1 s2.0 S0020740319311415 Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

International Journal of Mechanical Sciences 164 (2019) 105143

Contents lists available at ScienceDirect

International Journal of Mechanical Sciences


journal homepage: www.elsevier.com/locate/ijmecsci

Investigations on hole-flanging by paddle forming and a comparison with


single point incremental forming
Lemopi Isidore Besong∗, Johannes Buhl, Markus Bambach
Chair of Mechanical Design and Manufacturing, Brandenburg University of Technology, Cottbus - Senftenberg, Konrad - Wachsmann - Allee 17, Cottbus D-03046,
Germany

a r t i c l e i n f o a b s t r a c t

Keywords: The use of dedicated dies to carry out hole-flanging operations is costly for small batch production. As a result,
Paddle forming hole-flanging by single point incremental forming (SPIF) was put forward because it offers advantages such as low
Hole-flanging equipment cost, high formability, and flexibility compared to conventional hole-flanging. However, drawbacks
Incremental sheet forming
of hole-flanging by SPIF such as poor geometrical accuracy and long process times are obstacles to its industrial
Deformation analysis
take-up. This paper presents a novel incremental hole-flanging method which uses paddle-shaped tools rotating
at high speeds while being fed axially into the sheets to form flanges, with the aim of preserving the advantages of
incremental forming while reducing the drawbacks related to process time. The results of the experiments show
higher formability for the paddle forming process compared to hole-flanging by multi-stage SPIF and conventional
hole-flanging. The high formability in paddle forming is attributed to the presence of shear which delays fracture.
Furthermore, heat generated due to friction between the paddle and metal blanks leads to a rise in the blank
temperature which enhances formability. Optical strain measurements of the formed flanges reveal that the
paddle speed and feed have major influences on the forming mechanism of the process and the process borders.
A maximum hole expansion ratio (HER) of 3.3 was obtained from the experiments in a process time of ∼2 s,
showing that the process combines high formability and low process times.

1. Introduction flanging produces flanges with good geometrical accuracy in a short


process time and is hence suitable for mass production. The flanges are
Many industrial applications of sheet metal components require formed in one press stroke or by multi-stage hole-flanging. In multi-
flanges as bearing seats, for tapping threads or to serve as position- stage hole-flanging, several forming operations are carried out to ob-
ing aids. Hole-flanging is a metal forming process used to expand the tain the flange. The blanks are progressively deformed and fed for-
diameter of holes in metal sheets with the aim of creating hollow cylin- ward from one tool set to another until the desired flange geometry
drical or cone-shaped flanges. The holes are usually created by shearing is achieved [5]. This requires the use of many expensive die sets which
or drilling [1]. The metal sheet is clamped firmly between a die and a makes the process economically feasible only for high volume manu-
blank holder. The flange is formed by a rigid punch which results in cir- facturing, such as in the automobile industry. As recently reviewed by
cumferential stretching at the hole edge and bending at the die radius. Yang et al. [6], there is a trend towards the production of variants and
Extensive research has been carried out on the process. Johnson et al. smaller lot sizes, which require more flexible production processes. This
[2] used experiments to investigate the mechanics of hole-flanging of paper introduces hole-flanging by paddle forming as a new process that
circular cups. They found that excessive circumferential strains lead to offers the advantages of flexible incremental forming such as low tooling
fracture at the edges of the flanges. Similarly, Hyun et al. [3] studied costs and high formability while keeping process times low compared to
the hole-flange ability of high strength steel sheets and proved that frac- conventional incremental forming operations such as hole-flanging by
ture was a result of localized necking due to circumferential stretching. SPIF.
Kacem et al. [4] examined failure prediction in hole-flanging and stated The rest of this chapter is divided as follows: Section 1.1 reviews
that amongst other parameters, ironing influences failure. hole-flanging by SPIF and its limitations. Hole-flanging by paddle form-
Flanges are usually made by conventional press forming operations ing is introduced as an alternative to SPIF in Section 1.2. The paddle
which make use of dedicated dies and punches. Conventional hole- shape used in the study is presented in Section 1.3.


Corresponding author.
E-mail address: besonlem@b-tu.de (L.I. Besong).

https://doi.org/10.1016/j.ijmecsci.2019.105143
Received 3 April 2019; Received in revised form 7 August 2019; Accepted 7 September 2019
Available online 9 September 2019
0020-7403/© 2019 Elsevier Ltd. All rights reserved.
L.I. Besong, J. Buhl and M. Bambach International Journal of Mechanical Sciences 164 (2019) 105143

Fig. 1. Setup of hole-flanging by paddle forming. (a) Start of the process. (b) Hole expansion. (c) Start of the hole-flanging process. (d) Flange forming. (e) End of
the process.

1.1. Single point incremental forming Using the strategy with an increasing flange diameter proposed by
Cui and Gao [5], Bambach et al. [12] obtained an HER of 5.5 on a 1 mm
Today’s markets are highly competitive and manufacturing enter- thick mild steel (DC04) which is significantly greater than the HER of ∼2
prises have to scale down the size of production batches to meet a wide achieved by conventional hole-flanging. Similarly, Silva et al. [13] used
variety of customer demands while remaining profitable. This requires optical deformation analysis to prove that formability is higher in hole-
the implementation of mass customization. Customization refers to the flanging by SPIF compared to conventional hole-flanging. Cristino et al.
capacity to provide individually designed products without compromis- [14] studied the forming limits of square-shaped flanges and showed
ing cost efficiency [7]. High costs are associated with the design, man- that the deformation mechanism is different at various portions of the
ufacture, usage, and disposal of dies and punches in conventional hole- flange. Petek et al. [15] proposed a variant of hole-flanging by SPIF
flanging. Jeswiet et al. [8] state that the use of dedicated forming equip- which uses a multi-stage backward tool path approach. The tool size,
ment is expensive for small batch production because of the few parts vertical and horizontal step sizes were found to have major effects on
manufactured. Consequently, there has been growing interest in the use the flange height and average flange thickness.
of incremental sheet forming (ISF) processes by manufacturing enter- Despite the advantages offered by hole-flanging using SPIF, the pro-
prises in the last decades, which reduce or eliminate the need for ded- cess is not widely used in industry because of the long process time [16],
icated tools and draw-upon low-cost equipment [9]. In addition, these geometrical inaccuracies such as bulges at the clamped end [17] and
processes have higher formability and flexibility compared to conven- conicity at the flange edge [12]. The flanges are usually produced us-
tional forming processes [10]. A recent review by Duflou et al. [11] gives ing several deformation stages which require the use of long tool paths
a current overview of the development of ISF. and consequently a long process time. Hence, accelerated incremental
Hole-flanging by multi-stage SPIF was introduced by Cui and Gao hole-flanging processes are needed. Laugwitz et al. [18] showed that
[5] to replace conventional hole-flanging in small batch production. incremental hole-flanging can be carried out at high speeds. The set-up
The process uses tools with hemispherical tips which follow specific tool used in the research draws upon a lathe where the blank rotates at high
paths and deform clamped sheet metals with pre-existing holes. Three speeds, i.e. the set-up is designed for basic research rather than for in-
tool path strategies were employed in [5] to study the effects of the dustrial application. Borrego et al. [19] investigated the production of
forming strategies on the length and thickness distribution of the formed flanges using single stage SPIF to reduce process time. An HER of 1.66
flanges. The first tool path strategy involved increasing the flange diam- was obtained using a 1.6 mm thick 7075-O aluminum sheet formed by
eter, the second was based on increasing the forming angle and the third a tool of 10 mm radius rotating at 1000 rpm. Presently, no solution is
strategy was a combination of the first two strategies. The first tool path available that combines the advantages of incremental forming (forma-
strategy yielded the flanges with the highest formability and uniform bility, flexibility) and conventional hole-flanging (short process time,
wall thickness. good geometrical accuracy).
The HER is often used as a measure of formability in hole-flanging
processes. The HER is the ratio of the maximum inside flange diameter 1.2. Paddle forming
of the finished flange to the minimum hole diameter required to form a
flange without blank failure (cracks). It is calculated using Eq. (1). The literature presented above establishes the need to reduce the
𝐷𝑓 process time and improve the geometrical accuracy of flanges made by
HER = (1) SPIF. Allwood and Shouler [20] recognized that sheet metal parts could
𝐷𝑖
be formed using paddle shaped tools, which have a line contact com-
Where Df is the inner diameter of the flange and Di is the initial diameter pared to a single point contact. The main advantages of paddle forming
of the hole. are a short process time and increased formability. For forming of three
L.I. Besong, J. Buhl and M. Bambach International Journal of Mechanical Sciences 164 (2019) 105143

Fig. 2. (a), (b) Sample paddle shapes from


[20]. (c) Side view of the paddle used in
the present study. (d) Bottom view of the
paddle used in the present study.

dimensional parts, paddle forming received much attention. It does, 1.4. Present research
however, seems to be highly suitable for hole flanging, where it could
combine flexibility, high formability and a short process time. Based on the limitations of the existing hole-flanging processes to
This research investigates the application of the paddle forming produce flanges in small batch production using a cost-effective method
process to axisymmetric hole-flanging. The axisymmetric shape of the while maintaining short process time, good geometrical accuracy and
flange permits high-speed paddle rotation with little kinematic effects high formability, experiments were conducted on axisymmetric paddle
acting on the paddle since rotational motion is self-stabilizing. The forming with the following aims:
length of the tool path and process time are significantly reduced in the
• to obtain the maximum formability (HER) using hole-flanging by
hole-flanging process because the flange is formed by the axial displace-
ment of the paddle compared to long tool paths required in hole-flanging paddle forming.
• to find the process limits (range of paddle speeds and feeds needed
by SPIF. The setup of axisymmetric hole-flanging by paddle forming is
shown in Fig. 1. to form flanges with no cracks).
• to investigate the effect of paddle feed and speed on the hole-flanging
As in hole-flanging by SPIF and conventional hole-flanging, the blank
is clamped by a blank holder and a die. The paddle rotating at high speed process.
• to determine the evolution of forces and temperature during the pro-
moves axially and expands a hole at the center of the blank to form a
flange. Multi-stage forming, which is used in hole-flanging by SPIF, can cess.
• to compare paddle forming to conventional hole-flanging and hole-
be mimicked in hole-flanging by paddle forming using a paddle which
has multiple sections (edges). In this case, the paddle makes contact flanging by SPIF for maximum HER and process time.
• to determine the deformation mechanisms in the hole-flanging pro-
with the milled hole at edge 1 or 2 (see Fig. 2c), depending on the hole
diameter Di . This causes expansion of the hole diameter and bending at cesses in order to explain the differences in process formability. This
the clamped end of the sheet as shown in Fig. 1b. As the paddle moves is done by the use of finite element analyses.
further into the die, edges 3 and 4 make contact with the sheet and
This paper is organized as follows: Section 2 presents the materials
increase the diameter of the hole as can be seen in Fig. 1c. The paddle
and methods used for the experiments. In Section 3, the results obtained
moves into the die, stretches the blank and increases the flange length,
from the paddle forming experiments are presented. Section 4 com-
see Fig. 1d. At the end of the process, the full flange length is obtained as
pares paddle forming to conventional hole-flanging and hole-flanging
shown in Fig. 1e. The paddle then returns to its original position at the
by SPIF. The dominant deformation mechanisms in the processes are in-
top of the set-up, as shown in Fig. 1a. The forming time is determined
vestigated in Section 5 using a combined numerical-analytical approach.
by the feed of the paddle and the flange length.
In Section 6, a discussion on formability is presented and used to ex-
1.3. Paddle shape plain the results of the experiments. This is followed by conclusions in
Section 7.
Candidate paddle designs in the original work of Allwood and
Shouler [20] include an axi-symmetric male paddle shown in Fig. 2a 2. Experiments
and a possible paddle geometry, see Fig. 2b. The effects of paddle shape
on flange geometrical accuracy and crack occurrence along the length The flanges studied in this work were made using 3 forming pro-
of the flanges formed by paddle forming was recently investigated using cesses, conventional forming, SPIF and paddle forming, with the aim of
finite element (FE) analyses by the present authors [21]. Concave pad- obtaining the minimum hole size needed to form flanges with no cracks.
dles, such as the one proposed by Allwood and Shouler [20] in Fig. 2b, The blanks used in the experiments were cut from the same stock of
form flanges in which material points in the middle of the flange length aluminum sheets (EN AW-6181-T1) purchased from Novelis to prevent
are subject to biaxial stretching. This indicates a high tendency for fail- variation in the mechanical properties of the blanks. This aluminum al-
ure. However, there was no bulging at the clamped end of the flange, loy is used in the construction of automobile body frames due to its high
which led to a good geometric accuracy. A convex-shaped paddle caused strength-to-weight ratio, good formability, and corrosion resistance. The
bulge formation at the clamped end of the sheet. The elements along the mechanical properties of the sheets were obtained by conducting ten-
flange length were in the safe zone of the forming limit diagram (FLD). sile tests following DIN EN 10,002-1 tensile test standards on a Zwick
The paddle shape in this study was designed so that the blank is Z250 tensile testing machine. The material obeyed the following hard-
deformed with 4 forming angles to mimic multi-stage hole-flanging by ening law: 𝜎 = 506.2 𝜀0.23 . Other mechanical properties were found to
SPIF with 4 stages, see Fig. 2c. Each angle on the edge of the paddle be as follows: Young’s modulus = 65 GPa, yield strength = 165 MPa, ul-
represents a forming stage in hole-flanging by multi-stage single point timate tensile strength = 353.8 MPa and percentage elongation at frac-
ISF. The first and third forming edges have an angle of 31.5° to the axis ture: 24%.
of symmetry of the paddle, while the second and fourth forming edges To carry out optical strain measurements on the outer surfaces of the
have an angle of 56°. This paddle shape was obtained by modifying the flanges after forming, grid points were etched on the blanks by electro-
concave paddle used in a prior work [21] to prevent crack occurrence chemical etching using aluminum electrolyte 8090 produced by Östling
along the flange length (see Appendix 1). The paddle was designed to GmbH, Solingen, Germany. The grids had a point distance of 2 mm and
make a sliding contact with the blank along a line as shown in Fig. 2d. the points were 1 mm in diameter.
L.I. Besong, J. Buhl and M. Bambach International Journal of Mechanical Sciences 164 (2019) 105143

Fig. 3. (a) Clamping set up for the hole-flanging exper-


iments. (b) Punch used for conventional hole-flanging.
(b) Single point incremental forming tool. (c) Paddle
shaped tool.

Fig. 4. Tool path strategies for hole-


flanging by SPIF. (a) Increasing flange di-
ameter (strategy 1). (b) Increasing flange
angle (strategy 2).

Table 1 2.2. Hole-flanging by SPIF


Test plan of conventional hole-flanging.

Forming strategy Tool diameter [mm] Vertical speed [mm/min] The flanges were made using an increasing flange diameter tool path
strategy (strategy 1) with 4 forming stages, see Fig. 4a. The flange di-
Single stage 40 500
ameters were increased in the following steps: 28 mm, 32 mm, 36 mm
Multi Stage 25 500
30 500 and 40 mm. Tools of 10 mm and 20 mm diameter were used to investi-
35 500 gate the effect of tool size on the maximum HER. The tools were rotated
40 500 at a speed of 2000 rev/min in one set of experiments to study the ef-
fect of tool rotation on the maximum HER. In the other experiments,
the tools were not rotated but they could rotate freely on the bearings
of the machine tool holder during the forming processes. The HER of
Holes of different diameters (Di ) were milled at the center of the flanges obtained by the 2 tool path strategies were compared: increas-
sheets used in the hole-flanging experiments. A blank holder and a die ing flange diameter (strategy 1) and increasing flange angle (strategy
having internal diameters of 42 mm were used to hold the blanks as 2). For strategy 2, only the 20 mm diameter tool with no rotation was
shown in Fig. 3a. The blanks had a size of 80 mm × 80 mm with a thick- used. The forming angles of the flanges were 30°, 50°, 70°, and 90°, see
ness of 0.8 mm. The same blank holder and die were used for all the Fig. 4b. The 2 tool paths had a step size of 0.5 mm. A tool travel speed
tests. The setup was mounted on a dynamometer to measure the form- of 4000 mm/min was used to perform the experiments. The tool paths
ing forces. A thermocouple was glued on the bottom (non-contact) sur- were derived by modeling the forming steps using the software Creo
face of the blanks to measure the blank temperature during the pro- 3.0 workbench. The tool paths of the 10 mm diameter tool are shown in
cesses. The forces and temperatures were recorded using a measuring Fig. 4. The numbers in Fig. 4 indicate the forming stages. The schedule
system purchased from Hottinger Baldwin Messtechnik GmbH, Darm- of the hole-flanging by SPIF experiments is presented in Table 2 with
stadt, Germany. The experiments were conducted on an Arrow 1000 the variables for each sequence of experiments.
vertical milling center manufactured by Cincinnati Milacron. Cold form-
ing oil Raziol CLF 100 was used as lubricant in all the tests. The punch 2.3. Paddle forming experiments
used in conventional hole-flanging is shown in Fig. 3b. Fig. 3c shows
the SPIF tool and the paddle is presented in Fig. 3d. The paddle-shaped tool with a diameter of 40 mm was clamped by a
tool holder, see Fig. 5a. The paddle rotating at high speeds was fed along
its axis of symmetry (z-axis) into the die and used to deform the blanks
until the flanges were formed. The paddle feed and speed were varied
over a range of values to determine their influence on the deformation
2.1. Conventional hole-flanging of the flanges. A clearance of 1 mm was left between the paddle and die.
The tests were conducted on blanks with a hole diameter Di of 12.1 mm.
The single stage conventional hole-flanging experiments were con- The Di was increased for subsequent tests with the aim of determining
ducted using a 40 mm diameter punch, see Fig. 3b. Multi-stage conven- the range of paddle feeds and speeds needed to form flanges and hence
tional hole-flanging was carried out in 4 forming steps using punches to identify the process limits. The hole diameters were: 12.1 mm, 13 mm,
of different diameters (25 mm, 30 mm, 35 mm, and 40 mm). The punch 15 mm, 16 mm, 20 mm, 26.7 mm and 30 mm, corresponding to HERs of
feed was 500 mm/min in all experiments. An overview of the experi- 3.3, 3, 2.7, 2.5, 2, 1.5 and 1.3 respectively. Two repetitions were car-
ments is presented in Table 1. ried out for conventional hole-flanging and hole-flanging by SPIF while
L.I. Besong, J. Buhl and M. Bambach International Journal of Mechanical Sciences 164 (2019) 105143

Table 2
Plan of experiments of hole-flanging by SPIF.

Tool path sequence Forming strategy Tool diameter [mm] Forming angle [0] Part diameter [mm] Rotational speed [rev/min]

1 Strategy 10 90 28 0
1 10 90 32 0
10 90 36 0
10 90 40 0
2 Strategy 10 90 28 2000
1 10 90 32 2000
10 90 36 2000
10 90 40 2000
3 Strategy 20 90 28 0
1 20 90 32 0
20 90 36 0
20 90 40 0
4 Strategy 20 90 28 2000
1 20 90 32 2000
20 90 36 2000
20 90 40 2000
5 Strategy 20 30 40 0
2 20 50 40 0
20 70 40 0
20 90 40 0

Fig. 6. Variation of forces during paddle forming.


Fig. 5. (a) Set up of paddle forming experiment. (b) Paddle tool.

3 repetitions were carried out for paddle forming to make the results 3.2. Temperature development during paddle forming
statistically significant.
Friction due to the relative motion of the paddle and the metal blanks
3. Results of paddle forming experiments caused heat to be generated at the contact points of the paddle and
metal blanks. Additional heat was generated by plastic deformation of
3.1. Forming forces the blanks since paddle forming is characterized by high strain rates
and short process times. Both heat sources led to a rise in blank tem-
The forces acting on the paddle during the paddle forming process perature. Thermocouples glued at about a quarter of the flange length
of a flange with a HER = 3.3 formed at a speed of 6000 rev/min and a from the edge of the flange were used to measure the blank temper-
feed of 1000 mm/min are shown in Fig. 6. The results show negligible ature during the forming processes. Fig. 7a shows the location of the
forces acting in the x- and y-directions (Fx and Fy ) since the rotation of thermocouple on the outer surface of the blanks before the hole-flanging
the paddle along the z-axis has a self-stabilizing effect in these directions processes and Fig. 7b shows its location after the processes. The temper-
and the displacement of the paddle is along the z-axis. In the z-direction, atures were recorded as close as experimentally possible to the flange
the force increased until the low peak in Fig. 6 which corresponds to edge. The temperatures were highest at the flange edge (since the pad-
the end of contact with the second edge of the paddle. The increase dle makes a higher number of contacts with the blank edge) and reduced
in Fz is due to bending of the blank at the clamped edge. There was a towards the clamped end. The number of rotations of the paddle in con-
slight drop in Fz which stayed constant for about 0.2 s before starting tact with the blanks during the hole-flanging process were calculated.
to increase again. As the paddle moved further into the die there was A high number of paddle rotations at the same paddle feed implies that
a sharp increase in Fz . A maximum force of 1050 N was measured at a more heat is generated at the contact points of the blank and paddle.
paddle position corresponding to Fig. 1c. The increase in Fz is because A maximum blank temperature of 370 °C was measured for a flange
paddle edge 4 has a large angle (same as edge 2) and higher forces with an HER of 3.3 formed at a paddle speed of 6000 rev/min and
are required to continue bending and stretching the sheet. From the an axial feed rate of 500 mm/min, see Fig. 7c. The paddle was in con-
high point (Fig. 1c), there was a progressive drop in Fz to zero as the tact with the blank for approximately 420 rotations during the forming
flange length increased since the blank length being stretched to form of this flange. The blank temperature reduced to 330 °C for the flange
the flange progressively reduces. formed at a reduced speed of 5000 rev/min at the same feed. The paddle
L.I. Besong, J. Buhl and M. Bambach International Journal of Mechanical Sciences 164 (2019) 105143

Fig. 7. (a) Blank before the hole-flanging process.


(b) Blank after the hole-flanging process. (c) Varia-
tion of temperature during the paddle forming pro-
cess.

Fig. 8. (a) Flange obtained from experiments. (b)


Variation of sheet thickness along flange length.

made contact with the blank for about 350 rotations in this case. Simi- 3.4. Distribution of principal strains
larly, there was a drop in maximum blank temperature to 320 °C for a
flange formed using a higher paddle feed of 1000 mm/min and speed of Principal strain measurements of the non-contact surfaces of the
6000 rev/min. The paddle made contact with the blank for about 210 blanks were used to study the material flow and mode of deforma-
rotations. The trend indicates a drop in blank temperature as the paddle tion of the flanges. This strain analysis method is based on the assump-
speed reduces and the paddle feed increases since the paddle is in con- tion that the volume of the blank remains constant and no shear defor-
tact with the blanks for a fewer number of rotations. The heat loss due mation occurs during the forming process. The designation ‘principal
to conduction and radiation was considered negligible given the short strains’ in this research actually refers to ‘logarithmic principal surface
process time. stretches’ since deformation in paddle forming is characterized by sig-
For the flanges with an HER of 3, a maximum blank temperature nificant shear and nonlinear strain paths [22]. The existence of these
of 320 °C was measured. Maximum blank temperatures of 290 °C and deformation mechanisms makes the 2D state plane stress assumptions
250 °C were measured for the flanges with HERs of 2.7 and 2.5. The based on Marciniak–Kuczyński (M-K) analysis [23] invalid. The strain
blank temperature increased in all experiments, however, lower tem- measurements were carried out at the start and end of the flanging pro-
peratures were recorded for larger Di . This is because for larger Di , the cess. The strains of the flanges were determined based on the strains of
blank area in contact with the paddle reduced and the paddle made con- the original blanks at the start of the process. These strain measurements
tact with the blanks for a fewer number of rotations, thus generating less cannot be used to describe the sequential evolution of the strains dur-
heat. ing the forming process but rather show the deformation of the flange
with respect to the original blank. The evolution of principal strains dur-
3.3. Flange thickness distribution ing the paddle forming process has been investigated for various paddle
shapes by the use of FE analyses [21]. The major principal strain of the
The optical strain analysis measurement system argus® was used flange in Fig. 8a is shown in Fig. 9a while the minor principal strain is
to determine the thickness reduction and deformation mode (principal presented in Fig. 9b.
strains) of the flanges. The system uses grids on the non-contact sur- The distribution of the principal strains (surface stretches) in the FLD
faces of the blanks to measure the surfaces stretches after forming, see of the flange in Fig. 8a is shown in Fig. 10. The principal strains are pre-
Fig. 8a, from which deformation measurements and the sheet thickness sented based on the legend (color shading) of the major strains in Fig. 9a.
are computed. A representation of the percentage thickness reduction of The strains were considered along a line section shown in Fig. 10. Point
a flange with an HER of 3.3 obtained at a paddle feed of 1000 mm/min A represents the clamped part of the sheet. The sheet is not deformed
and speed of 6000 rev/min is shown in Fig. 8b. The results show thick- in this area during the process. Point B indicates the bent edge of the
ening of the blank at the bent area of the flange of about 35%. Thinning flange, see Fig. 9b. This point corresponds to the wrinkling region (point
occurs along the flange length with a minimum percentage thickness B) in the FLD in Fig. 10. Bulging and wrinkling can occur at B if the sheet
reduction of 64 at the flange edge. is not properly clamped. At point C (about a third of the flange length),
L.I. Besong, J. Buhl and M. Bambach International Journal of Mechanical Sciences 164 (2019) 105143

Fig. 9. (a) Major strain distribution. (b) Minor strain distribution.

Fig. 10. FLD of the flange formed at a speed of 6000 rev/min and feed of
Fig. 11. Effect of paddle speed and feed on the process limits of the flanges with
1000 mm/min.
HER = 3.3.

the minor principal strain is at its maximum value, 0.14, see Fig. 9b. The
flange undergoes biaxial stretching in this zone as shown in Fig. 10. Bi- 3.5.1. Influence of paddle feed on the principal strains of the flanges
axial stretching is influenced by paddle shape, as shown in [21]. Paddle The effects of change in paddle feed on the principal strains are
speed and feed also controls biaxial stretching as will be demonstrated shown in Fig. 12. For a flange formed at a feed of 500 mm/min and speed
in Section 3.5. Point D shows a point at about half the flange length. The of 6000 rev/min (experiment 1 in Fig. 11), the paddle made contact
flange is deformed in-plane strain mode in this region. At the edge of the with the blank for about 420 rotations during the forming process and
flange (point E) the strains tend towards uniaxial tension as observed in the maximum temperature of the flange was 370 °C. The flange length
Fig. 10. was 15.05 mm. The principal strains in Fig. 12a show biaxial stretching
The accuracy of the strain measurement reduces towards the edge at point C and near plane strain at points D and E. The flange showed
of the flanges due to the large distance between the points on the no cracks at its edges. The paddle feed was increased to 1000 mm/min
grid which is caused by severe deformation at the flange edges. The (experiment 2 in Fig. 11), the paddle made contact with the blank for
widespread of the points on the FLDs towards the edge of the flange about 210 rotations. A maximum blank temperature of 320 °C was mea-
demonstrates severe deformation. sured. The flange length was 16.10 mm which is more than the flange
length in experiment 1 and indicates a higher tendency for the paddle
to draw the blank into the die at high feeds. Biaxial stretching increased
3.5. Effects of paddle speed and feed on principal strains at point C. Points D and E moved slightly away from plane strain, see
Fig. 12b. This flange was at the process limit. Very small cracks with
Flanges with the maximum formability (HER = 3.3) were chosen lengths under 0.1 mm could be seen at the edge of the flange. For the
for the experimental strain analyses because the strain measurements of flanges formed at a higher paddle feed of 1500 mm/min (experiment
these flanges are very sensitive to changes in paddle speed and feed since 3), the maximum blank temperature reduced to 130 °C in 140 rotations
high formability requires the best forming conditions. The principal of the paddle. This flange had cracks with lengths greater than 1 mm
strains of the flanges were compared to understand the effect of paddle at the edge; see Fig. 12c for the principal strains. At the highest pad-
speed and feed on process mechanics, crack occurrence and wrinkling. dle feed, 2000 mm/min (experiment 4), the paddle made approximately
The paddle feed was increased from 500 mm/min to 2000 mm/min 100 rotations in contact with the blank and the maximum blank temper-
while the speed was kept constant at 6000 rev/min in one set of ex- ature was 100 °C. Similarly, the flange had cracks, the FLD is shown in
periments (experiments 1–4 in Fig. 11). Experiment 5 was conducted to Fig. 12d. The pictures of the flanges and principal strain measurements
determine the minimum paddle feed required to form the flanges at the are provided in Appendix 1.
same paddle speed. In the other experiments, the paddle feed was kept As can be seen from the FLDs in Fig. 12, the strains along the flange
constant at 500 mm/min while the rotational speed was reduced from length (points C and D) move to the right and slightly upwards as the
6000 rev/min to 4000 rev/min (experiments 1-4′ in Fig. 11). The com- paddle feed increases. This variation in the surface stretches is due
bination of paddle speed and feed that produced flanges with no cracks to a higher rate of deformation of the material per unit time as the
are represented by the green squares in Fig. 11, while the experimental paddle feed increases. This leads to increased biaxial stretching of the
trials which formed flanges with cracks at the edges of the flanges are flanges at points C and D. As a result, flanges formed at high paddle
shown by the red circles in Fig. 11. feeds (1500 mm/min and 2000 mm/min) had cracks at their edges. The
L.I. Besong, J. Buhl and M. Bambach International Journal of Mechanical Sciences 164 (2019) 105143

Fig. 12. Change in principal strains of the flanges due to variation of paddle feed. (a) 500 mm/min feed (no crack). (b) 1000 mm/min feed (no crack). (c)
1500 mm/min feed (cracks). (d) 2000 mm/min feed (cracks).

3.5.2. Influence of paddle speed on the principal strains of the flanges


The influence of the paddle speed on the principal strains and process
border is presented in Fig. 14. The flanges formed at a paddle speed of
6000 rev/min and feed of 500 mm/min (experiment 1) had no cracks,
see the FLD in Fig. 14a. The result is the same as in Fig. 12a. The pad-
dle speed was reduced to 5000 rev/min (experiment 2′ in Fig. 11). A
maximum blank temperature of 330 °C was measured and the paddle
made about 350 rotations in contact with the blank. A flange length
of 13.95 mm was measured which is the shortest for the flanges. This
flange had no cracks and marked the process limit. There was a shift in
the strain points at C, D, and E towards the left in the FLD, as shown
in Fig. 14b. Flanges formed at a lower paddle speed of 4500 rev/min
had cracks (experiment 3′ in Fig. 11), see Fig. 14c for the principal
strains. The maximum blank temperature reduced to 300 °C and the pad-
Fig. 13. Influence of paddle feed on the principal strains of flanges formed at a dle made 320 rotations in contact with the blank. Similarly, the flanges
speed of 6000 rev/min. formed at a paddle speed of 4000 rev/min had cracks. The principal
strains are shown in Fig. 14d. The paddle made approximately 280 ro-
tations in contact with the blank and the maximum blank temperature
presence of the cracks at the edges leads to the spread of strain points at was 270 °C (see experiment 4′ in Fig. 11). The pictures of the flanges
the edges of the flanges, see Fig. 12c for a paddle feed of 1500 mm/min and strain measurements are provided in Appendix 1.
and Fig. 12d for a paddle feed of 2000 mm/min. There was a change in The shift of the strain points in the FLD to the left as the paddle speed
the distribution of the strain points between D and E from plane strain reduced from 6000 rev/min to 5000 rev/min is probably due to lower
towards uniaxial tension as the paddle feed increased. forming temperatures and less stretching of the material at a lower pad-
The highest temperature (410 °C) was measured at the lowest pad- dle speeds. However, the principal strains moved to the right and up-
dle feed, 250 mm/min (see experiment 5 in Fig. 11). The paddle made wards at points C and D for the speeds below 5000 rev/min, see Fig. 14c
about 840 rotations in this case and the flanges had cracks at the edges, and Fig. 14d. This indicates biaxial stretching at point C for lower pad-
probably due to the high blank temperature, which led to softening and dle speeds which leads to crack occurrence at the edges of the flanges
necking. In addition, repeated sliding contact between the paddle and formed at speeds of 4500 rev/min and 4000 rev/min. The minor strains
blank could lead to wear of the blank surface and hence promote failure. at the bend (point B) moved slightly towards zero from negative values.
A representation of the surface stretches shows the displacement of This shows less wrinkling at the bend of the flanges formed at speeds of
points C and D towards the right and upwards as the paddle feed in- 4000 rev/min and 4500 rev/min, which is due to a higher tendency for
creases, see Fig. 13. The full lines indicate experimental trials of the the paddle to draw the blank into the die at low paddle speeds. A paddle
flanges formed without cracks while the dotted lines indicate the flanges feed of 500 mm/min could be high for paddle speeds of 4500 rev/min
formed with cracks. and 4000 rev/min, since the paddle makes contact with the blanks for
L.I. Besong, J. Buhl and M. Bambach International Journal of Mechanical Sciences 164 (2019) 105143

Fig. 14. Change in principal strains due to variation of paddle speed. (a) 6000 rev/min speed (no cracks). (b) 5000 rev/min speed (no cracks). (c) 4500 rev/min
speed (cracks). (d) 4000 rev/min speed (cracks).

formed in the same time). The temperature was high enough to per-
mit forming flanges without cracks. However, there is a displacement of
the principal strains to the right on the FLD for the flanges with cracks
formed at lower speeds (4500 rev/min and 4000 rev/min). The strains
moved to the right on the FLD at lower paddle speeds because there
is more stretching of the blank in the direction of paddle feed per ro-
tation of the paddle. This is accompanied by a reduction in shear and
blank temperature during forming which causes a drop in formability
and leads to biaxial stretching in the middle portion of the flange and
hence crack formation. The full lines indicate the flanges with no cracks
while the dotted lines indicate the flanges with cracks.

3.6. Influence of the ratio paddle speed/feed on flange formability

Fig. 15. Influence of paddle speed on the principal strains of flanges formed at The ratio 𝛽 = speed/feed (units = rev/mm) is varied in machining op-
a feed of 500 rev/min. erations to accomplish roughing and finishing cuts. The paddle speed
and feed can be equated to the speed and feed of cutting tools in hole
drilling operations. 𝛽 was calculated for the paddle speeds and feeds
lower number of rotations (320 and 280). The paddle feed was reduced used in the hole-flanging experiments with an HER of 3.3 to find out if
to 250 mm/min to increase the number of rotations the paddle makes 𝛽 can be used as a guideline in determining the feasibility of conducting
contact with the blank. The number of paddle rotations in contact with hole-flanging by paddle forming. The green squares show the flanges
the blanks increased to approximately 640 and 560 rotations at paddle with no cracks and the red circles represent the flanges with cracks in
speeds of 4500 rev/min and 4000 rev/min. Similarly, the paddle feed Fig. 16.
was reduced to 250 mm/min for 5000 rev/min paddle speed. However, As the paddle speed was varied from 6000 rev/min to 4000 rev/min
all the flanges formed at paddle feeds of 250 mm/min had cracks. This at a feed of 500 mm/min (experiments 1-4′ in Fig. 11), the maximum
could be due to higher wear of the blank surface at low paddle feeds temperature and 𝛽 reduced as shown in Fig. 16a. The 𝛽 values were 12,
which may lead to failure. 10, 9 and 8 for the paddle speeds 6000 rev/min, 5000 rev/min, 4500
A representation of how the surface stretches vary with the paddle rev/min and 4000 rev/min, respectively. The paddle formed an equal
speed is shown in Fig. 15. The principal strains move to the left of the volume of blank material in the same time interval during these tests
FLD as the paddle speed reduces from 6000 rev/min to 5000 rev/min. since the axial feed of the paddle was the same. However, for 𝛽 below
The flange formed at a paddle speed of 5000 rev/min was the short- 10, fewer rotations of the paddle per unit time led to more stretching of
est. The paddle deforms more material per rotation of the paddle at the the material in each rotation of the paddle which led to the development
speed of 5000 rev/min (since equal volumes of the blank material are of cracks. This is indicated by movement of the principal strains to the
L.I. Besong, J. Buhl and M. Bambach International Journal of Mechanical Sciences 164 (2019) 105143

Fig. 16. (a) Change in temperature and the ratio of speed/feed with paddle speed. (b) Change in temperature and the ratio of speed/feed with paddle feed.

right and slightly up in the FLDs, see Fig. 15. This shows that there is a
minimum 𝛽 needed to form flanges without cracks as the paddle speed
varies.
The paddle feed was increased from 250 mm/min to 2000 mm/min
at a constant speed of 6000 rev/min (experiments 5 and 1–4 in Fig. 11).
This causes a reduction of 𝛽 and temperature, see Fig. 16b. The flanges
with no cracks lie in the middle of the distribution of 𝛽. At the lowest
paddle feed, 250 mm/min (experiment 5 which corresponds to 𝛽 = 24
rev/mm), the cracks were probably due to thermal softening of the blank
and wear caused by repeated contact of the blank and paddle. Crack
occurrence at the higher paddle feeds (experiments 3 and 4 which cor-
respond to 𝛽 = 4 rev/mm and 𝛽 = 3 rev/mm) was due to the low number
of rotations of the paddle which stretched the material within a short
time period leading to cracks. This was accompanied by a drop in the
maximum temperature of the sheets. Thus, there is an optimum range Fig. 17. Variation of process formability with process temperature.
of 𝛽 needed to produce flanges with no cracks. The values of 𝛽 were
considered with respect to change in speed and feed independently. No
interaction was found for the combined effect of the change in paddle cracks. At paddle speeds of 1200 rev/min and feeds of 100 mm/min
feed and speed on 𝛽. only the flange with a 1.3 HER had no cracks. The blank temperatures
were 165 °C for 3.3 HER and 130 °C for 1.3 HER. All the flanges formed
3.7. Influence of varying the paddle speed and feed at a constant 𝛽 on at paddle speeds of 120 rev/min and 36 rev/min had cracks. The re-
formability sults of the experiments show that high temperatures are required to
form flanges and the required temperature increases with an increase
The ratio 𝛽 was kept constant at 12 rev/mm (experiment 1 in Fig. 11), in HER. Similarly, there is an increase in shear as 𝛽 increases from low
while the paddle speed and feed were increased from low paddle speeds paddle speeds and feeds to high paddle speeds and feeds.
and feeds ([36 rev/min]/[3 mm/min]) to high paddle speed and feed
([6000 rev/min]/[500 mm/min]). The paddle speeds and feeds used in
the experiments are shown in Eq. (2).
3.8. Variation of process limits with the HER
𝑠𝑝𝑒𝑒𝑑 36 rev∕ min 120 rev∕ min 1200 rev∕ min
𝛽 = = = = From analysis of the flanges with an HER of 3.3, the flanges formed
𝑓 𝑒𝑒𝑑 3 mm∕ min 10 mm∕ min 100 mm∕ min
3000 rev∕ min 6000 rev∕ min at a speed of 6000 rev/min and feed of 500 mm/min (experiment 1
= = (2) in Fig. 11) yielded the best results. A reduction in paddle speed below
250 mm∕ min 500 mm∕ min
5000 rev/min and an increase in paddle feed above 1000 mm/min led to
The experiments were conducted on blanks having different HERs. biaxial stretching and crack occurrence at the edges of the flanges. The
The maximum temperatures measured from the experiments are pre- process limits were found to be between paddle speeds of 5000–6000
sented in Fig. 17. rev/min and feeds of 500–1000 mm/min as shown in Fig. 18a. The best
There was a drop in blank temperature as 𝛽 reduces from high pad- forming conditions are needed to form the flanges at the maximum HER,
dle speeds and feeds to low paddle speeds and feeds and as the HER hence only a small range of paddle feeds and speeds can successfully
value reduces for the same paddle speed and feed. All the blanks formed form the flanges. The variations of the maximum process temperature
using paddle speeds of 6000 rev/min and feed 500 mm/min had no with change in paddle speed and feed for flanges with a HER of 3.3
cracks. The blank temperature reduced with reduction in the HER from are shown in Fig. 18a. To examine the robustness of the process, flanges
370 °C at 3.3 HER to 170 °C at 1.3 HER. The blanks formed at a paddle were made with smaller HERs, down to 1.5. The paddle feeds and speeds
speed of 3000 rev/min and feed of 250 rev/min had lower temper- corresponding to the process borders of the different HERs are shown
atures compared blanks formed at a paddle speed 6000 rev/min and in Fig. 18. Flanges with no cracks are represented in green while those
feed 500 mm/min. The blank temperatures reduced from 210 °C at 3.3 with cracks are represented in red. The lines represent the borders of the
HER to 155 °C at 1.3 HER. Only flanges with an HER below 2 had no paddle feeds and speeds required to form flanges for the various HERs.
L.I. Besong, J. Buhl and M. Bambach International Journal of Mechanical Sciences 164 (2019) 105143

Fig. 18. Experimental results for hole-flanging by the paddle forming process. (a) HER 3.3 including temperature variation. (b) HER 3. (c) HER 2.7 and 2.5 including
the wrinkling zone. (d) HER 2 and HER 1.5.

For holes of 13.3 mm diameter (HER = 3), paddle speeds in the range The results of the experiments demonstrate that as Di increases, the
5000–6000 rev/min yielded successful flanges. A maximum paddle feed paddle speed needed to successfully form the flange reduces while the
of 1500 mm/min was attained as the process border, see Fig. 18b. Hole paddle feed increases. That is the range of paddle speeds and feeds (the
diameters of 15 mm (HER = 2.7) required paddle speeds in the range of area on the graph) used to form the flanges increases as Di increases
4500–6000 rev/min and paddle feeds between 500 and 1500 mm/min (HER reduces). There is a drop in the maximum temperature as the Di
to produce flanges without cracks as seen in Fig. 18c. Wrinkling occurred increases, see Section 3.2.
for the flanges formed at paddle feeds of 500 mm/min and 750 mm/min. Assuming a paddle axial feed of 1000 mm/min is used to calculate
For a larger diameter of 16 mm (HER = 2.5), paddle speeds in the range the forming time (this feed works for all HER values), a process time of
4000–6000 rev/min and feeds of 500–1750 mm/min yielded flanges approximately 2 s is obtained. This is appropriate for short process cycle
with no cracks, see Fig. 18c. Similarly, wrinkling was observed for the times even in the automobile industry.
flanges formed in the low feed range (500 mm/min-750 mm/min). The
samples with wrinkles are indicated by the shaded area in Fig. 18c. A 4. Comparison of hole-flanging by paddle forming, SPIF, and
picture of a flange with wrinkles is shown in Appendix 1. The wrinkles conventional hole-flanging using experimental investigations
may be due to less friction between the paddle and blank as a result of
a smaller contact area of the blank in these experiments which reduces Hole-flanging by paddle forming was compared to conventional
the tendency of the paddle to draw the blank into the die. In addition, hole-flanging and hole-flanging by SPIF to find out the forming limits,
at low paddle feeds, friction between the paddle and blank reduces and process time and differences in forming conditions.
the paddle draws the blank slowly into the die thereby allowing wrinkle
formation. 4.1. Forming forces
The hole diameter was increased to 20 mm (HER = 2). Paddle
speeds in the range 4500–6000 rev/min and feeds in the range 500– The force in the z-direction (Fz ) during paddle forming (see Fig. 6)
2000 mm/min produced flanges without cracks, see Fig. 18d. For has a similar shape to Fz (punch load) measured in conventional hole-
sheets with an initial hole diameter of 26.7 mm (HER = 1.5), paddle flanging, as reported in [24]. The forces in the x and y directions were
speeds in the range 2000–6000 rev/min and feeds in the range 250– negligible. The forces during the hole-flanging operation by SPIF strat-
2500 mm/min were found to yield successful flanges, as shown in egy 1 with the 20 mm diameter tool rotating 2000 rev/min are shown
Fig. 18d. The flanges formed at high paddle feeds (2500 mm/min) had in Fig. 19. The forces oscillate between positive and negative values in
rough surfaces. Edge cracks were present for all the flanges formed at a the x and y directions. However, the force in the z-direction is unidirec-
paddle speed of 6000 rev/min and feed of 250 mm/min, except for the tional as shown in Fig. 19c. The magnitudes of all the forces increase
HER of 1.5. with an increase in the number of forming steps.
L.I. Besong, J. Buhl and M. Bambach International Journal of Mechanical Sciences 164 (2019) 105143

4.3. Process time

The durations of the three forming processes were compared to de-


termine which processes have forming times relevant to industrial appli-
cations. A high tool travel speed of 4000 mm/min was used in the SPIF
experiments to reduce the process time. The hole-flanging experiments
by SPIF needed significantly longer forming time than paddle forming
and conventional hole-flanging. For tool path strategy 1, (see Fig. 4a),
the 10 mm diameter tool required 105 s to form the flange while the
20 mm diameter tool needed 84 s. The 20 mm diameter tool required
87 s for strategy 2.
Hole-flanging by conventional forming and paddle forming needed
a forming time of ∼2 s, this makes them significantly advantageous in
terms of saving time. The process time for hole-flanging by paddle form-
ing and conventional stamping are the same since the same axial feed
was considered for both tools and they were displaced by the same dis-
tance. In addition to the forming time, conventional hole-flanging and
axisymmetric paddle forming require no tool path generation in the
product development phase, which significantly reduces process setup
time.

4.4. Hole expansion ratio

The Di was varied until the limits of the flanges with no cracks at
its edge were identified. The three processes were compared to find out
which process has the highest HER.

4.4.1. Conventional hole-flanging


The results of the conventional hole-flanging experiments are pre-
sented in Table 3. The crosses (x) show the flanges formed with cracks
at the edges while the zeros (0) indicate flanges with no cracks. Con-
ventional hole-flanging (by stamping) yielded flanges with a minimum
Di = 27 mm. Multi-stage (MS) stamping had no influence on the Di . The
25 mm punch could not be used in the multi-stage stamping experiments
because the value for Di (27 mm) was greater than 25 mm. Hence, multi-
stage conventional hole-flanging was carried out in 3 forming steps.
Fig. 19. Force variation in SPIF. (a) X direction. (b) Y direction. (c) Z direction.
4.4.2. Hole-flanging by SPIF
Table 4 presents the process limits of the hole-flanging operations by
SPIF. The flanges formed with the 10 mm diameter tool using strategy
1 and no rotation required a minimum Di of 26 mm. The same tool ro-
tating at 2000 rev/min formed flanges with a minimum Di of 25 mm.
The blanks formed by the 20 mm diameter tool with no rotation needed
minimum Di of 23 mm. The 20 mm tool with a rotational speed of 2000
rev/min formed flanges with a minimum Di of 20 mm. The results reveal
that the minimum Di needed to form flanges reduces with increase in
tool diameter and rotation of the tool. Similar conclusions were drawn
in [19] using single stage hole-flanging by SPIF. The lower effect of tool
rotation on formability reported in that study is probably because the
flanges were formed in a single stage and the tools were rotated at a
Fig. 20. Temperature variation in SPIF for a flange formed using a 20 mm tool lower speed. In addition, the larger hole sizes used in that study re-
rotating at 2000 rev/min. quired longer tool paths which led to less temperature gain between the
forming steps. Hence, there was a reduction in the gain in formability
due to less blank temperature rise.
In the other experiments, the 20 mm tool with no rotation was used
4.2. Temperature to form flanges using strategy 2 to determine the effect of the tool path
strategy on the minimum Di . A minimum Di of 25 mm was obtained from
The temperature change during the conventional flanging experi- strategy 2 which is greater than 23 mm obtained by strategy 1. This
ments was insignificant. This is the same for the SPIF tests with no rota- result is in agreement with previous research. Cui and Gao [5] stated
tion of the tool. Tool rotation in SPIF led to a rise in the temperature of that the HER is larger for strategy 1 than strategy 2 because strategy 1
the blanks due to friction at the contact with the tool. A maximum tem- promotes bending of the blank which prevents thinning while strategy
perature of 76 °C was measured on the blanks formed by hole-flanging 2 stretches the blank leading to thinning and crack formation.
using SPIF strategy 1 with the 20 mm tool rotating at 2000 rev/min.
The temperature development is shown in Fig. 20. Temperatures above 4.5. Summary of results
300 °C were recorded for paddle forming as shown in Fig. 7c.
The effect of the blank size was not considered in the comparison of Paddle forming yields flanges with the highest formability
the forces and temperature during the forming operations. (HER = 3.3) as shown in Table 5. Flanges formed by SPIF had a
L.I. Besong, J. Buhl and M. Bambach International Journal of Mechanical Sciences 164 (2019) 105143

Table 3
Results of hole-flanging by conventional hole-flanging.

Process Tool diameter (mm) Diameter of pre-cut, Di (mm)

18 19 20 21 22 23 24 25 26 27 28 29 30
Stamping 40 x x x 0 0 0
MS stamping 30, 35, 40 x x 0 0

Table 4
Results of hole-flanging by SPIF.

Tool path strategy Tool size (mm) Spindle (rev/min) Diameter of pre-cut, Di (mm)
18 19 20 21 22 23 24 25 26 27 28 29 30

Strategy1 10 0 x x x x 0 0 0 0
Strategy1 10 2000 x 0
Strategy1 20 0 x 0 0 0 0
Strategy1 20 2000 x x 0 0 0
Strategy2 20 0 x x 0 0

Table 5
Results of hole-flanging experiments.

Tool path strategy Tool D (mm) Spindle (rev/min) Di min (mm) HERmax Max flange height (mm) Process time in seconds (s)

Stamping 40 0 27 1.48 8.30 2


Multi-stage stamping 30,35,40 0 27 1.48 8.50 25
Strategy1 10 0 26 1.54 7.80 105
Strategy1 10 2000 25 1.6 8.35 105
Strategy 1 20 0 23 1.74 8.80 84
Strategy1 20 2000 20 2 9.95 84
Strategy2 20 0 25 1.6 8.20 87
Paddle forming 40 variable 12.1 3.3 16.10 2

Fig. 21. Flanges obtained by paddle forming, sin-


gle point incremental forming (SPIF), and conven-
tional hole-flanging.

maximum HER of 2. An HER of 1.48 was obtained in conventional flang- approach from [25] is employed here to determine the quantitative con-
ing. The longest flanges made by paddle forming and SPIF had lengths of tribution of each deformation mode for the investigated processes. The
16.10 mm and 9.95 mm. The flange lengths were 8.30 mm and 8.50 mm deformation mechanisms in hole-flanging by stamping, SPIF and paddle
for the flanges made by single stage conventional hole-flanging and forming are considered to be a combination of membrane stretching,
multi-stage conventional hole-flanging. The multi-stage strategy leads to bending and shear deformation. The plastic energy dissipated during
a slight gain in flange length. The process times are included in Table 5. the process can then be written as the sum of the energies dissipated in
The flanges with the longest flange lengths formed by the 3 processes these three deformation modes, expressed as:
are shown in Fig. 21.
𝐸𝑝𝑙𝑎𝑠𝑡𝑖𝑐 = 𝐸𝑚𝑒𝑚𝑏𝑟𝑎𝑛𝑒 + 𝐸𝑠ℎ𝑒𝑎𝑟 + 𝐸𝑏𝑒𝑛𝑑𝑖𝑛𝑔 (4)
5. FE analyses of the hole-flanging processes
The forces and moment resultants along with the total internal en-
FE simulations provide a convenient and cost effective method to ergy are output from the numerical simulations and are converted into
study the deformation mechanisms in forming processes. The plastic the energy terms using a Matlab script [25]. The forces and moment
energy density from FE simulations can be split to determine the quan- resultants can be mathematically expressed as:
titative contribution of the various deformation mechanisms present in 𝑡
forming processes. The mathematical formulation for obtaining the plas-
𝑆𝐹 = 𝜎𝑑𝑧 (5)
tic energy density is: ∫
0
𝑡

𝐸𝑝𝑙𝑎𝑠𝑡𝑖𝑐 = 𝜎 ∶ 𝜀̇ 𝑝𝑙 𝑑𝑡 (3) where the membrane forces SF = (SFxx ,SFyy , SFxy ), the membrane
∫ stresses 𝜎 = (𝜎 xx ,𝜎 yy , 𝜎 xy ), dz is the change in blank thickness and t is
0
the process time.
In this regard, Maqbool and Bambach [25] developed a combined
analytical and numerical approach to split the plastic energy dissipa- 𝑡
tion during SPIF as a contribution of energies from membrane stretch- 𝑄= 𝜎𝑑𝑧 (6)
ing, bending and shear deformation mode. The analytical and numerical ∫
0
L.I. Besong, J. Buhl and M. Bambach International Journal of Mechanical Sciences 164 (2019) 105143

Fig. 22. (a) Element chosen for energy density


analyses. (b) Energy density for conventional
hole-flanging. (c) Energy density for SPIF. (d)
Energy density for paddle forming.

where Q = (Qxz ,Qyz ) are the shear forces and 𝜎 = (𝜎 xz ,𝜎 yz ) are the shear shown in Fig. 22b. The energy dissipated due to bending constitutes less
stresses. than 10% of the total internal energy. Only about 2% of the total dis-
𝑡 sipated energy is due to shear. In hole-flanging by SPIF, the dissipated
𝑀= 𝜎𝑧𝑑𝑧 (7) energy due to bending has the most significant effect on the process me-
∫0
chanics followed by membrane stretching. The shear energy makes up a
where M = (Mxx ,Myy ,Mxy ) are the moments and 𝜎 = (𝜎 xx ,𝜎 yy , 𝜎 xy ) are
minimal portion (about 4%) of the total dissipated energy, see Fig. 22c.
membrane stresses
A high internal energy is dissipated in hole-flanging by SPIF compared
The membrane energy is calculated by evaluating values of
to the other processes. The reason for this difference is that SPIF is a
𝜎 obtained in Eq. (5) with the corresponding strain increments
multi-stage forming process. In each step, the energy increases predom-
d𝜀 = (d𝜀xx ,d𝜀yy ,d𝜀xy ) into Eq. (3). Similarly, the shear and bending ener-
inantly by local bending and unbending of the sheet and accumulates
gies are derived by substituting the stress from Eq. (6) and Eq. (7) and
to a larger value than in the other process variants. Hence it can be
the corresponding strains in Eq. (3). The energy due to blank thinning
assumed that the energy densities in Fig. 22c adequately represent the
is added to the shear energy because the effects of normal compressive
process mechanics in hole-flanging by SPIF. For paddle forming, most of
stresses on localization of deformation are similar to shear stresses [26].
the deformation energy is dissipated as bending energy. This is followed
5.1. Set up of FE analysis by membrane stretching. The energy dissipated due to shear is higher in
paddle forming (about 15% of the total energy at its maximum value)
The aluminum blanks were modeled as deformable isotropic shells than in conventional hole-flanging and hole-flanging by SPIF, as can be
obeying a power law hardening rule as obtained from the tensile tests. seen in Fig. 22d.
The shells were defined as thick shell elements based on the Belytschko–
Tsay element formulation with 5 integration points in the through thick- 6. Discussion
ness direction. This element type captures the through-thickness shear
and the stress-strain gradient. The blanks had a hole diameter of 27 mm The results of experiments in Section 3 and Section 4 reveal a
which correspond to the minimum hole size in conventional hole flang- difference in formability of hole-flanging by paddle forming and the
ing, see Fig. 21c. Clamping of the blanks was simplified by the use of other process variants. The energy densities from the FE simulations in
fixed boundary conditions. The same blank, die and clamping conditions Section 5 may be used to explain the differences in process formability.
are used in all the simulations. The paddle, punch, SPIF tool and die were In the following sections, firstly, the deformation mode in conventional
defined as non-deformable rigid bodies. The paddle forming simulation hole-flanging is reviewed and compared with the energy density ob-
was conducted using the paddle speed and feed in experiment 1, see tained from the process simulation. Secondly, the energy density from
Fig. 11. Conventional hole-flanging was performed at a punch feed of the simulation of hole-flanging by SPIF is used to explain the process
500 mm/min. The flange formed by SPIF was simulated using the 10 mm mechanism in hole-flanging by SPIF. Relevant literature is used to sup-
diameter tool with no tool rotation and tool path strategy 1. The tool port the results. Thirdly, the role of through thickness shear stresses in
path was established using a Matlab script based on the tool path of the enhancing formability is discussed. It is then used to justify the higher
experiments. A coefficient of friction of 0.1 was assigned at the contacts formability in paddle forming. Fourthly, the influence of temperature
between the parts. The processes were simulated using the implicit dy- on formability is discussed. The section ends with a discussion on trans-
namic FE code of LSDyna solver. Temperature changes during forming ferability of the high formability and deformation mechanism of paddle
were not included in the simulations. forming to other forming processes.

5.2. Results of the simulations 6.1. Deformation mechanism in conventional hole-flanging

Energy densities for an element close to the flange edge (where the Formability in conventional forming processes is usually represented
maximum blank stretch occurs and cracks are likely to develop) are used using necking limit curves derived from formability tests. The stresses in
for the analyses, see Fig. 22a. For conventional hole-flanging, the dissi- the plane of the sheet are considered to be far greater than the stresses
pated energy of the element is mainly due to membrane stretching as perpendicular to the plane of the sheet. Deformation occurs mainly due
L.I. Besong, J. Buhl and M. Bambach International Journal of Mechanical Sciences 164 (2019) 105143

existence of in-plane and out-of-plane normal and shear strains in SPIF


experiments by measuring the stretch at the edges of brazed aluminum
blanks of 3.1–3.3 mm thickness. Shear was maximum at the contact sur-
face of the tool and sheet. Thus, the plane stress assumptions used to
obtain the M-K forming limits cannot be used to describe the forming
mechanism of ISF due to bending, nonlinear strain paths, and significant
in-plane and through-thickness shear [26].
Shouler and Allwood [31] used specially designed tensile tests to
show that high membrane shear (𝜎 12 ) leads to increased formability.
Fatemi and Daraini [32] used numerical analyses to investigate the ef-
fect of normal and through thickness shear stresses on formability us-
Fig. 23. A 3D representation of the global and local coordinate sets used to ing a modified M-K model. The presence of normal and through thick-
interpret the forming strains in the hole-flanging processes. ness shear stresses enhanced formability by movement of the forming
limit curves upwards. Experimental investigations on annealed Al 1050
to in-plane stresses 𝜎 11 and 𝜎 22 (no in-plane rotation) and the out-of- showed that the presence of out-of-plane shear stresses (𝜎 13 and 𝜎 23 )
plane (through-thickness) stress, 𝜎 33 are assumed to vanish. In addition, lead to an increase in formability [33]. Similar conclusions were ob-
the strain paths are assumed to be linear during the forming process. tained by Eyckens et al. [34] using an analytical procedure. A physical
The feed of the punch in conventional hole-flanging is in the global explanation for increased formability is that the presence of through-
z-direction, see Fig. 23. Contact of the punch and sheet causes in-plane thickness stresses within the calculation of the yield stress ensures that
(membrane) stresses. These stresses lead to deformation of the blank at the beginning of plastic deformation the principal stresses in the
which leads to local stretching mostly in direction 2, thus the major plane of the sheet are lower than when no through-thickness stresses
strain 𝜀22 . The minor strain 𝜀11 is in direction 1. The membrane shear are present. This permits increased plastic deformation of the necking
stress 𝜎 12 is ignored since the punch does not rotate. The strains in the zone before failure occurs [35]. Out-of-plane normal and shear stresses
thickness direction (𝜀33 ) can be considered zero given that there is a mi- improve formability since they may act in planes which prevent void
nor change in (𝜀33 ) during the forming process and little shear is present. nucleation, crack growth and coalescence [33].
These assumptions largely agree with the energy densities obtained from Paddle rotation induces membrane shear strain, 𝜀12 and out-of-plane
the FE simulations in Fig. 22b. The dissipated energy is mainly due shear strains 𝜀13 and 𝜀23 . The axial feed of the paddle stretches of the
to membrane stretching of the sheet which represents the membrane blanks and causes strains 𝜀11 , 𝜀22 and 𝜀33 . Paddle forming processes
stresses 𝜎 11 , 𝜎 22 and 𝜎 12 . The dissipated energies due to bending and were specifically designed to induce high shear on the blanks at the slid-
through thickness shear have little effect on the deformation mecha- ing contact between the blank and paddle. The paddle makes a larger
nism. Hyun et al. [3] also state that deformation in conventional hole- contact area with the blanks compared to SPIF tools which permit more
flanging occurs by bending and stretching. Stretching promotes failure deformation of the blank by shearing. These assumptions agree with
because voids in the material easily coalesce to form cracks. the shear energy from the simulations, Fig. 22d. Deformation in paddle
forming is achieved by shearing, bending and stretching compared to
6.2. The dominant deformation mechanisms in hole-flanging by SPIF the deformation in SPIF, where bending and stretching prevail. Thus,
the higher formability of paddle forming compared to hole-flanging by
The dissipated energy in hole-flanging by SPIF reveals that the bend- SPIF is explained by the higher amount of shear present in hole-flanging
ing energy is more significant than the membrane stretching in the pro- by paddle forming. Hence, the assumptions of the total strain membrane
cess, see Fig. 22c. Through thickness shear has a very little role on the theory of rigid-plasticity used in conventional hole-flanging [36] can-
process. Silva et al. [13] stated that the higher formability of SPIF com- not be used for paddle forming given that the deformation energy due
pared to conventional hole-flanging is because the tool locally deforms to membrane stretching is less than 40% of the internal energy, see
the sheet and induces stretching and bending which agrees with the re- Fig. 22d.
sults in Fig. 22c. Similarly, Emmens and van den Boogaard [26] propose The bending energy (as a proportion) is highest in hole-flanging by
bending under tension as a reason for increased formability in SPIF. Cen- paddle forming where the radius at the paddle edge is 2 mm. The bend-
teno et al. [27] showed that bending leads to through-thickness stress ing energy remains the same for hole-flanging by SPIF with a 5 mm tool
variation and enhances formability in ISF processes. A through-thickness radius due to the accumulation of local bending and unbending. The
stress strain gradient enhances formability by delaying fracture in the lowest bending energy was observed in conventional hole-flanging with
inner surface of the bend due to the effect of the contact stress (stress the 20 mm punch radius. Thus, the bending energy reduces with an in-
triaxility) [26]. A small tool bends the blank more than the larger tool. crease in tool radius. Similarly, Morales-Palma et al. [37] showed that
This, in theory, should lead to higher formability of the flanges made bending reduces with an increase in tool radius.
using the smaller tool. However, too much bending by smaller tools The element used to analyze the deformation mechanism in the sim-
blocks material flow around the tool, which reduces formability in SPIF ulations corresponds to a point close to the edge of the flange with a
[28]. As a result, the 20 mm tool forms flanges with higher formability HER of 1.48 (conventional hole-flanging in Fig. 21c). However, as the
compared to the 10 mm tool. Furthermore, tool rotation increases the HER increases, the amount of shear energy in the paddle forming op-
formability of the flanges made by the 2 tool sizes. This is probably due erations increases, which explains the high formability in hole-flanging
to an increase in shear strains 𝜎 12 and 𝜎 23 . by paddle forming.

6.3. The significance of through-thickness shear on formability in 6.4. Influence of temperature


hole-flanging by paddle forming
No temperature change was detected for the flanges formed by con-
Research on ISF has highlighted the existence of in-plane and out-of- ventional hole-flanging and SPIF with no tool rotation. A maximum
plane normal and shear stresses/strains in these processes due to friction blank temperature of 76 °C was measured for the flanges made by
at the contact area of the blank and tool. Bambach et al. [29] observed the 20 mm diameter tool using SPIF strategy 1 and rotating at 2000
in-plane and out-of-plane normal and shear stresses in ISF by using FE rev/min. This is far less than a blank temperature of 370 °C obtained
investigations. The stresses were found to be dependent on the tool di- in paddle forming. The generated heat caused thermal softening of the
ameter and step size. Similarly, Jackson and Allwood [30] showed the sheets. Thermal softening is beneficial because it leads to a drop in the
L.I. Besong, J. Buhl and M. Bambach International Journal of Mechanical Sciences 164 (2019) 105143

yield strength of the formed material and consequently the forming pro- • A maximum HER of 3.3 was achieved for the flanges formed by
cess requires lower forming forces. Thermal softening also leads to an paddle forming. The flanges made by SPIF and conventional hole-
increase in formability. Naka et al. [38] showed that there is an increase flanging had HERs of 2 and 1.48. This shows higher formability for
in the forming limit strain due to increase of the blank temperature in paddle forming compared to the other processes. Shear and heat gen-
sheet metal forming. The high paddle speeds required to form the flanges erated during forming account for the increased formability.
demonstrate the effect of the gain in temperature due to friction at the • Hole flanging by paddle forming and convention hole-flanging re-
contact of the blank and paddle. However, high paddle speeds and low quire a process time of ∼2 s compared to a minimum of 84 s needed
feeds may also lead to easy failure of the sheet due to loss of strength in hole-flanging by SPIF.
of the sheet as it softens. This may be the reason for the occurrence of • FE simulations of the hole-flanging processes reveal a high shear en-
cracks in the flanges formed at a low paddle feed of 250 mm/min in ergy density in paddle forming compared to the other hole-flanging
Fig. 18. methods. This is used to explain the high formability in hole-flanging
by paddle forming.
6.5. Transferability of knowledge on paddle forming to other processes
Future research will seek to experimentally investigate the effect of
The aim of this section is to propose how the gain in formability due paddle shape on the principal strain distribution of the flanges with the
to the process mechanics in paddle forming can be transferred to other aim of obtaining the paddle shape with a maximum HER. The range
forming processes to improve their formability. of paddle speed and feed required to form flanges using the paddle
shape with the maximum HER will be investigated. Flange geometries
• The high formability obtained in paddle forming is due to the in- obtained by paddle forming will be compared to ideal CAD geometries
fluence of significant shear in the process mechanics compared to to investigate the geometrical accuracy of flanges obtained by the pro-
the other flanging processes (SPIF, punch-based). This suggests that cess. Furthermore, formability tests will be conducted on the blank ma-
shear should be intentionally induced to forming processes when terial at high strain rates and temperatures to observe their effects on the
possible to improve formability. Similar suggestions have been made forming limits and to develop a model for formability. Finally, thermo-
in [31] and [33]. Tool rotation is an easy method to apply shear in mechanical FE simulations will be conducted to study the influence of
axisymmetric parts. High-speed tool rotation may be used to improve temperature on the process mechanics.
the formability especially for hard to form materials.
• Shear and temperature were proposed as the reasons for increased Acknowledgements
formability in paddle forming. Emmens and van den Boogaard
[26] cited cyclic loading, bending under tension, contact stress The depicted results were achieved to a great extent in the project
amongst other mechanisms, as reasons for the increased formability “Roboterbasiertes Kragenziehen” (Ref.-No. AiF 20457BG), which is fi-
in ISF processes. The contribution of the individual process mech- nanced and supervised by the European Research Association for Sheet
anisms to increased formability in ISF processes remains complex. Metal Working (EFB). In the scope of the program to promote Industrial
The results presented in this paper indicate that under certain con- Collective Research they were funded by the German Federation of In-
ditions, shear and temperature account for increased formability. dustrial Research Associations (AiF) with means of the Federal Ministry
of Economic Affairs and Energy (BMWi) on the basis of a decision by
7. Conclusions the German Bundestag.
This funding is gratefully acknowledged.
The present research was carried out to investigate axisymmetric
hole-flanging by using paddle-shaped tools. The axisymmetric nature
Supplementary materials
of the process was utilized to attain high-speed rotation of the paddle.
Optical strain measurements were carried out on the flanges formed by
Supplementary material associated with this article can be found, in
paddle forming. Flanges made by paddle forming, conventional hole-
the online version, at doi:10.1016/j.ijmecsci.2019.105143.
flanging, and SPIF were compared for maximum HER and process time.
From the results of the investigations, the following conclusions can be
made: Appendix

• The success of hole-flanging by paddle forming is highly dependent Supplementary data


on paddle speed and feed. For large HER values, high paddle speeds
and low feeds are required to form flanges. The range of paddle speed References
and feed needed to form flanges increases with a decrease in HER
(increase Di ). [1] Konieczny A, Henderson T. On formability limitations in stamping involving sheared
• An increase in paddle feed leads to displacement of the principal edge stretching. SAE Technical Paper. 2007-01-0340.
[2] Johnson W, Chitkara NR, Ibrahim AH, Dasgupta AK. Hole-flanging and punching of
strains (principal surface stretches) of the flanges to the right and circular plates with conically headed punches. J Strain Anal 1973;8(3):228–41.
upwards in the FLD. This indicates an increase in biaxial stretching [3] Hyun DI, Oak SM, Kang SS, Moon YH. Estimation of hole flangeability for high
as paddle feed increases which leads to cracks on the flanges formed strength steel plates. J Mater Process Technol 2002;130-131:9–13.
[4] Kacem A, Krichen A, Manach PY, Thuillier S, Yoon JW. Failure prediction in the
at high feeds. Biaxial stretching increases at low paddle speeds for hole-flanging process of aluminium alloys. Eng Fract Mech 2013;99:251–65.
the same paddle feed which leads to crack formation. [5] Cui Z, Gao L. Studies on hole-flanging process using multistage incremental forming.
• Friction at the contact points between the paddle and blank leads CIRP J Manuf Sci Technol 2010;2(2):124–8.
[6] Yang DY, Bambach M, Cao J, Duflou JR, Groche P, Kuboki T, Sterzing A, Tekkaya AE,
to a rise in blank temperature. A maximum process temperature of
Lee CW. Flexibility in metal forming. CIRP Ann 2018;67(2):743–65.
370 °C was obtained for a flange with an HER of 3.3 formed at a [7] Loginova O, Wang XH. Customization: ideal varieties, product uniqueness and price
feed of 500 mm/min and speed of 6000 rev/min. High forming tem- competition. Econ Bull 2009;29(4):2573–81.
[8] Jeswiet J, Micari F, Hirt G, Bramley A, Duflou J, Allwood J. Asymmetric single point
peratures correlate with good formability. The maximum blank tem-
incremental forming of sheet metal. CIRP Ann Manuf Technol 2005;54(2):623–50.
perature decrease with increase in paddle feed, reduction in paddle [9] Ambrogio G, de Napoli L, Filice L, Gagliardi F, Muzzupappa M. Application of in-
speed and increase Di . cremental forming process for high customised medical product manufacturing. J
• A decrease in paddle speed leads to a reduction in wrinkling at the Mater Process Technol 2005;162-163:156–62.
[10] Allwood JM, King GPF, Duflou J. A structured search for applications of the incre-
bend of the flange due to an increased tendency for the paddle to mental sheet forming process by product segmentation. Proc Inst Mech Eng B J Eng
draw the sheet into the die at low speeds. Manuf 2006;219(2):239–44.
L.I. Besong, J. Buhl and M. Bambach International Journal of Mechanical Sciences 164 (2019) 105143

[11] Duflou JR, Habraken AM, Cao J, Malhotra R, Bambach M, Adams D, Vanhove H, [26] Emmens WC, van den Boogaard AH. An overview of stabilizing deformation mecha-
Mohammadi A, Jeswiet J. Single point incremental forming: state-of-the-art and nisms in incremental sheet forming. J Mater Process Technol 2009;209(8):3688–95.
prospects. Int J Mater Form 2017;11(6):743–73. [27] Centeno G, Bagudanch I, Martínez-Donaire AJ, ML García-Romeu ML, C Vallellano C.
[12] Bambach M, Voswinckel H, Hirt G. A new process design for performing hole-flang- Critical analysis of necking and fracture limit strains and forming forces in single–
ing operations by incremental sheet forming. Procedia Eng 2014;81:2305–10. point incremental forming. Mater Des 2014;63(20):20–9.
[13] Silva MB, Teixeira P, Reis A, Martins PAF. On the formability of hole-flanging by [28] Al-Ghamdi KA, Hussain G. Threshold tool-radius condition maximizing the forma-
incremental sheet forming. Proc IMechE 2013;227(2):91–9. bility in SPIF considering a variety of materials: experimental and FE investigations.
[14] Cristino VAM, Montanari L, Silva MB, Martins PAF. Towards square hole-flanging Int J Mach Tools Manuf 2015;88:82–94.
produced by single point incremental forming. Proc IMechE 2014;229(5):380–8. [29] Bambach M, Hirt G, Junk S. Modelling and experimental evaluation of the incremen-
[15] Petek A, Kuzman K, Fijavž R. Backward drawing of necks using incremental ap- tal CNC sheet metal forming process. Proceedings of the 7th COMPLAS, Barcelona,
proach. KEM 2011;473:105–12. Spain 2003, April 7–10.
[16] Bastos RNP, de Sousa RAJ, Ferreira JAF. Enhancing time efficiency on single point [30] Jackson K, Allwood Julian. The mechanics of incremental sheet forming. J Mater
incremental forming processes. Int J Mater Form 2016;9(5):653–62. Process Technol 2009;209(3):1158–74.
[17] Voswinckel H, Bambach M, Hirt G. Improving geometrical accuracy for flanging by [31] Shouler DR, Allwood JM. Design and use of a novel sample design for formability
incremental sheet metal forming. Int J Mater Form 2015;8(3):391–9. testing in pure shear. J Mater Process Technol 2010;210(10):1304–13.
[18] Laugwitz M, Voswinckel H, Hirt G, Bambach M. Development of tooling concepts to [32] Fatemi A, Dariani BM. The effect of normal and through thickness shear stresses on
increase geometrical accuracy in high speed incremental hole flanging. Int J Mater the formability of isotropic sheet metals. J Braz Soc Mech Sci Eng 2016;38(1):1–13.
Form 2018;11(4):471–7. [33] Allwood JM, Shouler DR. Generalised forming limit diagrams showing increased
[19] Borrego M, Morales-Palma D, Martínez-Donaire AJ, Centeno G, Vallellano C. Exper- forming limits with non-planar stress states. Int J Plast 2009;25(7):1207–30.
imental study of hole-flanging by single-stage incremental sheet forming. J Mater [34] Eyckens P, van Bael A, van Houtte P. Marciniak–Kuczynski type modelling of the
Process Technol 2016;237:320–30. effect of through-thickness shear on the forming limits of sheet metal. Int J Plast
[20] Allwood JM, Shouler DR. Paddle forming. a novel class of sheet metal forming pro- 2009;25(12):2249–68.
cesses. CIRP Ann Manuf Technol 2007;56(1):257–60. [35] Allwood JM, Shouler DR, Tekkaya AE. The increased forming limits of incremental
[21] Besong LI, Buhl J, Bambach M. Paddle shape optimization for hole-flanging by pad- sheet forming processes. KEM 2007;344:621–8.
dle forming through the use of a predefined strain path in finite element analysis. J [36] Wang NM, Wenner ML. An analytical and experimental study of stretch flanging. Int
Mach Eng 2019;19(2):83–98. J Mech Sci 1974;16:135–43.
[22] Allwood JM. Sheet metal forming with six components of strain. Micromechanics [37] Morales-Palma D, Vallellano C, García-Lomas FJ. Assessment of the effect of the
Seminar. Friday 8 May 2009. Department of Engineering. University of Cambridge. through-thickness strain/stress gradient on the formability of stretch-bend metal
[23] Marciniak Z, Kuczyński K. Limit strains in the processes of stretch-forming sheet sheets. Mater Des 2013;50:798–809.
metal. Int J Mech Sci 1967;9(9):609–20. [38] Naka T, Torikai G, Hino R, Yoshida F. The effects of temperature and forming speed
[24] Krichen A, Kacem A, Hbaieb M. Blank-holding effect on the hole-flanging process of on the forming limit diagram for type 5083 aluminum–magnesium alloy sheet. J
sheet aluminum alloy. J Mater Process Technol 2011;211(4):619–26. Mater Process Technol 2001;113(1–3):648–53.
[25] Maqbool F, Bambach M. Dominant deformation mechanisms in single point incre-
mental forming (SPIF) and their effect on geometrical accuracy. Int J Mech Sci
2018;136:279–92.

You might also like