannot be used at present for
toolpath optimisation. In contrast, the explicit analysis was
performed in about 35 min on a 2.6 GHz Pentium IV single
processor machine. Therefore, although the implicit FEA
provides better results concerning geometry and thinning,
the explicit FEA offers a reasonable accuracy with
computation times that allow for a computer-based
toolpath optimization. See table 6 for results.
As reported earlier, the explicit method is generally
considered to be ill-suited for the prediction of springback.
Three types of springback can be defined in AISF:
a continuous local springback that occurs on every
displacement of the tool,
a global springback that occurs after the final
unloading and dismounting from the clamps,
a global springback after trimming (if done).
If the last two types of springback are to be modeled, this
can be performed by means of an implicit solution
procedure that is employed after an explicit analysis step
to analyze the forming process. However, the first type of
springback as described in [69] cannot be modeled by
means of a combination of explicit and implicit solution
procedures as this would require a continuous interchange
between the two codes. Since a purely implicit approach is
not economically feasible for larger models, a viscous
damping can be used [93] to improve the solution behavior
of the explicit method for the prediction of springback for
the single point forming of a pyramid frustum.
The outcome of the FE analysis is given in Figure 45.
Modeling the complete process (e.g. for a cone or a
pyramid) using continuum elements is uneconomical.
However, it is necessary to use continuum elements when
the local stresses under the tool are of interest. For
instance, shell elements based on Kirchhoff or MindlinReissner theory do not allow for a representation of the full
3D stress field. (Although they provide good results in the
prediction of geometry and thickness.)
Consequently, some researchers [69, 94] consider a strip
of material that is meshed with brick elements. The strip
can be thought of as a plane, side wall of a pyramidal
frustum.
The authors found that:
the plastic strain increases stepwise under the action of
the tool and
each increase in plastic strain is accompanied by
compressive stresses.
This can be seen in Figure 46 where a plot of the
equivalent plastic strain (blue curve) and the triaxiality
ratio (red curve) are given. Triaxiality is the ratio of the
hydrostatic to the deviatoric stress. Figure 46 also shows
that negative peaks in the course of the triaxiality coincide
with and increase in plastic strain. Thus, compressive
stresses are superimposed whenever the tool deforms the
reference elements.

0.0

0.1
0.2
0.3
normalised process time

0.4

Figure 46: Equivalent plastic strain and triaxiality ratio [68].


These compressive stresses are thought to be the reason
for the high forming limits observed experimentally. This
assumption has been studied further [69, 72] using the
Gurson-Tveergaard-Needleman (GTN) model to analyze
the damage evolution during forming. Because of the
stretching of the sheet, the damage evolution must be
assumed to be anisotropic. Consequently, the isotropic
GTN model cannot be used to quantitatively predict the
damage evolution of the sheet. Despite these restrictions,
the GTN model gives a good qualitative conformance to
experimental results in that it predicts the amount of
damage that occurs is increased when the size of the tools
are increased.

For some part designs, a multi-stage SPIF procedure must be


used for steep flanges [61, 83]. Starting with a shallow preform,
the wall angle is increased in several stages to yield a final wall
angle of 80-90; see Figure 35. Multi-pass forming creates
compressive stresses between the stages, and the sheet can
wrinkle or fold under the action of the tool, e.g. when the
increase in angle is too large between the stages. An initial
investigation of the prediction of wrinkling has been
successfully conducted [93], where an explicit code has been
used to compare two variants of the multistage forming of an
(almost) square box. The FEA is found to be in good qualitative
conformance to experimental results, indicating that the
material folds for one of the strategy variants, which can also be
observed in the corresponding experiments. Furthermore, He
et al [107] have used FEA to predict the forces acting on the
tool during forming of an AA 3003 truncated cone. The
simulated values overestimate the experimental ones by 30%.
This may be due to the assumption of isotropic hardening.
4.10 Process Accuracy
The importance of dimensional accuracy for sheet metal
oriented rapid prototyping in general, and incremental
forming processes specifically, is clearly illustrated by
Allwood et al. in [99]. In the market study presented in this
paper the process window for important process features
is matched with the requirements for a significant number
of potential part categories in demand by the market. The
major shortcoming today for the asymmetric incremental
forming methods can be situated at the achievable
accuracy level: while most applications would impose a
dimensional accuracy of less than 1 mm, and a significant
number of part categories would require 0.5mm as
tolerance level, the accuracy levels reported by
researchers typically exceed these limits.
Bramley reports an accuracy of 1.5 mm for symmetrical
and 2 mm for asymmetrical parts produced by SPIF
using a toolpath from a milling oriented CAM module [103].
Along the outer edge of the workpiece, the deviations from
the preset part geometry can be limited by the use of a
support plate (see Figures 2 and 43). Otherwise this zone
typically shows even larger deviations between the CAD
geometry and the achieved part dimensions [91, 103].
The use of a support plate cannot eliminate similar
problems due to strong, localized slope gradients at other
locations in the workpiece. Reprocessing after reversal of
the workpiece is suggested as a means to correct
significant, unwanted deformation along the edge of the
part in case no support plate is used [103, 105].
For TPIF of mild steel parts Hirt et al. [68] report deviations
of order of magnitude 2 to 3mm, depending on the part
geometry, when using a straight forward toolpath as
generated by standard CAM contouring algorithms.
Besides the obvious machine tool inaccuracies, causes of
deviation from the theoretical workpiece shape, that
correspond to the programmed toolpath are: springback in
the material, geometric distortion due to stress
propagation (see Figure 47, point marked with A) and
over-spinning effects that result in unwanted bulging [105].
When determining the process accuracy, the following
typical features of asymmetric incremental forming need to
be taken into account. The systematic thinning that
characterizes the process can influence accuracy, since
the dimensional errors induced by thinning are typically of
the same order of magnitude as the imposed tolerances.
This implies that accuracy should be determined for either
the convex or concave face of the workpiece. Depending
on the selected face, a toolpath should compensate for the
anticipated thinning.

In contrast to conventional press work, the non-flatness of


the blanks can be an important source of non-processrelated inaccuracy, since part coordinates are typically
determined with the unprocessed blank surface as a
datum plane.

Figure 47: Section view of the solar oven part, created


without backing plate (above, shown in red) and with a
backing plate (below, shown in blue). The CAD model is
given in black and the preshape used in a double pass
toolpath strategy that is explained later is in green [81].
(mm)
Average
Deviation

Test 1 Test 2

Absolute 1.327

Avg. Pos. Dev.

1.113

Test 3 Test 4

1.317

0.955

0.379

1.513

1.067

0.305

Avg. Neg. Dev.

-1.55

-0.58

-0.79

-0.47

Minimum

-4.47

-1.78

-2.75

-2.06

Maximum

4.137

5.398

4.223

0.980

Sigma ()

1.727

1.221

1.169

0.508

Table 7: Accuracy for different toolpath strategies [105].

Figure 48: Determination of virtual target part geometry


based on measurement feedback [68].
The influence of different process parameters and toolpath
strategies on part accuracy has been reported. Ambrogio
et al. [91] demonstrate the positive influence of smaller
tool diameters and pitch size on the geometric discrepancy
between the workpiece and the target geometry.
Duflou [105] compared different SPIF toolpath strategies
for a given geometry. They concluded that double pass
processing with an upward contour finishing pass provided
the best overall accuracy (see Table 7). All test parts were
AA 3003 with a tool with a diameter of 12.7 mm. The
spindle speed was set so the tool rolls over the material
and the feedrate was set to 2000 mm/min.
Some corrective measures based on different feedback
strategies have been proposed to improve part accuracy.
Hirt et al. [68] proposed an iterative method in which the
completed part is measured as a reverse engineering
exercise, providing data for an adjusted toolpath
generation. The proposed method is based on comparison
of the actually formed part geometry with the target model,
mirroring the measured points around the target geometry
with a scale factor, and using these

points to generate new, virtual target geometry. This virtual


part geometry forms the basis for the determination of an
improved toolpath. Using a scale factor of 0.7 was found
to provide optimal results for part made of DC04, 1.5 mm.
Ambrogio et al. [106] use an in-process measurement
system that allows the determination of deviation between
the anticipated intermediate part geometry and the actually
realized intermediate shape. Per layer (incremental
toolpath contour) the observed deviations are measured to
correct the toolpath geometry for the next contour.
The proposed system has been tested with a discrete point
contact measurement system, used interactively, thus
simulating the availability of real in-process measurement
equipment. The toolpath optimization algorithm has been
tested with pyramid part geometry. The author claims
significant accuracy improvements. No quantitative output
is however available to evaluate the achievable
dimensional accuracy.
5
EXAMPLES OF APPLICATIONS
The major advantage of asymmetric incremental forming is
it can be used to make asymmetric parts, quickly and
economically, without using expensive dies. Shapes used
to demonstrate the abilities of the process are shown in
Table 8. Some of the shapes illustrated have been used to
conduct springback experiments, and in determining the
maximum draw angle , others are just for demonstration
of process abilities.
The asymmetric single and two point incremental forming
processes are still in their infancy. Much research work
remains to be done and to do this appropriate shapes are
needed to develop: FLDs, springback models and models
to test accuracy. Standard shapes are used to determine
what the draw angles should be for different materials and
thickness. Standard shapes are needed to compare
experimental results and for testing the process for
maximum speeds of deformation.
When testing materials for the maximum draw angle a
truncated cone as shown in Figure 15 was agreed upon
[59]. This shape has also been used by others [31, 32, 33,
59, 64, 66]. A pyramid shape has also been used on many
occasions to demonstrate the asymmetric abilities of the
new process plus it has become useful in determining the
springback that occurs for different material thickness and
for conducting pre and post process material tests [59, 64,
65, 67, 77]. Jadhav [100] has also used the pyramid to
conduct studies of twisting in a shape once forming is
completed.
There is not agreement on which shape or shapes should
be used to develop FLDs for the new processes. Filice
[32] and Hirt [78] have used a pyramid. Young is a
proponent of using several shapes each of which has
elements contained in most part shapes [62].
5.1 Rapid Prototype Examples
Making Rapid Prototypes, with sheet metal as the base
material, giving a part that can be used directly in the
function for which it is intended, is one of the major
advantages of using this Incremental Forming process.
The parts shown in table 9 are for designs that were made
as Rapid Prototypes for the automotive industry. They are
for the reflective surface of prototype headlights, for the
first two cases. The third case is for a heat/noise shield,
which is used over exhaust manifolds.
Table 10 shows examples of rapid prototypes made for
non-automotive applications. The first example is of a solar

Cross

Hexagon

5 lobe
shape

Faceted
cone

Multi-shaped surface

V shaped
tub

Cone

Dome

Hyperbola

Truncated pyramid

Table 8: Shapes used to demonstrate the viability of the


process and for experiments.
oven cavity for use in developing country applications. The
last two are for the same manufacturer of custom
motorbikes; the first part is for a motorbike seat and the
second is part of a gas tank.
5.2 Custom manufacture of a solar oven
The SPIF process has made it possible to manufacture an
aluminum solar oven cavity, economically without dies.
The ability to make a sheet metal cavity, inexpensively,
has allowed designers of the solar oven to redesign other
parts of a product, thereby reducing the cost and labour in
making a solar oven. A solar oven has been designed for
export to developing countries. Originally the oven cavity
was made from fibreglass, and painted black. The major
drawback, was the fibreglass wall was 7 mm thick, heavy,
time consuming to build, and labour intensive. The
possibility of using dies to form sheet metal into solar
cooking cavities was investigated and found to be very
expensive. This lead to considering SPIF of sheet metal as
a method of manufacture. Figure 49 shows how the
assembled solar oven works. The red shape is the part
which is formed by SPIF. A three dimensional model was
made in a Unigraphics environment and Figure 50 shows
the model and increment details. The total depth of the
model is 74 mm. The downward step for each pass was
set up differently for the two sections shown; z = 0.4 mm
from A to B, and z =0.3 mm from B to C.
The method of programming included a z-level profile
following part contour, which is available in Unigraphics,
was used.
Details of the process set-up in Unigraphics are:
Maximum drawing angle: set to 65o, everywhere.
Forming tool: 25.4 mm diameter, highly polished.
Forming speed (feedrate): 1125 mm/min.
Additional note: the forming tool rotates freely, and
is not matched to the machine feedrate.
1.3 mm thick AA 8008-0 Aluminum.
The final, successful result is shown in table 10. Although
the customer was unconcerned by springback, the crosssection was measured to find out how much springback
actually occurred; see Figures 47 and 51.

z
z

IGES file
x

z = 0.4 mm

y
74 mm

2002 vehicle
headlight

2002 vehicle
headlight

Automotive
heat/noise
shield

A
B

z = 0.3 mm

Figure 50. Illustration of the CAD and toolpath setup [ 36].

Finished part

inside

inside
outside

aluminized
Table 9: Rapid prototypes for automotive industries [29].

IGES file

Solar oven
cavity.

Motorbike
seat

Motorbike
gas tank

Finished part
inside

outside
surface

outside

gl a

ss
lid

Table 10: Non-automotive rapid prototype examples [104].

cooking
cavity

side view - oven


Figure 49: Assembled solar oven operation. The cooking
cavity is shown in red [36].

Figure 51: The x-z profile of CAD file and formed product.
The outside of the solar oven is shown as an inset. The
springback can be observed [36].
5.3 A Medical Application of SPIF
The medical industry can benefit greatly by using the SPIF
process to either produce Rapid Prototypes or one-off
parts. The following is an application of asymmetric
incremental forming of a customized medical product, i.e.
an ankle support. The steps to go from the request for a
part to the final manufacture are shown in Figure 52.
The application addressed here represents a Reverse
Engineering (RE) application; it permits the reproduction of
high accuracy models called from complex geometries. By
employing RE, it is possible to create lifelike 3D models
representing many different objects including parts of a
human body.
In this application, SPIF was used to produce a part that is
used as an ankle support.
The process includes the following steps:
first, the three-dimensional scanning of a human ankle
with a laser (non contact inspection) is carried out in
order to obtain the morphological and dimensional
information of the object;
next, once the data is available in a digital format, a
cloud of points is built, from which is possible to
develop the surfaces that individualize the shape ankle;
finally, through the application of CAD/CAM Systems, an
ISO part program is generated. This is handed over to
the numerical control machine, which carries out the
Incremental Forming operation.
Laser scanning is used because it has a large number of
advantages, such as: low operation costs and easy
acquisition; small dimensions; real time results.
The application addressed here was carried out with a
system that works according to triangulation principles.
This permits the measurement of small objects, with a
scan in six tenths of a second. The system permits
recording of the object including surface features such as
colour and luminosity.
The next step is surface building or editing with software.
This uses multiple observations during scanning, and since
the object investigated is observed from many different
angles markers are used on the surface. The software
develops the parametric surfaces, which are used to obtain
the solid model of the ankle.

CAD Model

Real Model

Laser Scanner

Virtual Model

Production

Request for an
Ankle Support

CAD/CAM
interface

shape embedding

`````
Figure 52: The manufacture of an ankle support, from request, to scanning of a live subject (reverse engineering), to setting up
a solid model and CAD drawing, to embedding the shape for toolpath planning, to creating a toolpath for manufacture, to
checking for accuracy [59].
Strength
coefficient

Hardening
exponent

Anisotropy

Tensile
Strength

Elongation
%

K = 545
MPa

n = 0.27

= 2.01

R = 290
MPa

A% = 50

Table 11: Properties of deep drawing quality steel [59].


The ankle support was shaped starting from a surface
slightly greater than the scanner recording, in order: a) to
obtain a clearance between the instep and collar; b) to
create space for an inside coating.
At the end of this step a real curve grate, shape, is
obtained. This is imported into a CAD/CAM system to
develop the CAM step. The toolpath is compiled and
generated in order to carry out the SPIF operation on a
CNC milling machine. The ankle support is divided into two
symmetrical parts, splitting the solid model and creating
two separate part programs. The position of the model in
the space is determined, with the specific aim of avoiding
surface inclinations that go beyond the incremental forming
limits in terms of the maximum drawing angle, max. Once
the right positioning is found, the part program, is
generated.
A CNC, three-axis milling machine is used for the forming
process. The work volume is 560X410X510. The program
is loaded into the machine memory and the part is formed
incrementally, making two matching, half ankle supports.
The sheet material is DDQ (Deep Drawing Quality) steel, 1
mm thick. Its properties are shown in Table 11. The tool
feed and rotating speed are 1000 mm/min, and 500
rev/min. The depth step, z, is 0.5 mm, while the tool
diameter is 11 mm. The forming time is half an hour.
Figure 52, shows a completed DDQ steel half ankle.
Process accuracy has been checked, by repeating the
foregoing procedure, with a part formed by Incremental
Forming. A laser scan of the part gives a new set of cloud
points that can compared with the original surface part
program. The results obtained show a maximum deviation
error of 0.5 mm.
6
THE BUSINESS CASE FOR AISF
As a relatively new process, AISF is not currently in regular
commercial use, but all developers of the process have
suggested potential application areas, and in some cases
have made test products to demonstrate feasibility [31, 32,
33, 60]. Table 12 lists the range of applications that have
been considered by developers which spans small to large
products, and products with more or less complexity.

The most significant benefit from AISF arises in low


volume production where existing processes require
specialized tooling. The cost and lead time of tooling
manufacture leads to very high component costs for small
runs of new designs, and all developers of AISF have
identified the attraction of AISF for prototyping or low
volume manufacture. In order to provide more precise
guidance on identification of candidate applications, two
economic models have been developed to explore the
trade off between fixed costs (particularly those associated
with tooling) and direct costs. Hirt and Ames [97] consider
the direct costs as comprising materials, labour and a
machine overhead rate, with tooling as the fixed cost.
Micari [59] uses more detail, with direct costs comprising
materials, labour, power, two overhead rates and some
batch setup costs, and fixed costs comprising tooling, NC
programming and machine setup. Both authors find that
AISF is likely to be attractive for production where new
tooling would be required and total batch sizes are less
than around 600.
Such analysis gives a means to support decisions on
investment in AISF. However, calculations of utility based
on this form of cost model have some limits:
the precise cost data required to apply the economic
evaluation is often unavailable, except in very large
companies. Even when it exists, it is unlikely to
reflect true cost: labour costs may be higher or lower
than a set rate depending on the capacity and
utilisation rate at the factory; machine overhead rates
are always estimates, as they can only be calculated
accurately in arrears, once utilisation rates are known.
the value of the ISF process is unlikely to be solely as
a direct replacement for an existing process. Existing
product designs and their associated process choices
have been selected based on available production
technology. The value of a novel technology such as
ISF is likely to be achieved when product designs
change to reflect the benefits and capability of the
process.
cost is not the only determinant of process selection.
Constraints on tolerance, surface finish, sheet
thickness and residual stress levels are equally
important in determining process choice.
These concerns are well known in more general analysis
of investment in advanced manufacturing technology. They
are discussed at more length by Ordoobadi and Mulvaney
[98] who claim that purely economic assessments are
generally pessimistic as they fail to capture the system
wide benefits of the new technology. In an attempt to

provide a broader assessment of the value of AISF,


Allwood, King and Duflou [98] have proposed an analysis
of AISF based on product segmentation. They gathered a
broad data set from 15 UK companies manufacturing
products from sheet metal, considering a total of 28
products. The data was largely related to the product (or
component or application) rather than the existing
processes used to make it, and included Figures for direct
costs and lead times, tooling costs and lead times, product
design and quality specifications. The data set was plotted
on a series of segmentation charts which could also be
used to display a process window for the capability of
AISF. Analysis of the charts suggests that AISF is
generally applicable for high value low volume products,
and that the volume for which AISF is attractive increases
with tooling costs. Generally, for batch sizes where AISF
was economically attractive, it also showed a lead time
advantage over conventional production. The product
segmentation approach also allows validation of process
capability. Of the products examined in the study, only
one a one-off cowling panel for a historic aeroplane
could be made by existing AISF techniques. However, if
the AISF process window could be expanded by improving
tolerance to around 0.3 mm, with feature definitions to 1
mm radius and allowing material thickness up to 3 mm, the
process would be attractive for more than half of the
applications considered.
As AISF technology matures, and receives wider
exposure, it can be considered by designers seeking the
advantage of its unique features, and this will generate
new applications. Furthermore, the avoidance of any
significant tooling costs in AISF, and the relatively low
asset value of the equipment suggests that AISF would be
attractive for distributed manufacturing. This opens a
further set of novel applications where supply chains that
currently depend on stock-holding and mass distribution
(such as the car body repairs and after-sales market) could
be reorganised via a network of low cost AISF processes
with distribution of data rather than material. This has
obvious benefits both in reducing the economic and
environmental burdens of distribution, and in offering
increased customisation without additional cost.
Clearly table 12 relates to actual applications however
the intention of table 12 is to include as the range of
applications being considered is wider than the set of
applications that have actually been tested.
7 RESEARCH OPPORTUNITIES AND CHALLENGES
Many potential applications have been listed in the
foregoing section. Nowadays, aspects such as aesthetics
and ergonomic quality have on increasing importance. This
gives rise to new products, which have always originated
in a designers mind but have not always been economical
to implement. Now, with AISF a design concept for a
shape can be implemented, immediately and economically
to give either a rapid prototype or a production part.
Potential applications in markets that include insurance
claims can be explored. For instance, sheet metal parts
can be stored electronically and then made as required for
automotive replacements, or parts can be reverse
engineered and then made on a three-axis CNC mill.
The architectural industry also has great promise for this
application. Custom made sheet metal parts can be made
with individualized patterns.
Although AISF is now viable, many challenges remain. A
set of useful guidelines now exist for the engineering
designer when using this new process. However, these
need to be expanded. The process is now established as a

useful tool for producing rapid prototypes which can be


used in situ. One challenge is the issue of springback
which keeps accuracy at 1.5 mm.
Application areas:
Automotive body panels (prototype, low-volume eg
motorsport, and after-sales)
Other automotive sheet metal parts structural, or
non-aesthetic
Architectural bespoke formwork, decorative
panels
Customized white goods
Reflectors and casings for lighting
Dental bespoke dental crowns
Housings and fairings for aerospace
Ship hull plates
Table 12: Potential applications areas for AISF [99].
Models that will allow a designer to achieve greater
accuracy are needed. Another need is the development of
Forming Limit Diagrams for different combinations of
materials and sheet thickness. Both the Design Engineer
and Manufacturing Engineer need these to determine if the
process is applicable for a design. The challenge is to
develop Forming Limit Diagrams, which can be used for
many combinations, including: type of material, sheet
thickness, forming tools and forming speeds. In addition,
research work shows there is a relationship between
maximum draw angle, max, and the major strain, max.
Another challenge is to develop a comprehensive system,
which gives an overview including shape prediction, FLDs
and max and compensates for springback.
Traditional stamping processes will still be used for mass
produced parts, for economic reasons. A challenge is to
make asymmetric incremental forming more viable for
mass production. With higher speeds and improvements in
accuracy this new process can be an alternative to
stamping especially with its flexibility. Modelling and on line
control are challenges along with the development of
algorithms which will compensate for springback.
There are many areas of manufacturing in which
asymmetric incremental sheet forming can be applied. One
example is the Biomedical area for which an example has
been given. For Biomedical applications it will be a
challenge to find ways of manufacturing the part while
meeting the clean room requirements of legal jurisdictions,
as discussed by McAllister and Jeswiet [102]. In the end,
all challenges are limited only to the imagination of the
Design Engineer and the Manufacturing Engineer.
8 SUMMARY
Asymmetric incremental sheet forming, with a single point
doing the forming, is a viable process for making
complicated shapes from sheet metal. It makes the
forming of sheet metal a flexible operation and is an easy
operation for all facilities having access to a three axis
CNC mill. There are two variations to the process, one
with a single point doing the forming and the other with two
points where one point is either a partial or full die. The
process has tremendous potential and there are many
future possibilities where it can be used.

Guidelines have been established for designers and


manufacturers who wish to form prototype sheet metal
parts or run a low volume production runs.
The parameters for which guidelines have been
established are: thickness of the sheet metal relative to
the maximum wall angle in a part (draw angle, ), the size
of the incremental step down, z, the speed of deformation
and size of the forming tool.
Design and manufacturing guidelines
Formability:
Formability increases with higher sheet-to-workpiece
relative velocities, with a trade-off of higher surface
roughness.
Formability decreases with thinner sheet.
Smaller tool size gives increased formability.
Anisotropy has an influence upon formability, with
greater formability being achieved with smaller diameter
tools in the transverse direction.
Sheet formability decreases with increasing increment
step size, z.
The effect of increment step size, z:
Large increment steps, z, give a higher roughness.
The increment step size, z, can influence not only the
surface roughness but also cause an orange peel effect.
The size of the orange peel effect can be influenced by
the incremental step size, x and y, and the draw
angle.
The effect of the draw angle :
There is a limitation on the maximum draw angle that
can be formed in one pass.
With increasing draw angle , the sheet thickness
reaches a minimum value where fracture occurs as a
consequence.
There is a strong dependence of the deformed sheet
thickness on the draw angle , which can lead to
inhomogeneous thickness distributions in the final part.
Knowing the parameter max for a material at a specific
thickness, a designer can take the first step in deciding if
a sheet metal part can be made in one pass without
tearing, or if a two pass or multiple pass sequence
should be used.
Plasticity:
Twisting can occur if forming is unidirectional.
Plastic strain increases in steps, with tool increments.
Each increase in plastic strain is accompanied by
compressive stresses. Cyclic loading can produce
residual stresses (due to repeated bending in
combination with the Bauschinger effect).
Springback is not any worse in this process compared to
other sheet metal forming operations. Three types of
springback can be defined in asymmetric incremental
sheet forming:
A continuous local springback that takes places on every
displacement of the forming tool,
A global springback that occurs after the final unloading
and dismounting from the clamps,
A global springback after trimming (if done).
Forces:
Peak forces can be observed in the area where failure
occurs at maximum draw angles, max,
Increasing the vertical step size z increases forces,
Larger tool diameters increase forming forces.
Increasing vertical step size z has a much lower impact
when compared to draw angle and tool size.

In addition, strategies need to be developed to increase


accuracy. This work is ongoing by members of CIRP and
will be reported in the future.
In summary, the flexible forming system with short lead
times envisioned by Schmoeckel [1] and the single point
forming system patented by Leszak [2] are now a reality
with the new ASIF processes that are now available.
Guidelines now exist for the manufacturing designer and
the new processes can be done by virtually any facility
having access to a three-axis CNC mill.
9
ACKNOWLEDGEMENTS
The authors are grateful to the UK EPSRC and the
University of Bath Innovative Manufacturing Research
Centre for the provision of funds to enable a workshop
meeting to be held at the University of Cambridge in
October 2004.
The authors have also been supported by research grants
in their individual countries and we wish to acknowledge
their support: the Institute for the Promotion of Innovation
by Science and Technology in Flanders (IWT); the Natural
Sciences and Engineering Research Council of Canada.
The information contained in this paper would not have
been possible without our graduate students: G. Ambrogio,
J. Ames, M. Bambach, E. Hagan, M. Ham, S. Jadhav, G.
Owen, A. Szekeres, and D. Young.
There are many others who have participated in seminars
that have been held in preparation for this paper, many
thanks to them.
10
[1]
[2]
[3]

[4]

[5]

[6]
[7]

[8]

[9]

REFERENCES
Schmoeckel, D. Developments in Automation,
Flexibilization and Control of Forming Machinery.
Annals of CIRP vol. 40/2/1992; 615.
Leszak, E. Patent US3342051A1, published 196709-19. Apparatus and Process for Incremental
Dieless Forming.
Mller, H., Enzmann, H., 1998, Potentials of Rapid
Prototyping Techniques for the Manufacture of
Prototype Sheet Metal Forming Tools, Proc.
European Conf. on Rapid Prototyping and
Manufacturing, Aachen, 1998, pp 337-350.
Mller, D. H., Mller, H., 2000, Experiences using
Rapid Prototyping Techniques to Manufacture Sheet
Metal Forming Tools, Proc. ISATA Conference,
Dublin, 2000, p 9.
Jrgensen, T. H., Moos, N., 2002, Applications of RP
Techniques for Sheet Metal Forming, State-of-theArt, 2002, RAPTIA Thematic Network report,
www.raptia.org .
Nakagawa, T., 1993, Recent Developments in Auto
Body Panel Forming Technology, Annals of the
CIRP, Vol. 42(2)1993, pp 313-317.
Franke, V., Greska, W., Geiger, M., 1994, Laminated
th
Tool System for Press Brakes, Proceedings of 26
International CIRP Seminar on Manufacturing
Systems - LANE 94, Meisenbach, 1994, pp. 883
892.
Himmer, T., Techel, A., Nowotny, S., Beyer, E.,
2002, Recent Developments in Metal Laminated
Tooling by Multiple Laser Processing, Proc. Internat.
Users Conf. on Rapid Prototyping, Rapid Tooling &
Rapid Manufacturing, Frankfurt/Main December,
2002, ISBN 3-8167-6227-1, p. B-5/3.
Hagan, E. and Jeswiet, J. A review of conventional
and modern single point sheet metal forming
methods. IMECHE part B, J. of Engineering
Manufacture, 2003 vol 217 No B2. Pp 213 - 225.

[10] Brown, J. Advanced machining technology


handbook, 1998 McGraw Hill.
[11] Anon., Process developments put metal forming in a
spin. Metallurgia, vol. 54, p 454, 482, Oct. 1987
[12] Sortais, H.C., Kobayashi, S., Thomsen, E.G.
Mechanics of Conventional Spinning. Trans.ASME,
series B, Journal of Engineering for Industry, vol. 85,
pp. 346-350, 1963.
[13] Reagan, J., Smith, E.
Metal Spinning, 1991,
(Lindsay Publications, Bradley, IL).
[14] Kobayashi, S., Hall, I.K., Thomsen, E.G. A Theory of
Shear Spinning of Cones. Trans.ASME, series B,
Journal of Engineering for Industry, vol. 83, pp. 485495, 1961.
[15] Kalpakcioglu, S.
On the Mechanics of Shear
Spinning
Trans.ASME, series B, Journal of
Engineering for Industry, vol. 83, pp. 125-130, 1961.
[16] Anon., Playback spinning doubles productivity.
Metallurgia, vol. 51, p. 386, Sept. 1984.
[17] Kohne, R. Better dimensional accuracy in metal
spinning. Strips Sheets Tubes. vol 3, pp. 31-34,
1986.
[18] Kegg, R.L. A New Test Method for Determination of
Spinnability of Metals
Trans.ASME, series B,
Journal of Engineering for Industry, vol. 83, pp. 119124, 1961.
[19] Kalpakcioglu, S. A Study of Shear Spinnnability of
Metals, Trans.ASME, series B, Journal of
Engineering for Industry, vol. 83, pp. 478-484, 1961.
[20] Avitzur, B., Yang, C.T. Analysis of Power Spinning of
Cones.
Trans. ASME, series B, Journal of
Engineering for Industry, vol. 82, pp. 231-245, 1960
[21] Ismail, N., Mahdavian, S. M., Hydrodynamic
lubrication in metal spinning. Tribology in
Manufacturing Processes, vol. 5, pp. 119-123, 1994.
[22] Kitazawa, K., Wakabayashi, A., Murata, K., Yaejima,
K.
Metal flow phenomena in computerized
numerically controlled incremental stretch-expanding
of aluminum sheets. J. of Japan Institute of Light
Metals, vol. 46, pp. 65-70, 1996.
[23] Kitazawa, K., Wakabayashi, A., Murata, K.,
Yakejima, K. Metal flow phenomena in computerized
numerically controlled incremental stretch-expanding
of aluminum sheets. J. of Japan Institute of Light
Metals, vol. 46, pp. 65-70, 1996.
[24] Kitazawa, K., Nakane, M. Hemi-ellipsoidal stretch
expanding of aluminum sheet by CNC incremental
forming process with two path method. J. of Japan
Institute of Light Metals, vol. 47, pp. 440-445, 1997.
[25] Kitazawa, K. Limit strains for CNC incremental
stretch-expanding of aluminum sheets. J. of Japan
Institute of Light Metals, vol. 47, pp. 145-150, 1997.
[26] Powell N. and Andrew, C. Incremental forming of
flanged sheet metal components without dedicated
dies. IMECHE part B, J. of Engineering Manufacture,
1992, vol. 206, pp 41 47.
[27] Matsubara, S. Incremental Backward Bulge Forming
of a Sheet Metal with a Hemispherical Tool. J. of the
JSTP, vol. 35, pp. 1311-1316, 1994.
[28] M. Bambach, G. Hirt, S. Junk. Modelling and
Experimental Evaluation of the Incremental CNC
Sheet Metal Forming Process, VII International
Conference on Computational Plasticity, COMPLAS
2003, Barcelona 7-10 April 2003.
[29] Jeswiet, J., Hagan, E. Rapid Proto-typing of a
Headlight with Sheet Metal, Proceedings of Shemet,
April 2001, pp 165-170.
[30] Kim, T.J., Yang, D.Y. Improvement of formability for
the incremental sheet metal forming process.

[31]

[32]
[33]

[34]

[35]

[36]

[37]
[38]

[39]
[40]

[41]
[42]

[43]

[44]
[45]
[46]

[47]
[48]
[49]

International Journal of Mechanical Sciences, vol.


42, pp. 1271-1286, 2001.
Leach, D., Green, A. J., Bramley, A. N., A new
incremental sheet forming process for small batch
th
and prototype parts. 9 International Conference on
Sheet Metal, Leuven, pp. 211-218, 2001.
Filice, L., Fantini, L., Micari, F. Analysis of Material
Formability in Incremental Forming, Annals of the
CIRP, vol. 51/1/2002: 199-202.
Jeswiet, J., Hagan, E. Rapid Prototyping Non-uniform
Shapes from Sheet Metal Using CNC Single Point
Incremental Forming. NAMRI/SME 2003 transactions,
vol. XXXI, pp 65 69.
Jadhav S., Goebel R., Homberg W., Kleiner M.:
Process optimization and control for incremental
sheet metal forming, Proceedings of the International
Deep Drawing Research Group Conference, IDDRG,
Pg. 165-171, Bled, Slovenia, May 2003.
Micari, F. and Ambrogio, G. A Common shape for
st
conducting
Incremental
Forming
Tests.
1
Incremental Forming Workshop, University of
Saarbrucken, 9 June 2004. On Cdrom.
Jeswiet, J., Duflou, J., Szekeres, A., Levebre, P.
Custom Manufacture of a Solar Cooker a case
study. Journal Advanced Materials Research, Vols. 68, May 2005, pp 487-492.
Hirt, G. Tools and Equipment used in Incremental
st
Incremental Forming Workshop,
Forming. 1
University of Saarbrucken, 9 June 2004. On Cdrom.
Amino, H., Lu, Y., Maki, T., Osawa, S., Fukuda, K.
Dieless NC Forming, Prototype of Automotive Service
Parts; Proceedings of the 2nd International
Conference on Rapid Prototyping and Manufacturing
(ICRPM), Beijing, 2002.
Aoyama, S., Amino, H., Lu, Y., Matsubara, S.
Apparatus for dieless forming plate materials,
Europisches Patent EP0970764, 2000.
Allwood, J. M., Houghton, N. E., Jackson, K. P. The
th
design of an Incremental Forming machine, 11
Conference on Sheet Metal, Erlangen, 2005; pp 471 478.
Kopp, R., Ball, H-W. Recent Developments in Shot
rd
Peen Forming. 3 International Conference on Shot
Peening, Garmisch-Partenkirchen, 1987; 297 - 308.
Meyer, R., Reccius, H. Shot Peen Forming of NC
machined parts with integrated stringers using large
balls. 3rd International Conference on Shot Peening,
Garmisch-Partenkirchen, 1987; 327 334.
Kopp, R., Wustefeld, F., Linnemann, W. High
Precision Shot Peen forming, 5th International
Conference on Shot Peen Forming. Oxford 1993; 127
138.
Kopp, R., Schulz, J. Flexible Sheet Forming
Technology by Double-sided Simultaneous Shot
Peen Forming. Annals of CIRP Vol. 51/1/2002.
Iseki, H.: Flexible and Incremental Bulging of Sheet
Metal Using High-Speed Water Jet, JSME
International Journal, Series C, Vol. 4, 2001:486-493.
Jurisevic, B., Heiniger, K. C., Kuzman, K. Junkar, M.
Incremental Sheet Metal Forming with a High Speed
Water Jet. Proceedings of the International Deep
Drawing Research Group conference, May 11, 2003;
139 148.
Geiger, M. and Vollertsen, F. Mechanisms of Laser
Forming. Annals of CIRP Vol 42/1/1993; 301 304.
M. Geiger. Synergy of Laser Material Processing and
Metal Forming. Annals of CIRP Vol. 43/2/1994; 563
570.
Kim, J. and Na, S.J. Development of irradiation
strategies for free curve laser forming. Optics and

[50]

[51]

[52]

[53]
[54]

[55]

[56]

[57]
[58]
[59]
[60]

[61]

[62]

[63]

[64]

[65]

Laser technology vol. 35, issue 8, Nov. 2003, Pages


605-611.
Hennige T., Holzer S., Vollersten F., Geiger M. On
the working accuracy of laser bending, Journal of
Materials Processing Technology (1997) vol.71 422432.
Otsu M., Fujii M., Osakada K.: Controlled laser
forming of sheet metal with shape measurement and
using database, Proc. of Metal Forming 2000; 433438.
Male, A. T., Li, P. J., Chen, Y.W., Zhang, Y. M.
Flexible forming of sheet metal using plasma arc.
Journal of Materials Processing Technology, vol 115,
2001; 61- 64
Weck, M. Werkzeugmaschinen-Fertigungssysteme
Band 2 Konstruktion und Berechnung; 5. verbesserte
Auflage, VDI-Verlag, Dsseldorf, 1994.
Westkmper, E., Schaaf, W., Schfer, T.
RoboshapingFlexible
Inkrementelle
Blechumformung
mit
Industrierobotern;
Werkstattstechnik online, Jahrgang 93, 2003, H .
Schafer, T, Schraft, R.D. Incremental sheet forming
by industrial robots using a hammering tool. AFPR,
Association Francais de Prototypage Rapid, 10th
European Forum on Rapid Prototyping, Sept 14,
2004.
Meier, H., Dewald, O., Zhang, J. Development of a
Robot-Based Sheet Metal Forming Process. In: steel
research, Issue 1/2005, Dusseldorf: Verlag
Stahleise.
Stewart, D. A Platform with Six Degrees of Freedom,
UK Institution of Mechanical Engineers Proceedings
1965-66, Vol 180, Pt 1, No 15.
Iseki, H.; Kato, K.; Kumon, H.; Ozaki, K.: Flexible and
Incremental Sheet Metal Bulging Using a Few
Spherical Rollers; TR. JSME, 59(565-C), 2849-2854.
Micari, F. Single Point Incremental Forming: recent
results. Seminar on Incremental Forming, 22
October, 2004, Cambridge University. CdRom.
Hirt, G. Junk, S., Witulski, N. Incremental Sheet
Forming: Quality Evaluation and Process Simulation.
th
7 ICTP International Conference on Technology of
Plasticity,
October
27-November
1,
2002,
Yokohama, Japan, paper no. 343.
Hirt, G., Junk, S., Bambach, M., Chouvalova, I.
Process Limits and Material Behaviour in
Incremental Sheet Forming with CNC-Tools,
THERMEC 2003, International Conference on
Processing & Manufacturing of Advanced Materials Processing, Fabrication, Properties, Applications,
July 7-11, 2003, Legans, Madrid, Spain.
Young, D., and Jeswiet, J. Forming Limit Diagrams
for Single Point Incremental Forming of Aluminum
Sheet. IMECHE part B, J. of Engineering
Manufacture, vol. 219 part B 2005. pp 1 6.
Hagan, E., and Jeswiet, J. Analysis of surface
roughness for parts formed by CNC incremental
forming. IMECHE part B, J. of Engineering
Manufacture. Vol. 218 No. B10, 2004. pp 1307
1312.
Hirt, G., Junk, S., Bambach, M., Chouvalova, I.,
Ames, J. Flexible CNC Incremental Sheet Forming:
Process Evaluation and Simulation. VI Conferncia
Nacional de Conformao de Chapas, 15.10 2003,
Porto Alegre/RS, Brasilien; ed. Lirio Schaeffer;
Brasul Ltda., 2003, pp.30-38.
Jeswiet, J. Incremental Single point Forming with a
Tool Post. Proceedings of Shemet 2001, April 2,
2001, K.U. Leuven; pp 37 42.

[66] Jeswiet, J., Hagan, E., Szekeres, A. Forming


Parameters for Incremental Forming of Aluminum
Sheet Metal. IMECHE part B, J. of Engineering
Manufacture. 2002 Vol 216, pp 1367 1371.
[67] Fratini, L., Ambrogio, G., Di Lorenzo, R., Filice, L.,
Micari, F. The Influence of mechanical properties of
the sheet material on formability in single point
incremental forming. Annals of CIRP vol 53/1/2004; p
207.
[68] Hirt, G., Ames, J., Bambach, M., Kopp, R. Forming
Strategies and Process Modelling for CNC
incremental Sheet Forming. Annals of CIRP vol
53/1/2004; p 203.
[69] Bambach, M., Hirt, G., Junk, S. Modelling and
Experimental Evaluation of the Incremental CNC
th
International
Sheet Metal Forming Process. 7
Conference on Computational Plasticity, 2003,
Barcelona; p 1.
[70] Kim, Y. H and Park, J. J. Effect of process
parameters on formability in incremental forming of
sheet metal. Journal of Materials and Processing
technology; vol. 130-131, 2002, pp. 42-46.
[71] Voce, E. J. Inst. Met. 1948 74:537.
[72] Lievers, W.B., Pilkey, A.K., Lloyd, D.J. Using
Incremental Forming to Calibrate a Void Nucleation
Model for Automotive Aluminum Sheet Alloys. Acta
Materiala 52 (2004) 3001-3007.
[73] Johnson, W. and Mellor, P.B. Plasticity for
Mechanical Engineers. 1962 van Nostrand
Reinhold Co.
[74] Hosford, W.F. and Caddell, R.M. Metal Forming.
1983 Prentice-Hall.
[75] Hecker, S.S. and Stout, M.G., Strain Hardening of
Heavily Cold Worked Metals, 1982 ASM Materials
Science seminar, ASM Metals Park Ohio.
[76] Amino, H., Lu, Y., Ozawa, S., Fukuda, K., Maki, T.
Dieless NC Forming of Automotive Service Panels,
Advanced Technology of Plasticity (2002), Vol.2.
[77] Hagan, E., Jeswiet, J. Effect of Wall Angle on Al 3003
strain hardening for parts formed by CNC incremental
forming. IMECHE part B, J. of Engineering
Manufacture; 2003 vol 217 No B11; 1571 1581.
[78] Hirt, G., Junk, S., Chouvalova, I. Herstellung von
Prototypen und Kleinserien komplexer Bauteile mit
inkrementeller Blechumformung, 9. Schsische
Fachtagung Umformtechnik, Dresden, 2002.
[79] Junk, S., Hirt, G., Chouvalova, I. Forming Strategies
and Tools in Incremental Sheet Forming, Proceedings
th
of the 10 International Conference on Sheet Metal,
Belfast, 2003.
[80] Jeswiet, J., Duflou, J., Szekeres, A. Forces in Single
Point and Two point Incremental Forming. Journal
Advanced Materials Research, Vols. 6-8, May 2005,
pp 449 - 456.
[81] Duflou, J., Szekeres, A., VanHerck. Force
Measurements for Single Point Incremental Forming:
and experimental study. Journal Advanced Materials
Research, Vols. 6-8, May 2005, pp 441 448.
[82] Nyahumwa, C. and Jeswiet, J. A Friction Sensor for
Sheet Metal Rolling. Annals of CIRP vol. 40/1/1991,
pp. 231-234.
[83] Young, D. and Jeswiet, J. Wall Thickness Variations
in Single Point Incremental Forming. IMECHE part B,
J. of Engineering Manufacture. Vol. 218 No B11,
2004; pp 204 210.
[84] Kitazawa,
K.,
Hayashi,
S.,
Yamazaki,
S.
Hemispherical stretch-expanding of aluminum sheet
by computerized numerically controlled incremental
forming process with two path method. Journal of

[85]

[86]
[87]
[88]

[89]

[90]

[91]

[92]

[93]

[94]

[95]

[96]

Japan Institute of Light Metals, vol. 46, pp. 219-224,


2001.
Kim, T.J., Yang, D.Y. Improvement of formability for
the incremental sheet metal forming process. Int. J.
of Mechanical Sciences, vol. 42, pp. 1271-1286,
2001.
Thomsen, E.G., Yang, T.Y., Kobayashi, S.
Mechanics of Plastic Deformation in Metal
Processing, 1965, (The Macmillan Company).
Jeswiet, J., Incremental Single Point Forming.
Transactions of North American Manufacturing
Research Institute; vol. XXIX, 2001, pp 7579.
Hagan, E. and Jeswiet, J. Analysis of surface
roughness for parts formed by CNC incremental
forming. IMECHE part B, J. of Engineering
Manufacture. Vol. 218 No. B10, 2004. pp 1307
1312.
S. Junk, G. Hirt, I. Chouvalova. Forming Strategies
and Tools in Incremental Sheet Forming.
th
Proceedings of the 10 International Conference on
Sheet Metal SHEMET 2003, 14-16 April 2003,
Jordanstown, UK, pp. 57-64.
Ambrogio, G., Costantino, I., De Napoli, L., Filice, L.,
Fratini, L., Muzzupappa, M. On the influence of some
relevant process parameters on the dimensional
accuracy in Incremental Forming: a numerical and
experimental investigation. Journal of Materials
Processing Technology; special edition for AMPT
2003, pp 501-507.
Ambrogio, G., Filice, L., Fratini, L., F. Micari. Some
relevant correlation between Process parameters
and Process Performance in Incremental Forming of
metal sheet. Proc. of the 6th Conference Esaform.
28-30 April 2003. Salerno Italy, pp. 175 178.
Hirt, G., Ames, J., Bambach, M. A new forming
strategy to realise parts designed for deep-drawing
by incremental CNC sheet forming, Steel Research
vol. 76 (2005) No. 2/3, pp 160-166.
Bambach, M., Hirt, G., Ames, J. Modeling of
optimization strategies in the incremental CNC sheet
metal
forming
process,
NUMIFORM
2004,
Columbus, Ohio, pp 1969-1974.
Kim, T. J. and Yang, D.Y. Improvement of formability
for the incremental sheet metal forming process,
International Journal of Mechanical Sciences 42
(2000) 1271-1286.
Iseki, H. An approximate deformation analysis and
FEM analysis for the incremental bulging of sheet
metal using a spherical roller, Journal of Materials
Processing Technology 111 (2001) 150-154.
Ambrogio, G., Filice, L. Fratini, L., Micari, F. Process
mechanics analysis in single point incremental
forming, NUMIFORM 2004, Columbus, Ohio.

Proceedings are on CD Rom.


[97] Hirt,, G., and Ames, J., Environmental and Economic
Benefits of CNC Incremental Sheet Forming for
Prototyping and Low Volume Production of Sheet
Components, 9th International Workshop on Ecology
and Economy in Manufacturing ICEM 2003, 18.
21. 10. 2003, Miskolc, Ungarn; No. 10, pp 70-79.
[98] Ordoobadi S M and Mulvaney N J, 2001,
Development of a justification tool for advanced
manufacturing technologies: system-wide benefits
value analysis, J Eng Tech Mgt, 18, 157-184.
[99] Allwood J M, King G P F, and Duflou J, 2004, A
structured search for applications of the Incremental
Sheet Forming process by product segmentation,
Proc I Mech E, Part B, J Eng Manuf, Vol 219, No.
B2, pp 239-244.

[100] Jadhav, S. Basic Investigations of the Incremental


Sheet Metal Forming Process on a CNC Milling
Machine. Dr.-Ing. Dissertation 2004, University of
Dortmund.
[101] Bramley, A. A new incremental sheet forming
process for small batch and prototype parts, 22
October, 2004, Cambridge University. CdRom.
[102] McAllister, P. and Jeswiet, J. Medical device
regulation for Manufacturers. IMECHE 2003 J.
Engineering in Medicine, Vol. 217 Part H; pp 459
467.
[103] Bramley, A.N., 2001, Incremental Sheet Forming
Process for Small Batch and Prototype Parts, in
Idee-Vision-Innovation Edited by F Vollersten and M
Kleiner, Verlag Meisenbach, 2001, ISBN 3-87525149-0, pp 95-102.
[104] Jeswiet, J. Recent results for SPIF. Seminar on
Incremental Forming, 22 October, 2004, Cambridge
University. CdRom.
[105] Duflou, J.R., Lauwers, B., Verbert, J., Tunckol, Y.,
De Baerdemaeker, H., Achievable Accuracy in
Single Point Incremental Forming: Case Studies. 8th
ESAFORM Conf. on Material Forming, Cluj Napoca,
April 2005, ISBN 973-27-1173-6, pp 675-678.
[106] Ambrogio, G., De Napoli L., Filice, L., Muzzupappa,
M., Improvement of Geometrical Precision in Sheet
Incremental Forming Processes, Proceedings of 7th
ASME Conference on Engineering System Design
and Analysis, July 19-22, 2004, Manchester, UK. no.
ESDA2004-58589, on CDRom.
[107] He, S., Van Bael, A., Van Houtte, P., Szekeres, A.,
Duflou, Henrard, C. and Habraken, A.M. Finite
element Modeling of Aluminum Sheets, J. of
Advanced Materials Research, vols. 6 8 (2005), pp
525 532.
[108] He, S., Van Bael, A., Van Houtte, P., Tunckol, Y,
Dufllou, J. R., Henrard, C., Bouffioux, C. and
th
habraken, A.M. Proc. Of 8 ESAFORM, April 27
29, 2005, Culj-Napoca, Roumania, pp 258 265.
[109] Henrard, C., Habraken, A.M., Szekeres, A., Duflou,
J.R., He, S., Van Bael, A. and Van Houtte, P. J. of
Advanced Materials Research, vols. 6 8 (2005), pp
533 542.

Вам также может понравиться