Anda di halaman 1dari 143

THE TRIBOLOGY AND FORMABILITY OF ZINC COATED STEEL SHEETS

SUBJECTED TO DIFFERENT STRAIN STATES

by
YOHAN JANG

Submitted in partial fulfillment of the requirements


For the degree of Master of Science

Thesis Adviser: Professor. Gary M. Michal

Department of Materials Science and Engineering


CASE WESTERN RESERVE UNIVERSITY

May, 2010

CASE WESTERN RESERVE UNIVERSITY


SCHOOL OF GRADUATE STUDIES

We hereby approve the thesis/dissertation of


Yohan Jang

candidate for the

(signed)

Master of Science

degree *.

Professor Gary M. Michal


(chair of the committee)
Professor Gerhard E. Welsch

Professor John J. Lewandowski

(date)

03/ 24/ 2010

* We also certify that written approval have been obtained for any
proprietary material contained therein.

Copyright 2010 by Yohan Jang


All rights reserved

TABLE OF CONTENTS
Table of Contents ..............................................................................................................
List of Tables ....................................................................................................................
List of Figures ...................................................................................................................
Acknowledgement ...........................................................................................................
Abstract ...........................................................................................................................

1. Introduction ....................................................................................................................1
2. Background .....................................................................................................................4
2.1 Static Friction Theory ..............................................................................................4
2.1.1 Coulomb and Amontons Friction Theory ....................................................4
2.1.2 Bowden and Tabors Friction Theory ...........................................................5
2.1.3 Orowan and Shaws Friction Theory .............................................................7
2.2 Dynamic Friction Theory ........................................................................................8
2.2.1 Stribeck Friction Theory ..............................................................................10
2.2.2 Daltons Friction Theory ..............................................................................12
2.3 Fracture Mechanics of Coated Steel Sheet ............................................................15
2.3.1 Wedging Mechanism Causing Powdering ....................................................16
2.3.2 Buckling Mechanism Causing Powdering ....................................................17
3. Experimental Procedure ................................................................................................20
3.1 Material Properties .................................................................................................20
3.1.1 Special Treated Zinc Coated (STZC) Materials...........................................20
3.1.2 Hot- Dip Galvanized (GI) Material..............................................................23

iv

3.1.3 Galvanneal (GA) Material............................................................................24


3.2 Measurement of Real Area of Contact .................................................................24
3.2.1 Systemic Manual Point Count Method ........................................................25
3.3 Mechanical Testing ..............................................................................................27
3.3.1 Measurement of Surface Friction Coefficient ..............................................27
3.3.1.1 Cup Test .......................................................................................28
3.3.1.2 Hemispherical Punch Test ...........................................................33
3.3.2 Measurement of Coating Adhesion Propensities ........................................38
3.3.2.1 Powdering Tests ...........................................................................38
3.3.3 Measurement of Formability ......................................................................40
3.3.3.1 Limiting Dome Height (LDH) Test ...........................................40
3.4 Surface Analysis ..................................................................................................40
3.4.1 Surface Analysis by SEM ..........................................................................42
3.4.2 Surface Analysis by ESCA ........................................................................42
3.5 Microstructural Characterization by SEM and EDS ............................................42

4. Results .........................................................................................................................44
4.1 Materials Characterization ...................................................................................44
4.1.1 Surface .......................................................................................................44
4.1.1.1 Surface Morphology ...................................................................44
4.1.1.2 Real Area of Contact ( Areal Fraction of Asperities ) ................51
4.1.1.3 Depth Profile of Alloying Elements ...........................................51
4.1.2 Microstructure ............................................................................................57
v

4.1.2.1 Cross sectional Morphology .......................................................57


4.1.2.2 Phase Identification of Coating Layers ......................................60
4.2 Surface Friction ..................................................................................................69
4.2.1 The Coefficient of Surface Friction under Drawing Conditions ...............69
4.2.2 The Coefficient of Surface Friction under Stretching Conditions .............72
4.3 Coating Adhesion Properties ..............................................................................78
4.3.1 Powdering under Drawing Conditions ......................................................78
4.3.2 Powdering under Plane Strain & Stretching Conditions............................79
4.4 Formability .........................................................................................................85
4.4.1 Limiting Dome Height Value ....................................................................85

5. Discussion ...................................................................................................................88
5.1 Factors affecting Surface Friction .......................................................................88
5.1.1 Surface Morphology ..................................................................................88
5.1.2 Strain State .................................................................................................90
5.1.3 Microstructure ............................................................................................92
5.1.4 Galvanizing and Galvannealing Conditions ..............................................95
5.2 Factors affecting Coating Adhesion .....................................................................99
5.2.1 Microstructure ............................................................................................99
5.2.2 Strain State ...............................................................................................103
5.2.3 Galvannealing Conditions ........................................................................107
5.2.4 Steel Substrate ..........................................................................................108
5.2.5 Surface Friction ........................................................................................111
vi

5.3 Formability ........................................................................................................112


5.3.1 Surface Friction .........................................................................................112
5.3.2 Coating Adhesion ......................................................................................116
6. Conclusions ..............................................................................................................118
7. Appendix 1 ...............................................................................................................123
Appendix 2 ...............................................................................................................124
Appendix 3 ...............................................................................................................125
8. References ................................................................................................................126

vii

LIST OF TABLES

Table 3.1

Galvannealing condition and mechanical properties for the STZC


Samples ................................................................................................... 23

Table 3.2

Mechanical properties for commercial IF-grade and IFP- grade .............23

Table 3.3

Mechanical properties for GI and GA material .......................................24

Table 4.1

Areal fraction of asperities at initial normal pressure( P ) = 0 .................52

Table 4.2

Atomic concentration, elemental ratio and zinc state ratio


observed by XPS at 5nm depth in the surfaces
of five different coated steel sheets .........................................................56

Table 4.3

Polynomial fits for punch load, P in kN, as a function of


contact radius, rc in mm ..........................................................................77

Table 4.4

Calculated coefficient of friction, ,under stretching condition ............77

Table 5.1

Predominant factors affecting surface friction under specific


conditions ................................................................................................97

Table 5.2

Calculated lattice misfit of interfacial layers [33,47,48].............................110

Table A.1

Prediction of the number of fields (n) to be observed


as a function of the desired relative accuracy and
of the estimated magnitude of the areal fraction of
the constituent ........................................................................................123

Table A.2

95% Confidence interval multipliers .....................................................123

Table A.3

Data used for calculation of ratio of a shear stress due to friction


to a shear yield strength of pure zinc coating and an interfacial
shear strength of pure zinc coating ........................................................124

viii

LIST OF FIGURES

Figure 2.1

Comparison between hard/hard contact and hard/soft contact .................7

Figure 2.2

Various static friction models : (a) Coulomb and Amontons model ;


(b) Bowden and Tabors model ; (c) Orowans model ; (d) Shaws
model ........................................................................................................8

Figure 2.3

Examples of static Friction (a) and dynamic Friction (b,c). Figure (a)
shows Coulomb friction which is independent of velocity. Figure (b)
shows there is a pre-sliding regime by microscopic motion before gross
sliding. Figure(c) shows both sliding velocity and frequency of velocity
oscillation lead a friction force. Figure (b) and (c) suggest that friction
is a function of either displacement or velocity........................................ 9

Figure 2.4

The Stribeck curve and the Stribeck effect by Stribeck model ...............11

Figure 2.5

Daltons friction model for coated steel sheet .........................................14

Figure 2.6

Trapped lubricant effect (a) no lubricant condition, (b) lubricant&


trapped condition; y = the shear yield strength of surface material ........15

Figure 2.7

Crack initiation of GA under the lateral tension condition .....................16

Figure 2.8

The mechanism of interfacial crack propagation ....................................17

Figure 2.9

The buckling failure mechanism .............................................................18

Figure 3.1

Fe-Zn diffusion couple and principle of microstructure control .............22

Figure 3.2

Transparent test grid sheet ......................................................................27

Figure 3.3

Schematic diagrams of cup test (A) and important parameters


in equation (3.2) (B) ................................................................................31

Figure 3.4

Preparations for cup test:(A-a) male die,(A-b) female die with


225mm diameter, respectively, (A-c) punch with 33mm diameter,
(B-a) 74mm blank, and (B-b) fully formed cup ......................................32

Figure 3.5

Generalized stretching forming operation [25] .......................................34

ix

Figure 3.6

Punch with 101.6mm diameter and die set with 225mm diameter and
132.6mm diameter draw bead for hemispherical punch test (A) and
important parameters in equation (3.4) (B) .............................................37

Figure 3.7

Radius of the boundary of punch contact ( rc ) is plotted as a punch


displacement, for three test materials under different test
conditions [25] .........................................................................................38

Figure 3.8

Relationship between major engineering strain vs. minor


engineering strain and limiting dome height vs. blank geometry ...........41

Figure 4.1

SEM plan view images (200 ) of the STZC1 steel (a) and STZC2
steel (b) ....................................................................................................47

Figure 4.1

SEM plan view images (200 ) of the STZC3 steel (c) and GI steel
(d) ............................................................................................................48

Figure 4.1

SEM plan view images (200 ) of the GA steel (e) ................................49

Figure 4.2

SEM images of pattern (asperity) on the (a) STZC1, (b) STZC2, (c)
STZC3 (d) GI and (e) GA steels. Yellow arrow indicates surface
asperity ....................................................................................................50

Figure 4.3

Real area of contact of five coated steels at initial normal pressure=0 ...52

Figure 4.4

XPS (ESCA) Depth Profile of the (a) STZC1, (b) STZC2, (c) STZC3,
(d) GI and (e) GA steel coatings .............................................................55

Figure 4.5

Cross sectional images of five types of coating layer captured by


back scattered electron mode in SEM (2000 ) : (a) STZC1 (b)
STZC2 (c) STZC3 (d) GI and (e) GA steels. Arrows from (a) to (e)
indicate an Fe-Zn intermetallic phase formed through the thickness
and its average vertical height. Circles and ellipse in (e) indicate
vertical cracks and horizontal crack, respectively. ................................59

Figure 4.6

EDS element mapping through the thickness of the STZC 1 steel


(a) and energy spectrum (b) and chemical composition (at.%)
(c) from an Fe-Zn intermetallic phase (yellow arrow) in the steel
coating ....................................................................................................63

Figure 4.7

EDS element mapping through the thickness of the STZC 2 steel


(a) and energy spectrum (b) and chemical composition (at.%)
(c) from an Fe-Zn intermetallic phase (yellow arrow) in the steel
coating ....................................................................................................64

Figure 4.8

EDS element mapping through the thickness of the STZC 3 steel


(a) and energy spectrum (b) and chemical composition (at.%)
(c) from an Fe-Zn intermetallic phase (yellow arrow) in the steel
coating ....................................................................................................65

Figure 4.9

EDS element mapping through the thickness of the GI steel (a)


and energy spectrum (b) and chemical composition (at.%) (c)
from a pure zinc coating (yellow arrow) on the steel ............................66

Figure 4.10

EDS element mapping through the thickness of the GA steel (a)


and energy spectrum (b) and chemical composition (at.%) (c)
from a fully galvannealed zinc coating (yellow arrow) on the steel ......67

Figure 4.11

Fe-Zn-Al ternary phase diagram (isothermal section at 500C) [32] ....68

Figure 4.12

Punch load as a function of hold down load and punch displacement


of the (a) STZC1, (b) STZC2, (c) STZC3, (d) GI and (e) GA steels.....74

Figure 4.13

Coefficient of surface friction calculated as a function of punch


displacement for cups drawn from STZC1, STZC2, STZC3, GI
and GA steels .........................................................................................75

Figure 4.14

Punch load as a function of punch displacement and radius of


contact boundary of the (a) STZC1, (b) STZC2, (c) STZC3, (d)
GI and (e) GA steels ...............................................................................76

Figure 4.15

Powdering weight under the condition of drawing of the (a) STZC1,


(b) STZC2, (c) STZC3, (d) GI and (e) GA steels and (f) a comparison
of the powering rate in the five zinc coatings ........................................83

Figure 4.16

Powdering weight calculated as a function of blank width from LDH


tests of the (a) STZC1, (b) STZC2, (c) STZC3, (d) GI and (e) GA
steels and (f) a comparison of the powdering rate in the five zinc
coatings ..................................................................................................84

Figure 4.17

LDH curves for the (a) STZC1, (b) STZC2, (c) STZC3, (d) GI and
(e) GA steels and (f) curves normalized by setting each plain strain
condition as an origin .............................................................................87

Figure 5.1

Effect of surface morphology on surface friction ..................................89


xi

Figure 5.2

Effect of different strain state on surface friction. Note GI: galvanized,


GA: galvannealed, P-GA: galvannealed and phosphate treated, and
Bare: non-coated (Source: present study [0], Paik [24], Ghosh [25],
Fox et al.[39], Siegert et al. [40], Hira et al.[41] ) .................................91

Figure 5.3

A microstructure of zinc coating where an Fe-Zn intermetallic phase


was grown ideally ...................................................................................94

Figure 5.4

Effect of average vertical height of Fe-Zn intermetallic phase on


surface friction ........................................................................................97

Figure 5.5

Effect of galvannealing temperature on shear yield strength


of zinc coating [43, 44, 45] ..........................................................................98

Figure 5.6

Effect of average vertical height (a) and type (b) of Fe-Zn


intermetallic phase on coating adhesion for a drawing strain ................101

Figure 5.7

The Fe-Zn intermetallic phases and surface oxides acting as initial


crack sites ...............................................................................................102

Figure 5.8

Effect of strain state on coating adhesion in phase based (a) and


phase based coating (b) based upon strain states generated during

cup testing and LDH testing ..................................................................105


Figure 5.9

Schematic illustration of crack propagation of phase based- and


phase based coating as fracture driving strains are subjected
(Red line: crack initiation sites, blue line: crack propagation) ..............106

Figure 5.10

Effect of steel substrate on coating adhesion in the STZC steels ..........110

Figure 5.11

Effect of surface friction on formability in various zinc coated steel


sheets .....................................................................................................113

Figure 5.12

Effect of a modification of strain state by frictional constraint on


formability .............................................................................................115

Figure 5.13

Effects of coating layer and steel substrate on formability; a) effect


of average vertical height of Fe-Zn intermetallic phase, b) effect of
steel substrate (IF and IFP *) .................................................................117

xii

ACKNOWLEDGEMENTS

The author would like to thank Professor Gary M. Michal for giving this research
opportunity and lending his insight and experience for long time to this project. His
guidance and support have always encouraged me to overcome various challenges. I am
proud to have the opportunity to work with him. Also, I would especially like to thank Dr.
Doojin Paik for all of his time and effort while being an invaluable help through the
length of this work. My thanks also to CMCM staff, Christopher J. Tuma, SCSAM staff,
Reza Sharghi and graduate colleague, Joshua Katz for all of their advice and help.

I would like to thank my wife, Yewon Lee and lovely son, Josiah Jang who kept on my
side and my parents who prayed for my ways.

Finally, I would like to glorify my god who guided my life and provided all supports for
my needs throughout this pursuit of a masters degree.

xiii

The Tribology and Formability of Zinc Coated Steel Sheets Subjected to Different Strain
States

Abstract
by

YOHAN JANG

Five types of zinc coated sheet steels were evaluated for their surface friction,
coating adhesion properties and formability under different strain states. The coefficient
of friction values found under biaxial tensile strain were significantly larger than those
associated with the drawing strain states. It was found that surface morphology, strain
state, and shear yield strength of the zinc coating were direct factors affecting surface
friction. The - phase based coatings exhibited the lowest coating adhesion either at the
plane strain condition or at the stretching condition whereas the - phase based coating
exhibited that at the drawing condition. It was found that coating adhesion depends
directly on the microstructure of the zinc coating and the strain state. It was also found
that surface friction and coating adhesion could affect the formability of a zinc coated
steel sheet by influencing its strain state and fracture mechanism, respectively.

xiv

1. Introduction
Zinc coated sheet steels such as galvanized or galvannealed are gaining
importance in the automotive industry due to their good drawability and corrosion
resistance. The steel with a thin zinc layer can be produced by passing the steel through a
molten bath of zinc at a temperature of around 460 C and various kinds of heat
treatments through a cascade of furnaces can be applied after this. The coated sheet steels
consist of a coating layer (~10m thick on each side of the substrate) and steel that is
nominally 0.7mm thick. In spite of its modest thickness compared to the substrate, the
coating layer has a critical role in the whole performance of the coated sheet steel. It is
well known that the zinc based coating layer increases corrosion resistance because zinc
is anodic with respect to iron, the zinc acts as a sacrificial anode to protect the substrate
from corrosion. Another important role of the zinc layer which is not as well known is its
ability to increase formability. The outer zinc coating layer serves as a thin film of soft
metal on the substrate and it avoids hard metal to hard metal contact which greatly
increases friction. Usually, the coating layer increases the formability of the steel sheet by
lowering its surface friction. Additionally, the coating layer can improve paintability and
weldability through post heat treatments [1,2].
Formability is a key characteristic in the application of zinc coated steel sheet for
processes which avoid machining such as automotive panels, the outer body of
appliances, and parts for aircraft. Two key properties that contribute to formability are
surface friction and coating adhesion. The lower the surface friction, the higher the
formability that is possible. Surface friction is a characteristic of a particular system, used
1

to measure friction. Thus, it is very important to understand how friction is being


assessed and what system factors affect friction for different kinds of forming operations.
Coating adhesion affects formability and it depends upon the characteristics of the
coating layer. Usually, the better the coating adhesion, the higher the formability that is
possible. In the case of heat treated zinc coated steel sheets such as galvannealed steel,
the coating layer has several intermetallic phases and these can be fractured easily and
detached from the coating during a sheet forming operation. The process of brittle
fracture leads to mass loss commonly referred to as powdering. It results in reducing the
coating quality, causing damage to the forming tools or the steel sheet itself, and as a
result the lowering of formability. Coating adhesion should be taken into account to
optimize formability and it is very important to understand how the fracture takes places
on the substrate and what factors affect mass loss or powdering.
The objectives of this work are to evaluate surface friction and coating adhesion
of differently treated zinc coated sheet steels subjected to different strain states, to
analyze factors correlated with them, and subsequently to define optimal conditions in
terms of surface friction and coating adhesion for good formability
To do this, it is first required to understand how friction can be measured in major
sheet metal forming processes and what kind of coating fracture mechanisms occur
during such processes. A number of reviews of previous studies of surface friction
models and fracture mechanisms are presented as a part of the background in chapter 2.
Chapter 3 includes details associated with the experimental procedures. The
information regarding the materials used in this work is listed in section 3.1. The method
2

used to quantify the surface morphology of each material is described in section 3.2.
Several kinds of mechanical tests used for the measurement of friction, coating adhesion,
and formability are described in section 3.3. Several methods for microscopic surface
analysis and microstructural characterization are presented in section 3.4 and section 3.5,
respectively.
Chapter 4 includes the results of experiments described in chapter 3. The five zinc
coated materials are characterized in terms of both their surface and microstructural make
up in section 4.1. The surface frictional behavior of the five materials subjected to
different strain states is covered in section 4.2 and the coating adhesion behavior of the
materials subjected to different strain states is contained in section 4.3.
Chapter 5 discusses key issues uncovered through this study. Several factors
affecting surface friction and coating adhesion, empirical relationships among factors,
and fracture models for the five zinc coatings are discussed in section 5.1. The relevance
of surface friction and coating adhesion to formability, one of the biggest themes in this
study, is discussed in section 5.2.
Finally, chapter 6 presents the conclusions of this study about surface friction and
coating adhesion of zinc treated steel sheets subjected to different strain states and their
effects on formability.

2. Background
The friction between sheet metal and various tools is one of the most important
factors affecting the formability as well as the success of a sheet metal forming operation.
In this chapter, friction will be defined and categorized with models proposed from
various perspectives.
Friction is the tangential reaction force between two surfaces in contact. In other
words, if two bodies are placed in contact under a normal load, a finite force is required
to initiate or maintain sliding. This is the force of friction. This force is the result of many
different mechanisms, which depend on contact geometry and topography, properties of
the bulk, the material at the surfaces of the bodies, displacement, relative velocity of the
bodies, and the presence of lubrication. Many models have been developed to take into
account these various sources of friction. Important static friction models and dynamic
friction models that relate to sheet metal forming will be presented.
2.1 Static Friction Theory
2.1.1 Coulomb and Amontons Friction Theory
Static friction is the friction associated with sticking. It describes the friction force
at rest. The concept is opposed to dynamic friction. The early work concerning this form
of friction was done by Coulomb and Amontons. The main idea is that the frictional force
is independent of the area of contact of the sliding bodies and velocity (Coulombs law)
and proportional to the applied load (Amontons law). In general, contact occurred only
locally at the summit of the surface irregularities so that the real contact is very small and
the frictional force is almost independent of the apparent area of contact, but still
4

dependent on the real contact area. After Coulomb and Amontons had suggested the
postulates of frictional force, a conductivity measurement by Tabor proved their
postulates [3]. According to Tabors studies, the deformation of the surfaces is not elastic,
but mainly plastic and the metal flows under the applied load until the area of contact is
sufficient to support it. In the case of plastic flow, the area of contact, i.e. real contact
area, is directly proportional to the applied load, so that the frictional force is directly
proportional to the applied load. It is clear that the frictional resistance is determined
primarily by the real area of contact and that the load is important only insofar as it
affects this area from Coulomb, Amontons and Tabors friction theory. The Coulomb
friction can be described as
FC = FN

(2.1)

where FC is coulomb friction force, FN is normal load and is coefficient of friction.


2.1.2 Bowden and Tabors Friction Theory
After Coulomb and Amontons work, Bowden and Tabor investigated the
deformation of surface asperities using the theory of plasticity [4]. They compared the
deformation of a single asperity with hardness testing, and obtained a very useful model
of the friction mechanism even if the influence of tangential stresses, interaction between
asperities, strain hardening, and elastic effects were neglected. According to their model,
the asperities each carry a part Fi of the normal load FN . If the asperities start to be
deformed plastically when the contact area of each junction has grown large enough to
carry its part of the normal load, the contact area of each asperity junction is ai =

Fi
,
H

where H is the hardness of the weakest bulk material of the bodies in contact and the unit
for H is the Pascal. The total contact area ( Ar ) can thus be written=
Ar

F .
a
=
H
i

This relation holds even with elastic junction area growth, provided that H is adjusted
properly. For each asperity contact the tangential deformation is elastic until the applied
shear stress exceeds the shear strength y of the surface materials. The friction force, i.e.
plastic deformation forces of microscopical asperities in contact, is defined as FC = y Ar .
As a result, the friction coefficient can be expressed as =

FC y Ar y
. According
=
=
FN Ar H H

to this model, the friction coefficient is not dependent on the normal load or the velocity,
but on the properties of the surface material in the model.
The model implies that it is possible to manipulate frictional characteristics by
deploying surface films of suitable materials on the bodies in contact. As such a soft
coating layer on a coated sheet steel can be very important for decreasing the frictional
force. Figure 2.1 shows schematically how this is possible. (a) is hard/hard metal contact
and (b) is soft thin film layer/hard metal contact. For the case of (a), the high shear
strength of the hard metal makes the frictional force significant. However, in the case of
(b), the shearing takes place on the thin film which has a low shear strength. That is why
a coated steel sheet could have a low frictional force compared to non-coated steel sheet.
Even though the Bowden and Tabor model is close to understand the frictional
mechanism of coated steel sheets, it has still many limitations.

Figure 2.1: Comparison between hard/hard contact and hard/soft contact

2.1.3 Orowan and Shaws Friction Theory


Since Bowden and Tabor (Fig.2.2-b) established the adhesion theory based upon
the individual plastic deformation of the contact points, many authors have developed
new friction theories concerning metal forming processes using a similar idea to focus on
the deformation of surface asperities using the theory of plasticity. Whereas the friction
model from Bowden and Tabor is confined to frictional behavior at low normal stress,
Orowan (Fig.2.2-c) and Shaw (Fig.2.2-d) developed friction models at high normal
stresses. Because some metal forming operations such as forging, extrusion and
stretching occur under high normal stress and the zones from different asperities can
interact due to the high stress, a different model should be applied to understand frictional
behavior. They suggested that the friction stress ( c ) at low normal stresses is
proportional to the normal stress ( N ), but at high normal stress is equal to the yield stress
in pure shear ( y ) of the surface materials [5,6]. This implies that normal stress no longer
affects friction at sufficiently high stress because the ratio between real area of contact
and apparent area increases with increasing pressure and approaches one.
7

Figure 2.2: Various static friction models : (a) Coulomb and Amontons model ; (b) Bowden and
Tabors model ; (c) Orowans model ; (d) Shaws model

2.2 Dynamic Friction Theory


Steel sheets experience various displacements within a die during metal forming
operations such as drawing and stretching. This means that static friction is no longer
valid in practice. Many studies have been started to understand the nature of dynamic
friction since the 1950s [Fig.2.3]. The concept is that surface friction is a function of
displacement or velocity. Rabinowicz studied the transition between sticking and sliding
in his early work [Fig.2.3(b)] [7] and found that the maximum frictional force typically
occurs at a small displacement from the starting point. Dalton proposed that the
coefficient of surface friction is a function of time or distance in his studies of the
frictional behavior of zinc-coated steel sheet [8]. Stribeck addressed the velocity
8

dependence of the frictional force in the early 19th century [9]. Hess et al. performed
experiments on the velocity dependence of friction with a periodic time-varying velocity
superimposed on a bias velocity and observed an unsteady sliding under high frequency
of velocity oscillation causes a multi-valued friction during acceleration and deceleration
[Fig.2.3 (c)] [10].

Figure 2.3: Examples of Static Friction (a) and Dynamic Friction (b,c). Figure (a) shows
Coulomb friction which is independent of velocity. Figure (b) shows there is a pre-sliding regime
by microscopic motion before gross sliding. Figure (c) shows both sliding velocity and frequency
of velocity oscillation lead a friction force. Figure (b) and (c) suggest that friction is a function of
either displacement or velocity.

2.2.1 Stribeck Friction Theory


Usually, sheet steel is formed with the aid of a lubricant. The understanding of the
effect of lubrication on surface friction is one of the most important aspects of practical
metal working. Based on friction experiments on bearings, Stribeck expressed the
relationship between the friction coefficient [ f ], viscosity of the lubricating oil [ ],
normal load [ FN ], and velocity [ V ] in the Stribeck curve. According to his model, a
material undergoing a forming operation experiences three distinctive lubrication states
based upon increase in a factor, V / FN (oil viscosity sliding velocity / normal load)
known as Stribeck parameter. In hydrodynamic lubrication regime where the V / FN
factor is high, the fluid lubricant completely isolates the friction surfaces and internal
fluid friction alone determines the tribological characteristics. In other words, friction is
related to only viscous drag forces in the lubricating film. In this state, the friction
coefficient increases as the V / FN factor increases because the shear force of the fluid
increases due to the film thickness increase. In the elastohydrodynamic lubrication
regime where the V / FN factor is low, i.e. at low velocities and high pressure, the thin
lubricant is transformed into an amorphous solid phase due to the high pressure (pressure
effect) and an interaction between friction surfaces is significant (velocity effect). The
transformation of lubricant and significant interaction between surfaces lead to a high
friction coefficient because shear forces corresponding to a combination of solid lubricant
and solid surface are much higher than those corresponding to fluid lubricants. As a result,
the friction coefficient decreases as the V / FN factor increases in this state. In the
10

boundary lubrication regime where the V / FN factor is very low, lubricant thickness is
at a minimum and surfaces are in contact at microasperities. The friction coefficient is at
its highest value in this state because lubricant is not effective. Figure 2.4 shows the
Stribeck curve and the Stribeck effect which corresponds to the dip in the frictional force
at low velocities.

Figure 2.4: The Stribeck curve and the Stribeck effect by Stribeck model

11

2.2.2 Daltons Friction Theory


Recently, Dalton has suggested new surface friction models for sheet metal
forming [8]. Daltons model is the most practical and applicable one because it is a
hybrid model that includes possible events, which take care of certain aspects of the
friction force. It covers four cases with several factors, which are the sheet coating, tool
type, roughness, and lubricant additives. Figure 2.5 shows the new model schematically.
Region1 and 2 show a mixed film plateau. In this region, the lubricant forms a thin film
and this phenomenon has already been described by Stribeck. Model A is the case of
increasing surface friction and the contact of asperities is significant. In model B1 surface
friction stabilizes and then increases. The supply of surface film or lubricant is exhausted
in this region and stick-slip eventually occurs. In model B2 separation of surfaces allows
a decrease of surface friction. In this region surface damage is high. Model C is the case
of the formation of pressurized pockets by oil trapping as asperities continue to deform.
Surface friction is gradually decreased as the forming process continues. This model is
one of the ideal cases for metal forming. Wanheim and Bay [11] suggested that the
presence of trapped lubricant would result in a superimposed hydrostatic pressure acting
on the valleys of the asperities and could reduce friction in metal working processes
[Fig.2.6]. According to their theory, the lubricant pressure equals:

K =K 0 + K1PF + ......

PF 1

= (1 2 ) K1 1
K 0 K1
h0

12

(2.2)

where K is the bulk modulus of the lubricant, K 0 is the initial bulk modulus of the
lubricant, K1 is a unitless coefficient of pressure sensitivity , PF is the hydrostatic
compression applied, is the displacement of an asperity after deformation, and h0 is the
height of an asperity. As a result, the effect of trapped lubricant equals:

qtotal PF K 0 qdry
=
+
K 0 2k 2k
2k

(2.3)

where qtotal is the total pressure necessary to obtain a certain contact area, k is the bulk
modulus of the asperity, and qdry is the pressure necessary to obtain a certain contact area
without trapped lubricant. The first term corresponds to a contribution by lubricant
pressure and the second term corresponds to a contribution by applied pressure. Once
superimposed hydrostatic pressure is applied by trapped lubricant and the total normal
pressure required is constant, qdry decreases compared to the no lubricant trapped
condition. Because qdry is directly related to the real area of contact, a large reduction in
contact area could be achieved by trapped lubricant. It is evident that friction decreases as
real contact area decreases.
Daltons model and Bowden & Tabors model have directly contributed to the
understanding of the frictional behavior of coated sheet steel during forming operations.
Based on their fundamental model, extensive work has been done to modify the surface
coating layer and various lubricant have been used in order to obtain a low friction
coefficient, i.e. high formability. The frictional behavior of special treated zinc coated

13

material, one of the candidates to show high formability will be discussed in subsequent
chapters.

Figure 2.5: Daltons friction model for coated steel sheet

14

2.3 Fracture Mechanics of Coated Steel Sheet


Another key property that contributes to the formability of zinc coated steel sheets
is coating adhesion and fracture. Usually, heat treated coated steel sheets have a
deficiency in that the coating layer tends to powder and fall off when undergoing severe
press forming. The process of brittle fracture is referred to as powdering. It is possible
that powdering acts as a source of cracks which could induce local stress concentrations
and thus affect formability. Coated steel sheets have several intermetallic phases formed
after galvannealing. These phases act as the main source of powdering. The fundamental
fracture mechanisms associated with the coating layer will be discussed in this section.

Figure 2.6: Trapped lubricant effect (a) no lubricant condition, (b) lubricant& trapped
condition ; y = the shear yield strength of surface material
15

2.3.1 Wedging Mechanism Causing Powdering


A. T. Alpas et al. [12] proposed that a decohesion of the interfaces between
substrate and coating could be followed by a propagation of through thickness cracks
under uniaxial compressive stress condition when the strength of the substrate-coating
boundary is higher that the intrinsic strength of the coating. This is referred to as the
wedging mechanism. The powdering phenomenon from galvanneal steel sheet (GA) is a
very good example of the wedging mechanism. The coating layer of GA may consist of a
thick layer of phase ( FeZn10 , 86.5~91.8 at.% Zn ), a thin layer of phase ( Fe3Zn10,
68.5~82.5 at.% Zn ), columnar crystals of 1 phase (Fe11Zn40 ,75~81 at.% Zn ) and
blocky crystals of (FeZn13, 92.8~94 at.% Zn). Initially, cracks are not created in the
interface. Many pre-existing cracks nucleated during cooling of the coated steel sheet
from the galvannealing temperature to room temperature. These grow in the phase as it
is subjected to a tensile stress [Fig.2.7]. Once the cracks are propagated adjacent to the
interface between coating and substrate under lateral tension, a crack propagation along
the interface can be triggered under lateral compression. The decohesion along the
substrate- coating interface under lateral compression directly results in powdering.

Figure 2.7: Crack initiation of GA under the lateral tension condition


16

Alpas et al. [12] stated that the driving force for interface crack propagation is an
inhomogeneous distribution of residual stress caused by the formation of the 1 phase
with a columnar morphology. Usually, the 1 phase is formed over the planar phase
and grows into the phase by forming a columnar morphology. The crystals cause
residual compressive stress and generate interfacial shear stresses along the -1 and -
phase boundaries. As a result, a crack can easily propagate through the interface when
subjected to lateral compression [Fig.2.8].

Figure 2.8: The mechanism of interfacial crack propagation

2.3.2 Buckling Mechanism Causing Powdering


Another mechanism by which a coating can fail under lateral compression is
buckling. Buckling failure occurs if the strength of the interface is less than that of the
coating itself so that decohesion of the substrate-coating interface precedes the formation
of the through thickness cracks in the lateral tension condition [Fig.2.9]. There are
17

various sources that can initiate buckling. In the case of hot dip galvanizing or electro
galvanizing, the substrate can be exposed to an unclean environment just before the
coating layer is deposited. The deposited defects on the substrate can reduce adhesion
energy between the substrate and coating layer equivalent to roughly 150J/m2 in the case
of a zinc-steel interface. During the electro galvanizing process,
some amount of hydrogen can be injected between the substrate and the coating layer.
S.L. Amey et al. [13] showed that approximately 2 109 moles / cm 2 of hydrogen could
be absorbed into the steel in a commercial electro galvanizing line. A region debonded by
hydrogen could exhibit buckling when subjected to lateral compression.

Figure 2.9: The buckling failure mechanism

Janavicius P. V. et al. [14] modeled the growth of an initially buckled area and
derived an expected critical pressure equation as follows.
1/2

EGa
Pcrit =
as a t
2
2a (1 )

18

(2.4)

1/2

32Ga
Pcrit
=
2
3(1

a 3
4 a
a +

t 1 t

1/2

as a = t

(2.5)

1/4

Pcrit

G 3 Et
= 4.8 a 4
a

as a t

(2.6)

where a = radius of debonded region, E = elastic modulus of layer, t = thickness of layer


and Ga = adhesion energy. According to the model, the growth of an initially delaminated
area in the zinc coating depends on the size of the initial delamination, the adhesion
energy between the zinc and its steel substrate surface and the thickness and the elastic
modulus of the zinc layer. The thickness and elastic modulus of the zinc layer cannot be
varied significantly in practice. However, the size of the initial delamination and the
adhesion energy can vary substantially. The presence of irregularities in the steel
substrate can greatly influence both of those parameters. Such irregularities could include
microflaps, carbides, oxide inclusions not completely removed after the rolling process.
This means that the exposure of the substrate to an unclean environment before
galvanizing could affect both crack initiation and propagation according to the buckling
fracture mechanism.

19

3. Experimental Procedure

3.1 Material Properties


Special treated zinc coated (STZC) materials were tested in the current work.
These materials were produced by Pohang Iron and Steel Company and are material used
for automotive panel applications. The STZC steels were designed to investigate the
effects of surface morphology and an Fe-Zn alloy layer on the formability of coated steel
sheets. In addition to the three STZC steel sheets, conventional galvanized (GI) and
galvanneal (GA) coated steel sheets were also tested for comparison. Usually, galvanized
steel sheet has a low powdering rate and a high surface friction coefficient because its FeZn alloy layer is very thin so its coating does not exfoliate during forming operations and
the zinc coating layer on the steel sheets surface is continuous resulting in a contact area
that is very large. Thus, galvanized steel can be used as a standard for a high powdering
resistance and high friction material. Galvanneal steel sheet has a high powdering rate
and a low friction coefficient because its Fe-Zn alloy layer is mostly composed of phase
which is relatively thick and the alloyed coating surface morphology is in the form of
islands or mesas so that its contact area is very low. Thus, galvanneal steel can be used as
a standard for a low powdering resistance and low friction material during forming.
3.1.1 Special Treated Zinc Coated (STZC) Materials
The materials were modified differently in terms of surface texture and
microstructure. The surface texture is one of the key parameters that determines surface
friction and affects formability. After completing the galvanizing process, three types of
20

patterns were imprinted into STZC1, STZC2 and STZC3 via working rolls that were
patterned by an Electrical Discharge Texturing (EDT) method. Microstructure is one of
the key parameters that determines the powdering rate which in turn affects formability.
The main parameters that control the microstructure of a coating are galvannealing
temperature and time. Temperature controls diffusion length to a greater extent than time.
Thus, temperature was controlled to modify the thickness of the Fe-Zn alloyed layers.
STZC1,STZC2 and STZC3 were galvannealed in 452C, 465C and 516C , respectively
[Fig.3.1]. The Fe-Zn phases initially nucleate at the steel/coating interface and consume
the pure zinc phase, when the steel passes through the galvannealing furnace. The
initial Fe-Zn intermetallic compound grows rapidly. The phase grows at the expense of
the phase. Later, the phase can be consumed by the growing phase [15]. It is well
known that the brittle phase decreases the formability of galvannealed steel sheet. The
galvannealing time was controlled by the line speed of the rolling process and the short
time in the range of 1~5 seconds was designed to prevent transformation of phase to
phase. The galvannealing condition and mechanical properties for the STZC steels are
listed in Table 3.1. The substrates of the STZC materials are an interstitial free steel (IF
steel grade) and an interstitial free rephosphorized steel (IFP steel grade). Interstitial free
steels are a class of high formable steel with exceptionally low levels of interstitial
elements such as carbon and nitrogen (typically, total C 30ppm and total N 40ppm by
weight) and are mainly used for automotive application due to their high drawability [16].
Interstitial free rephosphorized steel provides a high yield and tensile strength. It has a
carbon content similar to an IF steel grade, but a much higher phosphate content
21

( 630ppm) leading to an increased yield strength and ultimate tensile strength due to
grain boundary hardening [17]. The rephosphorized steel grade was applied to only the
STZC3 material in order to evaluate how the substrate in a coated steel sheet affects its
total formability. The mechanical properties of a typical commercial IF-grade steel and a
IFP-grade are given in Table 3.2.

Figure 3.1: Fe-Zn diffusion couple and principle of microstructure control

22

Table 3.1 Galvannealing condition and Mechanical properties for the STZC samples

Material

Grade

STZC1
STZC2
STZC3

IF
IF
IFP

Yield
GA heating
Strength
temperature(C)
(MPa)
452
465
516

160
152
264

Ultimate
Tensile
Strength
(MPa)
306
310
396

Elongation
(%)

Line
Speed
(mpm)

45
46
35

110
138
120

Table 3.2 Mechanical properties for commercial IF-grade and IFP- grade

Substrate
IF
IFP

Ultimate
Tensile
Strength
(MPa)
150~200 300~350
260~320 350~410
Yield
Strength
(MPa)

ef (%)
35
23

3.1.2 Hot- Dip Galvanized (GI) Material


The galvanized steel sheet was intended to have the highest formability possible.
Basically, the interstitial free grade steel which has good drawability was used for the
substrate of the GI. Usually, the GI has a thin intermediate layer called the inhibition
layer between the steel substrate and the zinc coating. The layer consists of an Al-Fe
compound and acts as a diffusion barrier to suppress brittle intermetallic Fe-Zn
compound formation which affects negatively the adhesion quality of the coating layer
and subsequent formability [18]. To process the GI a high aluminum content was present
in the zinc bath in order to promote the formation of the inhibition layer. Mechanical
properties for the sheet are given in Table 3.3. After being coated with zinc, the material
was temper rolled.
23

3.1.3 Galvanneal (GA) Material


The galvanneal steel sheet was intended to have the lowest formability. The
interstitial free steel strip immersed in a zinc bath was immediately heated to 505C and
held for 20 sec in a galvannealing furnace in order to form a fully phase coating layer
containing around 10 wt. % Fe. The Al content of the zinc bath was maintained at a low
level of 0.126 wt. % to produce a relatively thin inhibition layer. After the galvannealing
process, the material was temper rolled. The mechanical properties for the GA sheet steel
are listed in Table 3.3.
Table 3.3 Mechanical properties for GI and GA material

Material

Grade

GI
GA

IF
IF

Yield
GA heating
Strength
temperature(C)
(MPa)
505

156
146

Ultimate
Tensile Elongation
Strength
(%)
(MPa)
291
47
294
47

Al content
in the zinc
bath (%)
0.236
0.126

3.2 Measurement of Real Area of Contact


The real area of contact is made up of a large number of small regions of contact,
called asperities or junctions of contact, where atom-to-atom contact takes place as
Bowden and Tabor suggested in their inelastic adhesive theory [3]. The real area directly
affects the surface friction of a coated steel sheet [11]. The droplet like bumps in the
STZC steels and the GI steel imprinted by the texturing process creates what are
effectively asperities. In the case of the GA steel, islands or mesas are effectively the
asperities. Under the assumption that the change of real contact area is not significant

24

during a metal forming process, the initial real contact area, i.e. the sum of asperities was
calculated by a simple statistic method, called Systemic Manual Point Count [19].
3.2.1 Systemic Manual Point Count Method
The method is based upon the stereological principle that a grid with a number of
regularly arrayed points, when systematically placed over an image of a two dimensional
section through the microstructure, can provide, after a representative number of
placements on different fields, an unbiased statistical estimate of the volume fraction of
an identifiable constituent or phase. The method also can be used to measure the areal
fraction of asperities.
The systemic manual point count method requires a test grid and images obtained
from a scanning electron microscope. The test grid is in the form of a transparent sheet
where a specified number of equally spaced points formed by the intersection of very thin
lines is imprinted [Fig.3.2]. The scanning electron microscope provides excellent depth of
field and magnification high enough to adequately resolve the microstructure. It is
recommended to choose a magnification that gives an average constituent size that is
approximately one half of the grid spacing. The number of images or fields depends on
the desired degree of accuracy for the measurement. Table A.1 gives a guide to the
number of fields or images to be counted as a function of total number of points in the
test grid (PT), the relative accuracy which means a measure of the statistical precision,
i.e.=
% RA

95%CI
100 , and the magnitude of the areal fraction.
PP

The measurement procedure is as follows. The test grid which has a total of 100

25

points was superimposed upon an image captured at 200 magnification in a scanning


electron microscope. The number of points ( Pi ) falling within the microstructural
constituent of interest, i.e. asperities or mesas was counted and averaged for 20 fields.
The average number of points expressed as a percentage ( PP ) is an unbiased statistical
estimation of the areal fraction of the asperities or mesas, i.e. real contact area. After five
values of PP were obtained, the precision of the areal fraction measurement was
determined by calculation of the 95% confidence interval. The equations for calculating
these values are as follows:

PP =

1 n
PP (i)
n i =1
1/2

1 n

[ Pp (i ) Pp ]2
s
=

n 1 i =1

s
95% CI = t
n

(3.1)

where Pp (i ) is the percentage of grid points in field i , s is estimate of the standard


deviation, 95%CI is 95% confidence interval, i.e. PP 95%CI , t is 95% confidence
interval multiplier and n is number of fields. Table A.2 lists 95% confidence interval
multipliers (t).

26

Figure 3.2: Transparent test grid sheet

3.3 Mechanical Testing


Mechanical testing was conducted to determine how frictional and powdering
properties affect the formability of coated steel sheets. Surface friction coefficients were
measured by cup tests and hemispherical punch tests. These tests evoke drawing strain
states and stretching strain states, respectively. The deformed cups were subsequently
used to assess the degree of powdering by measuring the quantity of coating loss. Finally,
the plane strain formability was evaluated using a limiting dome height test.
3.3.1 Measurement of Surface Friction Coefficient
Punch-to-sheet friction plays an important role in determining formability and

27

subsequently determining the success or failure of sheet metal forming operations. The
friction depends strongly on what kind of forming conditions are operative. The friction
values need to be measured under different forming conditions. The cup test was
designed to determine surface friction values under conditions corresponding to drawing
strain states. The hemispherical punch test was designed to determine surface friction
values under conditions corresponding to biaxial stretching strain states.
3.3.1.1 Cup Test
In a deep drawing operation, friction exists in three regions. These regions are the
flange, the die radius, and the punch profile radius. The regions have different frictional
conditions. Throughout the flange region, in which the sheet is held with a normal force
exerted by a blankholder, sliding friction is predominant, and small amounts of strain are
present. In the die radius and punch profile regions, sliding is negligible, and stretching
due to the local plastic deformation of the sheet prevails [20]. The sliding friction in the
flange region is what needs to be measured in this test because the friction in the flange
region determines the ease of supply of material due to plastic flow and subsequently
drawability in a deep drawing operation [21].
Based upon modeling of the forming loads during drawing of a cup, the
relationship between maximum forming load, hold down load and sliding friction
coefficient has been developed. According to K. Langes Handbook of Metal Forming
[22], one of the relationships is that maximum forming load or drawing load is a linear
function of the hold down force and the coefficient of sliding friction is a simple multiple
of that slope. Swift et al. [23] found that the contact angle between the punch and the
28

drawn blank and the instantaneous blank radius are functions of the punch displacement
or forming load.
As a result, an equation (3.2) for the forming load, P , which takes into account
drawing stress (1), stress due to friction in the flange, stress due to friction over the die
radius (2) and stress due to bending and unbending (3) was obtained from the continuum
models by Swift et al.

X FN
t
P 2 Rmt 1.1 fm1 ln
=
+

exp( ) + fm 2
2r0
Rm Xt

(1)

(2)

(3.2)

(3)

In equation 3.2 is the coefficient of surface friction, is the contact angle, X is the
instantaneous blank radius, t is the blank thickness, Rm is equal to the punch radius plus
0.75t , FN is the hold down load, r0 is die radius, fm1 is the average flow stress in the

flange area of the blank, and fm 2 is the average flow stress when the blank undergoes
bending and unbending. The contribution to the forming load of fm1 and fm 2 , which
depend on the work hardening coefficient ( n ), is very small compared to that of the
coefficient of surface friction ( ). The coefficient of surface friction can be obtained by
taking the derivative of P with respect to FN resulting in equation (3.3). The value of
is determined through this transcendental equation by measuring the slope of the forming
load versus hold down load for particular values of the parameter X and .
dP 2 Rm
=
exp( )
dFN
X

29

(3.3)

The coefficient of surface friction, , for each of the five types of steel sheets
was determined through cup tests [Fig.3.3-A]. Initially about thirty blanks (74mm dia.)
were sheared from each type of steel using a shearing punch and hold down die set
mounted on a dual actuator hydraulic deformation press [Fig.3.4-B]. The punch and die
set was then replaced by a cup forming punch and male and female die set on the same
dual actuator hydraulic press [Fig.3.4-A]. Five of the thirty blanks for each steel were
used to determine the range of hold down loads where cups could be formed without
wrinkling their side walls or fracturing their bottoms. The acceptable range of hold down
loads ( FN ) for the STZC series steels and the GA steel was 2 to 11kN.The corresponding
range for the GI steel was 3 to 11kN. The remaining 25 blanks were used to measure the
punch load as a function of punch displacement for 5 different hold down loads and
subsequently to determine the slope,

dP
. The blanks were formed into cups employing
dFN

a punch speed of 20mm/sec after a mill oil was applied to the blanks using a soft brush.
Once the slope,

dP
was determined for a given punch displacement, the coefficient of
dFN

surface friction was calculated using equation (3.3) because the values for Rm is constant
(17.03mm) and and X are well established functions of punch displacement [24].

30

Figure 3.3: Schematic diagrams of cup test (A) and important parameters in equation (3.2) (B)

31

3cm

Figure 3.4: Preparations for cup test:(A-a) male die,(A-b) female die with 225mm diameter,
respectively, (A-c) punch with 33mm diameter, (B-a) 74mm blank, and (B-b) fully formed cup

32

3.3.1.2 Hemispherical Punch Test


In a stretching operation, deformation takes place in three distinct regions. These
are metal stretching in contact with the punch (region 1), metal stretching in the
unsupported region (region 2), and metal flowing through the blank holding areas
(drawing region 3 ) [Fig.3.5]. In contrast to the stretching mode of deformation in region
1 and 2, metal flow takes place through circumferential contraction in region 3. Thus,
region 1 is an optimal area to consider the surface friction under the condition of pure
stretching.
Amit K. Ghosh developed a model for friction measurement that is free from
material property inputs under the condition of stretching of sheets over punch surfaces
and a method to determine an average value of the coefficient of punch-to-sheet friction
in stretching over a 50.8mm hemispherical punch [25]. The model uses three assumptions
to simplify the complexity of the plastic deformation occurring in contact with the punch.
The first assumption is that interfacial pressure between the punch and sheet is uniform at
any stage of deformation. The assumption is very important because the pressure is
directly related to the surface friction. If the variation of the pressure during deformation
were large, the acquisition of a constant friction coefficient would be impossible.
According Wang [26], the pressure distribution shows a relatively small variation over
the punch contact area. Thus, the assumption is very reasonable. The second assumption
is that the frictional force is proportional to the applied load, i.e. Amontons law applies.
This means that frictional stress is proportional to the interfacial pressure, = p . The
third assumption is that the inflection point on the load-displacement plot corresponds to
33

the attainment of the maximum pressure. This means that the average pressure ( p )
exerted by the punch on the sheet surface reaches a maximum. Using these assumptions
and geometrical expressions for key parameters [Fig.3.6B] the following expression can
be derived to determine surface friction.

2 Xr0

X ( 2 / r0 ) sin 1 (r0 / ) ( 2 r0 2 ) (2r0 / ( 2 r0 2 ))

(3.4)

In equation 3.4, is the surface friction coefficient, rc is the radius of the punch-sheet
interface boundary, r0 is the value of rc at the inflection point in a plot of loaddisplacement, is the punch radius, X is [ (dP / drc ) / P ]r , and P is the punch load.
0

Figure 3.5: Generalized stretching forming operation [25]

34

The hemispherical punch test was used to obtain the coefficient of surface friction
under the condition of stretching based upon the work of Ghosh [25]. The hemispherical
punch employed had a standard diameter of 101.6mm, i.e. =50.8mm, and the blanks
were secured by a 132.6mm diameter draw bead [Fig. 3.6A]. The clearance between the
blank holder die set opening and the punch was 2.03mm. A hold down load of 170kN
was applied to completely eliminate metal flow into the die cavity. A constant punch
displacement rate of 25.4mm/min was maintained for all of the hemispherical punch
testing. The five different types of steels were sheared to form square shaped
(178mm 178mm) blanks.
The independent variables in Figure 3.6(B) are X and r0 . Two procedures were
required in order to determine X . Initially, punch load ( P ) versus punch displacement
( x ) plots were obtained from the hemispherical punch tests of the five different types of
steels. Punch displacements continued until the blanks fractured. The punch load data
obtained was used to convert to plots of punch load ( P ) versus radius of the punch-sheet
interface boundary ( rc ). According to Ghosh [25] [Fig. 3.7], the radius of the boundary of
the punch contact ( rc ) can be reliably determined as a function of punch displacement
( x ) regardless of the material or the test conditions. The punch load ( P ) versus the
radius of the boundary of the punch contact ( rc ) data were fitted with a fourth order
polynomial with an index of fit of 0.9999 using ORIGIN 8.0 software in order to get a
fourth order equation ( P = f (rc ) ) from which the derivative value, dP / drc could be
obtained. The next procedure was to determine an inflection point mathematically.
35

The punch load ( P ) versus punch displacement ( x ) plots were fitted with a fourth order
polynomial and fourth order equations were obtained from the fitting in a manner similar
to the first procedure. The specific punch loads and displacements at the inflection point
were determined after taking the second order derivative of the fourth order equations
( P = f ( x) ). The inflection point was found by solving for the value of X at which the
second derivative of plot was zero, i.e. d 2 P / dx 2 = 0 .
Once the punch load and displacement at the inflection point were determined, the
radius of the boundary of the punch contact at the inflection point ( r0 ) was obtained using
the first procedure. As a result, the coefficient of surface friction was calculated using
equation (3.4).

36

3cm

Figure 3.6: Punch with 101.6mm diameter and die set with 225mm diameter and 132.6mm
diameter draw bead for hemispherical punch test (A) and important parameters in equation (3.4)
(B)

37

Figure 3.7: Radius of the boundary of punch contact ( rc ) is plotted as a function of punch
displacement, for three test materials under different test conditions [25]

3.3.2 Measurement of Coating Adhesion Propensities


Coating adhesion was assessed by determining the fracture characteristics of a
coating. Powdering tests were employed to measure the amount of coating fracture that
occurred due to a forming operation.
3.3.2.1 Powdering Tests
The powdering tests were conducted after the sheet steels were subjected to
drawing and stretching strain states. Cup testing was employed to measure the amount of
coating exfoliation under drawing conditions and hemispherical punch testing was
employed to measure the same quantity under stretching conditions.
38

The test sequence began by cleaning each blank (77mm diameter) with methyl
alcohol followed by weighing each blank. A mill oil was then applied using a soft brush.
The blank was weighed again after application of the oil to monitor and control the
amount of oil applied. Approximately 0.1g oil equivalent to a 12m oil film thickness
was applied to each blank. The oiled blanks were then formed into cups employing a
punch speed of 20mm/sec and hold down loads in the range from 2kN to 11kN. Each
fully formed cup was rinsed in solvent and dried. After cleaning, Scotch 3710 tape was
applied to all of the cups inside and outside surfaces to pull off any loose coating.
Finally, the cup was weighed in gram once more. The difference in weigh between the
initial weighing of the blank and the final weighing of the formed cup established the
extent of powdering of the coated steel caused by the drawing operation.
A similar experimental procedure was applied after hemispherical punch testing
to measure the amount of exfoliation of a coating that occurred under a stretching strain
state. Blanks 178mm long with widths that varied from 125mm to 140mm were designed
to define the plane strain stretching condition. Before the blanks were subjected to
deformation they were cleaned with methyl alcohol and weighed. After the blanks were
deformed they were again cleaned with methyl alcohol and then Scotch 3710 tape was
applied to both their convex and concave surfaces to pull off any loose coating. The
deformed blanks were weighed again and the difference in weight between the initial
weighing and the final weighing was used to measure the extent of powdering of the
coating caused by the stretching operation.

39

3.3.3 Measurement of Formability


In many automotive stampings the failure of steel sheets occurs under a condition
of plane strain (minor strain 0). A.K. Ghosh suggested a test procedure in which dome
heights are measured as a function of minor strain and the minor strain can be varied by
modifying the blank width [27]. The punch height where a blank fails under a plane strain
condition can be used as an effective predictor of formability. This test procedure is now
commonly known as the limiting dome height or LDH test.
3.3.3.1 Limiting Dome Height (LDH) Test
The first procedure associated with a LDH test was to establish the blank width
that would generate a plane strain deformation condition. The plane strain condition was
determined by conducting a series of hemispherical punch tests with blanks that were
178mm long and had widths that varied from 115 to 148mm in intervals of 3mm [Fig.
3.8]. The blank width at which the specimen fractured after experiencing the minimum
punch displacement corresponded to the case of plane strain deformation. After the plane
strain deformation blank width was established, five replicated LDH tests were run with
that blank width to determine average punch heights.
3.4 Surface Analysis
Scanning electron microscope was employed to characterize the texture, i.e.
roughness and patterns found on the outer surface of each of the five steels which
experienced the Electrical Discharge Texturing (EDT) process. Electron Spectroscopy for
Chemical Analysis, also known as X-ray Photoelectron Spectroscopy (XPS) was
employed to analyze the concentration and the distributions of the main elements and
40

Figure 3.8: Relationship between major engineering strain vs. minor engineering strain and
limiting dome height vs. blank geometry

the phases in the extreme outer surface of each of the zinc coatings. There were two
objectives to the analysis of the extreme outer surface. First, the frictional behavior of the
sheet is tied to its extreme outer surface at contact, i.e. tangent points. Knowledge of the
phases and composition at the extreme outer surface could be useful for friction control.
Second, it is known that the aluminum in the coating layer controls the formation of FeZn phases. However, the diffusion of aluminum to outer surface where it forms alumina
could affect weldability negatively [28]. Thus, the outer surface analysis with ESCA
provides some information about the weldability of the five coated steels.

41

3.4.1 Surface Analysis by SEM


A scanning electron microscope equipped with a field-emission gun (HITACHI
S4500) was employed to take images of the surface of the coating layers. The specimens
were sheared into pieces 10mm by 10mm and cleaned with acetone. Plan view images of
each zinc coating were taken at 250 to provide a global view of the surface features
and at 500 to resolve details of the surface structure. Detailed images of potential
contact points and particles were taken at 3000 .
3.4.2 Surface Analysis by ESCA
A PHI 5600 ESCA instrument with an aluminum x-ray source and an inert gas
sputtering source was employed to obtain depth profiles of the extreme outer surface of
each of the five steels. The specimens were sheared into pieces of 10mm by 10mm and
cleaned with acetone. Depth profiles including the elements zinc, oxygen, aluminum, iron
and carbon were obtained for a maximum sputtering depth of 40nm.
3.5 Microstructural Characterization by SEM and EDS
Scanning electron microscope (SEM) and energy dispersive x-ray spectroscopy
(EDS) analyses of cross sectional samples were used to identify the Fe-Zn phases present
in each of the steel coatings and evaluate the disposition of their inhibition layers.
Samples 10mm 10mm were sheared from each of the five steels and placed in standard
31.75mm diameter mounts. Samples of each steel were positioned to form a stack with
sheared edges parallel to one another. The stacks were assembled so that the sheared
edges were aligned and the samples in each stack were separated from each other by
placing a small piece of double sticky tape between each sample away from the sheared
42

edges. Epoxy resin was used as the mounting media. Grinding through 400, 600, and
1200 grit paper was carried out by hand making sure that back and forth motion over
each paper was performed with the interface between the mounting material and the zinc
coatings oriented parallel to the direction of motion. This procedure minimized smearing
of the coating. Polishing proceeded through 3m and 1m diamond slurries finishing
with a 0.05m alumina slurry. The polishing machine employed was adjusted to apply a
heavy, even pressure to each mount. After polishing the specimens were etched in a
solution consisting of 5% nitric acid and 95% methanol, i.e. Nital.
Cross sectional SEM images of the coating layer found on each of the five steels
were taken at 2000 using the back scattering electron (BSE) mode. Energy dispersive xray spectroscopy (EDS) analyses were conducted at several location throughout the
thickness of the zinc coatings in order to investigate the compositions of the Fe-Zn phases.
Elemental maps of the zinc coatings for the elements zinc, iron and aluminum were
obtained to investigate the distribution of those elements through the thickness of the
coatings.

43

4. Results

4.1 Materials Characterization


Five types of zinc coated materials were characterized with respect to surface and
microstructure. The surface is one of the most important factors affecting friction and the
microstructure of a coating is directly related to its adhesion propensities.
4.1.1 Surface
Surface morphologies were investigated by scanning electron microscopy at
various magnifications. Areal fraction of contact between a coated steel sheet and a die,
i.e. the sum of the contact with surface asperities when the normal pressure is roughly
zero was calculated using a systemic point count method with captured images taken with
a SEM. The depth profiles of alloying elements found in the outer surface of zinc coating
were investigated by ESCA.
4.1.1.1 Surface Morphology
Three different types of morphology were imprinted onto the STZC1, STZC2 and
STZC3 steels via a skin pass [Fig.4.1]. The first type of pattern imprinted on the STZC1
steel has a well rounded droplet-like shape [Fig. 4.1(a)]. The second type of pattern
imprinted on the STZC2 steel has a truncated droplet -like shape [Fig. 4.1(b)]. The last
type of pattern imprinted on the STZC3 steel has a elongated droplet- like shape [Fig.
4.1(c)]. The average distance between the pattern shapes decreases from the STZC1 steel
to the STZC3 steel. The GI and GA steels which underwent a temper rolling process
exhibited different surface morphologies. The temper rolling process flattened and
44

merged their surface patterns unlike the skin pass process and so formed a conglomerated
mesa-like surface morphology, i.e. hills with flat tops. The pattern on the GI steel has a
coarse mesa-like shape [Fig. 4.1(d)]. The pattern on the GA steel has a fine mesa-like
shape and the distance between pattern shapes is smaller than that found on the GI steel
[Fig. 4.1(e)].
A unit of pattern creates an asperity point, i.e. initial contact point with a die, and
its shape and distribution affect surface friction directly [29]. Figure 4.2 shows different
patterns formed via either the skin pass or the temper rolling process. An asperity point
on the STZC1 steel has a rounded droplet-like shape 20 m in diameter and it is
distributed very sparsely [Fig. 4.2(a)]. An asperity point on the STZC2 steel has a
truncated droplet -like shape with a 20m major axis and 10m minor axis [Fig. 4.2(b)].
The shape and distribution of its asperity points falls between those found on the STZC1
and STZC3 steels. An asperity point on the STZC3 steel has an elongated droplet- like
shape approximately 20m in size across its diagonal. Its asperity points are distributed
very densely [Fig. 4.2(c)]. An asperity point on the GI steel has a mesa-like shape with a
diagonal dimension of 50m and these form flat and large conglomerates that are
distributed sparsely [Fig. 4.2(d)]. An asperity point on the GA steel has a fine mesa-like
shape with a diameter of 6 to 7 m. These fine asperity points form flat and small
conglomerates densely distributed over the surface [Fig. 4.2(e)]. The matrix surrounding
these asperity points on the GA steel is comprised of rough valleys whereas those of the
STZC steels and the GI steel are fairly planar. These imply that the Fe-Zn intermetallic
crystals that comprise the GA steel affect the size of the asperity points, the size of
45

conglomerates, the spacing of conglomerates and the matrix structure surrounding the
asperity points.

46

Figure 4.1: SEM plan view images (200 ) of the STZC1 steel (a) and STZC2 steel (b)
47

Figure 4.1: SEM plan view images (200 ) of the STZC3 steel (c) and GI steel (d)
48

Figure 4.1: SEM plan view images (200 ) of the GA steel (e)

49

Figure 4.2: SEM images of pattern (asperity) on the (a) STZC1, (b) STZC2, (c) STZC3, (d) GI
and (e) GA steels. Yellow arrow indicates surface asperity
50

4.1.1.2 Real Area of Contact ( Areal Fraction of Asperities )


The real area of contact of the five coated steels at an initial normal pressure near
zero was calculated with the aid of equation 3.1. The areal fractions of asperities are
given in Table.4.1 and the real area of contact as a function of the steel type is shown in
Figure 4.3. According to these results, the real area of contact increases from the STZC1
steel to the STZC3 steel. This result is related to the curvature of the asperities and the
repeat of the asperities pattern rather than the size of the asperities. The shape of the
asperities on the STZC1 steel is hemispherical with a high curvature and a low contact
area. The shape of the asperities on the STZC3 steel is roughly that of a flat rectangle
which has a low curvature and a high contact area. The shape of the asperities found on
the STZC1 steel is more regular than those found on the STZC3 steel. As a result, the
asperities with the high curvature and the high regularity yield a lower real area of
contact. The values for the real contact area of the GI and GA steels are very high and
similar to the STZC2 and STZC3 steels. Asperity flattening by the temper rolling process
contributes to an increase in the real area of contact.
4.1.1.3 Depth Profile of Alloying Elements
The concentration profiles of the outer surfaces were acquired using a sputtering
process in an ESCA system. Figure 4.4 shows the distribution of elements as a function
of depth below the surface in the range of 0 to 40nm. The STZC steels and the GI steel
profiles contain signals from the Zn2p3/2, Al2p, O1s and C1s ESCA peaks. The lack of
an Fe2p signal in these profiles proves that the outer surfaces of these four coatings are
uniformly comprised of a pure zinc phase, i.e. an Fe-Zn intermetallic phase was not
51

present in the outer 40nm of these four coating. The GA steel contains a signal from the
Fe2p peak profile in addition to the elements found on the STZC steels and the GI steel
profiles. The presence of iron proves that the outer surface of the GA contains an Fe-Zn
intermetallic phase.

Table 4.1 Areal fraction of asperities at initial normal pressure( P ) = 0


PP

Steel Type
STZC1
STZC2
STZC3
GI
GA

95%CI
2.7
4.1
4.5
3.8
4.1

21.7
38.5
39.5
34
39

50

Real area of contact at P=0 (%)

Real area of contact at P=0


45
40
35
30
25
20
15

STZC1

STZC2

STZC3

GI

GA

Steel type
Figure 4.3: Real area of contact of five coated steels at initial normal pressure=0
52

A significant concentration of oxygen was found for all five types of coatings at a
depth from 0 to 15nm below their surfaces. A mixture of zinc oxide (ZnO) and aluminum
oxide (Al2O3) must be present in the outer surface [30]. Under an assumption that the
total oxygen in the outer surface is consumed by the formation of the zinc oxide and the
aluminum oxide and the states of zinc in outer surface are a metal (Zn0) and an oxide
(Zn2+), it is possible to derive ratios of zinc oxide to zinc metal for the five types of
coatings. The results are given in Table 4.2. The higher galvannealing temperature and
the longer galvannealing time produced a higher percentage of oxidized zinc in outer
surface. The STZC3 steel and the GA steel have high percentages of oxidized zinc. The
conclusion is that the galvannealing process is one of the main factors that determines
what oxidation processes occurs on the outer surface.
The Al2p profiles reveal that a considerable amount of aluminum segregation to
the surface occurred. The aluminum added to the zinc bath for the formation of the
inhibition layer diffused to the outer surface and formed an aluminum oxide (Al2O3) layer.
The free energy of formation of Al2O3 is much lower than that for ZnO and provides a
driving force for the formation of Al2O3 on the surface of each coating. The profile of the
GI steel where the content of aluminum is two times higher than that of others is similar
to that of the others. The GI steel experienced no elevated temperature thermal exposure
after its coating layer solidified. The thermal exposure associated with the galvannealing
process affected the distribution of aluminum oxide. The GA steel which experienced
complete galvannealing processing developed the thickest Al2O3 layer. The STZC
samples had surface aluminum oxide levels that were lower than the GI and GA steels as
53

shown in Table 4.2. The high carbon content in the range from 0 to 5nm in the C1s
profiles resulted from either surface contamination due to residual oil or other
hydrocarbons.

54

Zn2p3/2

Zn2p3/2

O1s
O1s

Al2p

C1s

Al2p

C1s

Zn2p3/2

O1s

Zn2p3/2

O1s

C1s
Al2p

C1s

Al2p

e
Zn2p3/2

O1s
Fe2p
Al2p
C1s

Figure 4.4: XPS (ESCA) Depth Profile of (a) STZC1 (b) STZC2 (c) STZC3 (d) GI and (e) GA
steel coatings

55

Table 4.2 Atomic concentration, elemental ratio and zinc state ratio observed by XPS at 5nm
depth in the surfaces of five different coated steel sheets

Steel
Type

STZC1

STZC2

STZC3

GI

GA

Element

Atomic
Concentration
(at.%)

Zn

52

Al

11

28

Zn

52

Al

31

Zn

42

Al

45

Zn

47

Al

14

33

Zn

34

Al

13

47

Zn:Al:O

ZnO:Zn

Galvannealing
Temp.(C)

9:2:5

2:8

452

13:2:8

4:6

465

10:2:12

9:1

516

7:2:5

3:7

5:2:7

8:2

505

20

56

Galvannealing
Time(s)

1~5

4.1.2 Microstructure
The microstructural characterization of five coated steel sheets was carried out
using scanning electron microscopy (SEM) analysis combined with energy dispersive xray spectroscopy (EDS). Cross sectional specimens were used to evaluate morphological
changes across the coating thickness. Quantitative analysis of the chemical composition
of the coatings was performed at select points. Elemental maps were obtained to
qualitatively determine the distribution of phases in each coating.
4.1.2.1 Cross sectional Morphology
Cross sectional SEM images of the coating layer found on each of the five steels
are shown in Figure 4.5. The images were taken using the back scattered electron (BSE)
mode. In the BSE mode a region with a higher average atomic number will appear
brighter than an area with a lower average atomic number. Because the five coatings are
predominantly zinc with an atomic number of 30 and the substrates are predominantly
iron with an atomic number of 26, the zinc coatings are brighter than their substrate steels.
The GA steels coating had an average thickness of nominally 8 to 10m equivalent to
about 70g/m2. The GI steels coating had an average thickness of nominally of 17 to
19m equivalent to about 140g/m2. The average thickness of the zinc coatings on
STZC1,STZC2, and STZC3 steels fell between those of the GA and GI steels, i.e.
corresponding to a coating weight of 120g/m2. The variability in coating thickness
observed in the cross sectional images is consistent with the patterned surface structure of
the coatings that were found in the plan view images as shown in Figure 4.1 and Figure
4.2.
57

Fe-Zn intermetallic blocky crystals were identified in the STZC samples. The
yellow arrows indicate the crystals in Figure 4.5(a~c). The STZC3 steel [Fig.4.5(c)] with
its high annealing temperature and longer hold time has thick and long crystals whereas
the STZC1 steel [Fig.4.5(a)] with its low annealing temperature and short hold time has
slender and short crystals. The longer the hold time and the higher the annealing
temperature during galvannealing process resulted in more diffusional transformation
within the coating layer. Individual Fe-Zn crystals are not evident in the GI [Fig.4.5(d)]
and the GA [Fig.4.5(e)] samples because the GI steel did not experience the
galvannealing process and the GA steel was fully transformed to a single Fe-Zn
intermetallic phase.
Vertical and horizontal cracks in the coating and at the steel interface were found
in the GA steel [Fig.4.5(e), yellow circles] whereas they are not identified in the STZC or
GI samples. It means that the Fe-Zn intermetallic phase formed via the full galvannealing
process had more of a brittle character than the other four coatings. In Figure 4.5 coating
debris above the outer surface of several of the coatings appears to have formed during
sample preparation.

58

6m
4m

10m
0m

11m

Figure 4.5: Cross sectional images of five types of coating layer captured by back scattered
electron mode in SEM (2000 ) : (a) STZC1 (b) STZC2 (c) STZC3 (d) GI and (e) GA steels.
(Arrows from (a) to (e) indicate an Fe-Zn intermetallic phase formed through the thickness and its
average vertical height. Circles and ellipse in (e) indicate vertical cracks and horizontal crack,
respectively.)
59

4.1.2.2 Phase Identification of Coating Layers


Elemental mapping of the main alloying elements in the five zinc coatings was
conducted by energy dispersive x-ray spectroscopy (EDS). These maps are shown in
Figure 4.6(a) through Figure 4.10(a). It is obvious that the predominant element in the
steel substrate is iron and that in the coatings is zinc from the Fe-K maps and Zn-K ,
respectively. The STZC samples contain regions with a slightly higher iron intensity
compared to the majority of their coating. These regions appear to be where Fe-Zn
intermetallic phases formed. The width and height of the regions are consistent with those
of phases in the zinc coatings visible in Figure 4.5 ( indicated by yellow arrows ). The
GA steel exhibits the highest Fe intensity and indicates that the galvannealing process
resulted in sufficient iron diffusion to form a coating completely comprised of an Fe-Zn
intermetallic compound.
The aluminum signals in the STZC samples showed increased intensity both in
regions where Fe-Zn intermetallic phases were formed and at the outer surface. Detecting
aluminum at the outer surface through EDS mapping is consistent with the result
obtained using XPS [Fig.4.4]. An understanding of the ternary diagram of the Fe-Zn-Al
system aids in understanding the increased aluminum signal near the interface between
the zinc coatings and their substrates. The ternary diagram is shown in Figure 4.11. The
phase diagram contains a Fe-Zn intermetallic phase Fe2Al5 which serves as the inhibition
layer (reddish area). The high aluminum content of the Fe2Al5 inhibition layer produces
the enhanced intensity of the aluminum signal found at the coating/ substrate interface of
several of the steel samples. The signal from the Fe2Al5 inhibition layer is strongest for
60

the GI sample and essentially nonexistent for the GA sample. The inhibition layer
prevents the dissolution of iron into the molten zinc and in turn suppresses the formation
of Fe-Zn intermetallic phases.
Analyses of selected points in the five zinc coatings were conducted using energy
dispersive x-ray spectroscopy (EDS). Typical spectra obtained from the five types of
steels are displayed in Figure 4.6 (b), (c) through Figure 4.10 (b), (c). The phases
expected in the coatings are the Fe-Zn intermetallic phases delta () and zeta () and the
zinc solid solution phase eta (). The phase is nominally FeZn10 and its solubility range
for Zn is 85.6 to 91.8 at. % according to the Fe-Zn binary phase diagram [31]. The
phase is nominally FeZn13 and its solid solubility range for Zn is 92.8 to 95.0 at.%. The
phase has a solid solubility range for Zn of 98.5 to 100 at. %. According to all of the EDS
spectra [Fig. 4.6(b) ~ 4.10(b)], the main elements in all five coatings were zinc and iron.
Aluminum, silicon, phosphorous, carbon, and oxygen were detected as trace elements.
Aluminum is added to the zinc bath to form the inhibition layer. Silicon and phosphorous
are elements whose signal may have originated from the interstitial free (IF) steel and
rephosphorized steel substrates. The presence of carbon seems to be attributed to surface
contamination such as hydrocarbons from the chamber, vacuum pump, and the surface of
the sample [33]. The oxygen signal might be a result of surface oxidation. Quantification
of the EDS spectra revealed that the zinc content ranged from 91 to 94 at. % and the iron
content ranged from 6 to 9 at. % in the STZC samples. That result indicates the Fe-Zn
intermetallic phase found in the STZC samples is almost certainly phase. Their higher
than expected measured iron content seems to be attributed to extra iron signal being
61

detected by the EDS system because the phases are very close to their steel substrates.
The average zinc content away from the phase was 98 at. % . The zinc coatings for the
STZC samples were comprised of a matrix phase containing particles of the Fe-Zn
intermetallic phase. The zinc content ranged from 87.1 to 88.2 at. % and the iron
content ranged from 11.8 to 12.9 at. % in the GA sample. The GA sample with an
average concentration of 87.8 at. % zinc and 12.2 at. % iron was clearly comprised of
essentially pure phase. The zinc content ranged from 98.4 to 98.7 at. % and the iron
content ranged from 1.3 to 1.6 at. % in the GI sample. The GI sample with an average
concentration of 98.6 at. % zinc and 1.4 at. % iron was comprised of essentially pure
phase.

62

STZC1, (a)

(b)

(b)

(c)

Figure 4.6: EDS element mapping through the thickness of the STZC 1 steel (a) and energy
spectrum (b) and chemical composition (at.%) (c) from an Fe-Zn intermetallic phase (yellow
arrow) in the steel coating
63

STZC2, (a)

(b)

(b)

(c)

Figure 4.7: EDS element mapping through the thickness of the STZC 2 steel (a) and energy
spectrum (b) and chemical composition (at.%) (c) from an Fe-Zn intermetallic phase (yellow
arrow) in the steel coating
64

STZC3, (a)

(b)

(b)

(c)

Figure 4.8: EDS element mapping through the thickness of the STZC 3 steel (a) and energy
spectrum (b) and chemical composition (at.%) (c) from an Fe-Zn intermetallic phase (yellow
arrow) in the steel coating
65

GI, (a)

(b)

(b)

(c)

Figure 4.9: EDS element mapping through the thickness of the GI steel (a) and energy
spectrum (b) and chemical composition (at.%) (c) from a pure zinc coating (yellow arrow) on the
steel
66

GA, (a)

(b)

(b)

(c)

Figure 4.10: EDS element mapping through the thickness of the GA steel (a) and energy
spectrum (b) and chemical composition (at.%) (c) from a fully galvannealed zinc coating (yellow
arrow) on the steel
67

Fe2Al5

Figure 4.11: Fe-Zn-Al ternary phase diagram (isothermal section at 500C) [32]

68

4.2 Surface Friction


The surface friction characteristics of the five zinc coatings were evaluated by
measuring their coefficients of friction under two strain states. Cup tests simulating
drawing forming conditions and hemispherical punch tests simulating stretching forming
conditions were conducted to obtain the coefficient of surface friction of each of the five
zinc coatings.
4.2.1 The Coefficient of Surface Friction under Drawing Conditions
A series of cup tests described in Section 3.3.1.1 was conducted to obtain the
coefficient of surface friction for the five zinc coated steels. Figure 4.12 shows punch
load profiles (red profiles with black data points) as a function of hold down load and
punch displacement for the five zinc coated steels. The data points with a blue color
indicate a linear relationship between punch load ( P ) and hold down load ( FN ) for all
punch displacements. That result implies that the relationship between the zinc coating
and die surfaces at every punch displacement obeys Amontons law, i.e. the frictional
force between the two surfaces is proportional to the load pushing them together. The
data points with a green color indicate a parabolic relationship, i.e. concave downward,
between punch load ( P ) and punch displacement ( x ). The punch load increases until the
punch displacement is 18mm. That result show that the load required to draw the sheet
into the die increases due to several reasons such as strain hardening caused by increased
plastic deformation and sliding friction. That regime corresponds to a pure drawing
condition. After that region, the punch load starts to drop in the range of 20mm to 30mm.
It means that the area withstanding the hold down load starts to decrease due to sliding of
69

material and the load loss due to a decrease of the area is much higher than the load
gained by strain hardening and sliding friction.
Maximum punch loads were applied at a punch displacement of 18mm in all five
zinc coated steels. This implies that the position at which the maximum loads are applied
is not a material factor, but a geometrical factor. The punch load was 32kN, 29kN and
37kN in the STZC1, STZC2 and STZC3 steels, respectively. The reason why the STZC3
steel exhibited the highest value for the punch load is that its substrate had the highest
tensile strength of the five steel types. The maximum punch load mainly depends on the
mechanical properties of the substrate because the thickness of the zinc coating is only 2
to 3 % of total thickness of a sample. The GI steel composed of a pure zinc coating
showed the lowest maximum pressure, 28kN whereas the GA steel and the STZC steels
with Fe-Zn intermetallic phases exhibited maximum punch loads in the range of 29kN to
37kN.
The slopes of forming load ( P ) vs. hold down load ( FN ) were obtained from
Figure 4.12 and used to calculate the coefficient of surface friction by applying Equation
3.2. The results are plotted in Figure 4.13. The coefficient of surface friction shifts as the
punch displacement increases in all five zinc coated steels. It implies that the forming of
the zinc coated steel sheet agrees with the dynamic friction theory described in Section
2.2.1 and 2.2.2. The traces of the coefficient of surface friction in Figure 4.13 are similar
to the prediction of the Daltons friction model (Fig.2.5). The asperities on the surface
start to deform plastically making the real contact area increase between punch
displacements of 10 to 20mm. This results in an increase in the coefficient of friction.
70

The constant value of the coefficient of friction then decrease for punch displacement
between 20 and 30mm is interpreted by Daltons model B and model C, respectively.
According to model B, the coefficient of friction is stable before it starts to increase or
decrease. According to model C, the coefficient of friction starts to drop as the forming
process continues. This can be explained by the trapped lubricant effect illustrated in
Figure 2.6. The effect means that the presence of trapped lubricant results in a
superimposed hydrostatic pressure acting in the valleys of the asperities which can reduce
friction in metal working processes.
The STZC1 samples and the STZC2 samples show a similar trace to that of the GI
samples whereas the STZC3 samples show a similar trace of that with the GA samples at
the pure drawing region. The coefficient of friction is significantly affected by the zinc
coating in that region. The STZC1 coating and the STZC2 coating are composed of
mostly the phase. As such they are similar to the zinc coating found on the GI samples.
The STZC3 coating contains a significant fraction of the Fe-Zn intermetallic phase
along with the phase. That coating microstructure is roughly similar to that found in the
GA samples. The variation in the coefficient of friction in the STZC1, STZC2 and the GI
samples is larger than that in the STZC3 and the GA samples. It implies that the shear
strength of the zinc coatings from the former are lower than that from the latter so that the
coatings in the former experience more severe plastic deformation, i.e.
( y ( Fe Zn ) y ( Zn ) ). The value of the coefficient of friction for the STZC1, STZC2 and
GI samples were between 0.09 and 0.11 and for STZC3 and the GA samples between
0.13 and 0.15. The latter group containing the Fe-Zn intermetallic and phases has a
71

higher range of coefficients of friction than the former group. The phenomenon can be
explained by Bowden and Tabors theory because the selected coefficients of friction
occur after a punch displacement that corresponds to considerable plastic deformation of
asperities and prior to a drop of friction by a lubricant effect, i.e. pseudo- dry condition .
According to their theory described in Section 2.1.2, the friction coefficient can be
expressed as =

FC y Ar y
when plastic deformation of surface asperities is
=
=
FN Ar H H

dominant and dry sliding contact occurs. As a result, the STZC3 and the GA samples
with a high shear strength ( y ( Fe Zn ) ) zinc coating develop higher surface friction under
drawing conditions than the STZC1, the STZC2 and the GI samples with their lower
shear strength ( y ( Zn ) ) zinc coating.
4.2.2 The Coefficient of Surface Friction under Stretching Conditions
The coefficient of surface friction under stretching conditions was calculated
following the procedure outlined in Section 3.3.1.2. Figure 4.14 shows the trace of the
punch load as a function of the punch displacement and the radius of the contact
boundary for the five zinc coated steel sheets. Because the relationship between the
punch displacement and the radius of the contact boundary ( rc ) does not depend on the
type of material, that relationship is the same for the five zinc coated steel sheets as
shown by the red dots in Fig.4.14. Using the data for the punch displacement vs. the
radius of the contact boundary and the data for the punch displacement vs. the punch load
obtained with the dual actuator hydraulic press as shown by the green dots in Fig.4.14,
the plots of the punch load vs. the radius of the contact boundary were obtained as shown
72

by the blue dots in Fig.4.14. Fourth order polynomial equations were fitted to the plots of
the punch load versus the radius of the contact boundary using the software ORIGIN. The
equations are given in Table 4.3. After the inflection points and values for X were defined
from punch load vs. punch displacement plots and the fourth order polynomial equations,
respectively, the coefficients of friction under stretching condition were determined. The
coefficient of friction are listed in Table 4.4
The coefficients of friction for the STZC1, the STZC2 and the STZC3 samples
were 0.21, 0.23 and 0.24, respectively. The coefficients of the GI and the GA samples
were 0.24 and 0.19, respectively. The values are higher than those obtained under the
drawing conditions. That result can be explained by the fact that the stretching forming
process produces a significant amount of new surface area which has intrinsically higher
frictional characteristics and also the much higher plastic deformation under stretching
conditions can lead to break out of the surface oil film [34]. Under stretching conditions,
it seems that the degree of plastic deformation of the surface asperities is very important.
The GA sample showed the lowest coefficient of friction under the stretching conditions.
It implies that the Fe-Zn intermetallic phase in the coating with its high shear strength
resists plastic deformation of asperities during the forming which suppress the breakdown
of the oil film and maintains low surface friction. As a result, in spite of the stretching
accompanying severe plastic deformation the GA sample maintains its oil pockets more
effectively compared to the other zinc coatings leading to its relatively low coefficient of
friction.

73

(b)

34

32

32

30

30

Punch Load (kN)

Punch Load (kN)

(a)

28
26
24
22
20

28
26
24
22

12
5

20

10

Pun 10 15
ch
Dis 20
pla
cem 25 30
ent
(mm
)

8
6
4
2

ld
Ho

35

wn
Do

ad
Lo

N)
(k

12
10

10

Pun
ch 15 20
Dis
pla
cem 25 30
net
(mm
)

(c)

8
6
4
35

ld
Ho

wn
Do

)
kN
d(
a
Lo

(d)

38
30

36

28

32

Punch Load (kN)

Punch Load (kN)

34

30
28
26
24
22
20

26
24
22

12

18

)
kN
8
d(
a
6
o
nL
4
ow
D
2
ld
Ho
10

10

Pun
ch 15 20
Dis
pla
cem 25 30
ent
(mm
)

35

20
5

12

Pun
ch 15 20
Dis
pla
cem 25 30
ent
(mm
)

)
kN
d(
a
Lo
10

10

8
6
4
35

ld
Ho

wn
Do

(e)
32

Punch Load (kN)

30
28
26
24
22
20

12

Pun 10 15
ch
Dis 20
pla
cem 25 30
ent
(mm
)

)
kN
d(
a
Lo
10

8
6
4
35

ld
Ho

wn
Do

Figure 4.12: Punch load as a function of hold down load and punch displacement of the (a)
STZC1, (b) STZC2, (c) STZC3, (d) GI and (e) GA steels

74

0.24

0.24

STZC1

STZC2

0.22

0.20

0.20

0.18

0.18

Coefficient of Friction

Coefficient of Friction

0.22

0.16
0.14
0.12
0.10
0.08
0.06
0.04
0.02

0.16
0.14
0.12
0.10
0.08
0.06
0.04
0.02

0.00

0.00
8

10

12

14

16

18

20

22

24

26

28

10

12

Punch Displacement (mm)

0.24

16

18

20

22

24

26

28

26

28

GI

0.22

0.20

0.20

0.18

0.18

Coefficient of Friction

Coefficient of Friction

0.24

STZC 3

0.22

14

Punch Displacement (mm)

0.16
0.14
0.12
0.10
0.08
0.06
0.04
0.02

0.16
0.14
0.12
0.10
0.08
0.06
0.04
0.02

0.00

0.00
8

10

12

14

16

18

20

22

24

26

28

10

0.24

12

14

16

18

20

22

24

Punch Displacement (mm)

Punch Displacement (mm)

GA

0.22

Coefficient of Friction

0.20
0.18
0.16
0.14
0.12
0.10
0.08
0.06
0.04
0.02
0.00
8

10

12

14

16

18

20

22

24

26

28

Punch Displacement (mm)

Figure 4.13: Coefficient of surface friction calculated as a function of punch displacement for
cups drawn from the STZC1, STZC2, STZC3, GI and GA steels

75

60
55
50
45
40
35
30
25
20
15
10
5
0

(b)

Punch Load (kN)

Punch Load (kN)

(a)

45 ry
40
35 da
30 oun
25 b

10

60
55
50
45
40
35
30
25
20
15
10
5
0

Pun 15
b
20
ch 20 25
15 act
Dis
pla 30 35
10 ont
)
5
(mm cem
f c (mm
ent 40 45 0
.o
)
d
Ra

Pun 15 20
20 ct
ch
ta
15
Dis 25 30
10 con m)
pla
35
f
5
c
(m
o
40
(mm em
d.
ent
45 0
)
Ra

(d)

Punch Load (kN)

Punch Load (kN)

(c)
60
55
50
45
40
35
30
25
20
15
10
5
0

45 y
40 ar
35 d
30 oun
25 t b

45
40 y
35 ar
30 und
25 o

10

Pun10 15
20 c
ch 20 25
15 nta )
Dis
10 co
pla 30 35
m
5
(m
of
(mm cem
ent 40 45 0 ad.
)
R

60
55
50
45
40
35
30
25
20
15
10
5
0

45 y
40 ar
35 d
30 oun
25 t b

Pun10 15
20 c
ch 20 25
15 nta )
Dis
10 co
pla 30 35
m
5
(m
of
(mm cem
ent 40 45 0 ad.
)
R

Punch Load (kN)

(e)
60
55
50
45
40
35
30
25
20
15
10
5
0

10

45 y
40 ar
35 d
30 oun
25 t b

Pun 15
20 c
ch 20 25
15 nta )
Dis
10 co
pla 30 35
m
f
5
(m
o
(mm cem
ent 40 45 0 ad.
)
R

Figure 4.14: Punch load as a function of punch displacement and radius of contact boundary of
the (a) STZC1, (b) STZC2, (c) STZC3, (d) GI and (e) GA steels

76

Table 4.3: Polynomial fits for punch load, P , as a function of contact radius, rc in mm
Steel Type

Condition

Punch Load in Kilo-Newton

STZC1

Oil

-0.0888+0.1492rc -0.0114rc2 +0.0024rc3-0.0000338rc4

STZC2

Oil

-0.0686+0.1311rc -0.0111rc2 +0.0022rc3-0.0000305rc4

STZC3

Oil

0.16612-0.3501rc+0.0709rc2-0.0011rc3 +0.0000154rc4

GI

Oil

-0.0629+0.1305rc -0.0093rc2 +0.0021rc3-0.000029rc4

GA

Oil

-0.0562+0.0993rc -0.0037rc2 +0.0019rc3-0.000026rc4

Table 4.4: Calculated coefficient of friction, ,under stretching condition


Steel Type

Condition

Test Rate

STZC1

Oil

25.4mm/min

0.21

STZC2

Oil

25.4mm/min

0.23

STZC3

Oil

25.4mm/min

0.24

GI

Oil

25.4mm/min

0.24

GA

Oil

25.4mm/min

0.19

77

4.3 Coating Adhesion Properties


Powdering was measured by the test procedure described in Section 3.3.2.1. The
powdering tests were conducted separately after the cup tests and the hemispherical
punch tests in order to assess the effect of the forming conditions corresponding to
drawing and stretching.
4.3.1 Powdering under Drawing Conditions
The powdering rate in the five zinc coated steels was measured after the steels
experienced various drawing strain conditions. The results are given in Figure 4.15. The
value of the powder weight is constant regardless of hold down load for all the coated
steels. It implies that the powering rate is independent of the pressure normal to the
forming direction. If the coating layer was fractured by the wedging mechanism
described Section 2.3.1, the normal pressure may have aided the suppression of either the
propagation of pre-existing cracks by lateral tension or interface cracks by lateral
compression [Fig.2.7-8]. However, the normal pressure is too low to affect the powdering
rate. As a result, the normal pressure under drawing conditions showed little influence on
the powdering rate.
The values of the powder weight of the STZC1, STZC2 and STZC3 samples were
0.002 0.0015 g/m2, 0.002 0.0015 g/m2 and 0.0035 0.0015 g/m2, respectively. The
STZC3 sample with a high volume percent of phase had a higher powdering value than
the STZC1 and STZC2 samples that had relatively low volume percents of phase. It
suggests that the coarse phase crystals in STZC3 sample might affect adversely the
powdering rate because the coarse phase with relatively high hardness acts as an easy
78

crack nucleation site causing a decrease in the fracture stress of the coating [35]. As a
result, anti-powdering characteristics might be deteriorated by an increase in the ratio of
phase to phase in the coating layer among short duration galvannealed steel sheets
containing a discontinuous phase. However, the GI composed of 100% phase
exhibited a higher powdering rate value of 0.006 0.003 g/m2 compared to the STZC
samples containing phase. This result implies that the powdering tendency is affected
by not only microstructure, but also other factors such as interfacial shear strength.
The powdering weight of the GA showed the highest value of 0.1 0.025 g/m2 for
the five zinc coated steels as expected. The result agrees with that from Inagaki et al. [36]
and Nakamori et al. [37]. According to their results, an increase in the volume fraction of
phase with an Fe concentration from 8 to 15 wt.% is responsible for the deterioration of
the powdering resistance. It suggests that the increase of hardness with increasing iron
content prevents the lateral tension and compression stresses from being relaxed and as
such facilitates the propagation of cracks. As a result, the phase with its relatively high
iron content causes increased levels of powdering.
4.3.2 Powdering under Plane Strain and Stretching Conditions
After the Limiting Dome Height (LDH) tests, a series of powdering tests was
conducted with the five coated steel sheets. One of advantages of the LDH test is that the
biaxial strain condition can be varied by modifying the blank width [Fig.3.8] so that the
powdering rate under different biaxial strain conditions can be easily compared. The
results are given in Figure 4.16. Generally the powdering values were similar to the ones
found under drawing conditions. The STZC samples had lower powdering values than
79

the GI sample. The STZC3 sample had a higher powdering value than the STZC1 and
STZC2 samples. The GA sample had the highest powdering values for both stretching
and drawing strain conditions.
The biaxial strain condition was modified by varying the blank width from
124mm to 139mm. That range of blank width produced conditions of drawing, plane
strain and stretching. The STZC1 sample developed a plane strain condition with a blank
width of 130mm. Its powder weight had the highest value at the plane strain condition.
The powder weights under stretching conditions were higher than that at a drawing
condition. The STZC2 sample achieved the plane strain condition at a 124mm blank
width. The powder weight under the plane strain condition for the STZC2 sample was
higher than the STZC1 sample. The STZC3 achieved the plane strain condition at a
127mm blank width. The powdering value at the plane strain condition for the STZC3
sample was higher than the values for the STZC1 and the STZC2 samples. This could be
explained by Smiths theory that addresses a brittle second phase as a crack source as
discussed in section 4.3.1. The GI sample achieved the plane strain condition at a 130mm
blank width. The powdering values were higher than the ones from the STZC samples
regardless of the strain condition. It seems that the low interfacial shear strength of the GI
steels coating resulted in a higher powdering rate. The GI showed the highest powdering
value at the plane strain condition like the STZC samples. The GA sample achieved a
plane strain condition at a 139mm blank. Its powdering values were much higher than the
ones from the STZC samples and the GI sample regardless of the strain condition. The
presence of the phase is the most important factor affecting the powdering rate
80

regardless of the strain condition. Unlike the STZC samples and the GI sample, the GA
sample exhibited the highest powdering rate at the drawing condition rather than the
plane strain condition. Knowledge of the fracture modes for a phase based - zinc
coating and the phase based zinc coating is needed to understand the difference. in
powdering behavior between the STZC samples and the GA sample. According to Parisot
R. et al. [38], intergranular cracking occurs very early during the deformation of an
phase based coating, leading to a relaxation of stresses. That crack type does not reach
the substrate whereas cleavage cracking that occurs under highly tensile conditions such
as plane strain and stretching condition reaches the coating/substrate interface. It implies
that the dominant facture mode for powdering in an phase based coating would be a
cleavage mode rather than an intergranular mode when the coating is fractured by a
wedging mechanism. It makes sense that an phase based coating showed the higher
powering rate at either plane strain or stretching conditions rather than drawing condition
because the cleavage cracks occur only in plane strain and stretching conditions.
However, the phase based zinc coating found in the GA sample shows a different
fracture mode. The phase based zinc coating seems to be much more sensitive to
whether a strain condition keeps a component of compression or not. It makes sense that
the powdering rate in the GA sample was higher under a drawing condition than under a
plane strain condition because the lateral contraction inducing compression is much
larger for a drawing condition as opposed to a plane strain condition. As a result, the
powdering in the phase based coating was due to crack propagation in a direction
parallel to the lateral compression rather than the crack propagation in the normal
81

direction, i.e. a compression controlled fracture mode determined the extent of


powdering.

82

(b)
0.010

0.009

0.009

0.008

0.008

Powder Weight (g/m2)

Powder Weight (g/m2)

(a)
0.010

0.007
0.006
0.005
0.004
0.003
0.002
0.001

0.007
0.006
0.005
0.004
0.003
0.002
0.001

0.000

0.000
0

10

12

Hold Down Load (kN)

(c)

10

12

(d)

0.010

0.010

0.009

0.009

0.008

0.008

Powder Weight (g/m2)

Powder Weight (g/m2)

Hold Down Load (kN)

0.007
0.006
0.005
0.004
0.003
0.002
0.001

0.007
0.006
0.005
0.004
0.003
0.002
0.001

0.000

0.000
0

10

12

Hold Down Load (kN)

10

12

Hold Down Load (kN)

(e)

(f)

0.20

0.12

STZC1
STZC2
STZC3
GI
GA

Powdering Weight (g/m2)

0.18

Powder Weight (g/m2)

0.16
0.14
0.12
0.10
0.08
0.06
0.04
0.02
0.00

0.10

0.08

0.06

0.04

0.02

0.00
0

10

12

Hold Down Load (kN)

10

12

14

Hold Down Load (kN)

Figure 4.15: Powdering weight under the condition of drawing of the (a) STZC1, (b) STZC2,
(c) STZC3, (d) GI and (e) GA steels and (f) a comparison of the powdering rate in the five zinc
coatings

83

(b)
0.020

0.018

0.018

0.016

0.016

Powder Weight (g/m2)

Powder Weight (g/m2)

(a)
0.020

0.014
0.012
0.010
0.008
0.006
0.004
0.002

0.014
0.012
0.010
0.008
0.006
0.004
0.002

0.000

0.000
126

129

132

135

138

123

126

(c)

132

135

(d)

0.020

0.020

0.018

0.018

0.016

0.016

Powder Weight (g/m2)

Powder Weight (g/m2)

129

Blank Width (mm)

Blank Width (mm)

0.014
0.012
0.010
0.008
0.006
0.004

0.014
0.012
0.010
0.008
0.006
0.004
0.002

0.002

0.000

0.000
126

129

132

135

126

138

129

132

135

138

Blank Width (mm)

Blank Width (mm)

(e)

(f)

0.10

0.12

STZC1
STZC2
STZC3
GI
GA

0.09
0.10

Powder Weight (g/m2)

Powder Weight (g/m2)

0.08
0.07
0.06
0.05
0.04
0.03
0.02

0.08

0.06

0.04

0.02

0.01
0.00

0.00
126

129

132

135

126

138

129

132

135

138

Blank Width (mm)

Blank Width (mm)

Figure 4.16: Powdering weight calculated as a function of blank width from LDH tests of the
(a) STZC1, (b) STZC2, (c) STZC3, (d) GI and (e) GA steels and (f) a comparison of the
powdering rate in the five zinc coatings

84

4.4 Formability
The limiting dome height value, an effective predictor of formability was
measured using the limiting dome height test procedure described in section 3.3.3.1.
4.4.1. Limiting Dome Height Value
Limiting dome height (LDH) values of the zinc coated steel sheets were measured
using a LDH testing procedure. The results are given in terms of limiting dome height as
a function of blank width in Figure 4.17. The blank width with the lowest LDH value
(LDH0) corresponds to the plane strain condition and that width was set as the origin in
the Figure 4.17(f). The lower width and higher width correspond to a drawing condition
and a biaxial stretching condition, respectively. As a result, Figure 4.17(f) shows how
major strain varies as the strain states changes drawing (negative minor strain) to biaxial
stretching (positive minor strain). In other words, the LDH curves reveal a portion of a
forming limit diagram. However, the LDH curves in Figure 4.17 (f) are not quantitative
with regard to values for minor and major strains.
The LDH curves for the STZC samples [Fig.4.17 (a)~(c)] have a sharp V shape.
It implies that the strain state strongly affects the LDH values, i.e. the formability of the
zinc coated steel sheets. The slopes of LDH data depend on not only upon the strain state,
but also the characteristics of the zinc coating. It implies that the properties of the zinc
coating layer should be taken into account when considering overall formability. The
average LDH0 values for STZC1, STZC2 and STZC3 samples were 40.91.4mm,
37.61.0mm, 54.21.0mm, respectively.
The LDH curves for the GI and GA steels [Fig.4.17 (d),(e)] have a U shape and
85

the differences in the LDH values among the various strain states are lower than that of
the STZC steels. It implies that the sensitivity to strain state for the forming limit of the
GI and the GA steels is lower than that of the STZC steels. This result seems to be due to
the inhomogeneity of the steel sheet through its thickness. The STZC steels have an
phase layer, phase layer and the steel substrate layer whereas the GI and GA samples
have either a pure phase layer or a full Fe-Zn intermetallic phase layer and the steel
substrate. As a result, the GI and GA samples with a more homogeneously layered
structure respond less to a variation of strain state. The average LDH0 values of the GI
and the GA samples were 53.71.0mm and 35.21.0mm, respectively.
The LDH curves were normalized by setting each plane strain condition at the
origin for comparison purposes as shown in Figure 4.17 (f). The STZC 3 and the GI
steels showed the maximum formability and the GA steel showed the minimum
formability regardless of strain state. The formability of the STZC1 and STZC2 steel was
between those of the GI and the GA samples. It should be noted that the formability is a
mechanical property related to many parameters such as strain hardening component (nvalue), normal anisotropy (r-value) and the surface friction coefficient. A comparison of
the formability of five zinc coated steel sheets will be addressed in more detail in the next
chapter.

86

(a)

(b)

LIMITING DOME HEIGHT (mm)

LIMITING DOME HEIGHT (mm)

46

47
46

45
45
44

44
43

43

42
42
41
41
120

122

124

126

128

40

130

132

134

136

138

140

104

108

112

BLANK WIDTH (mm)

(c)

116

120

124

128

132

136

140

144

138

140

BLANK WIDTH (mm)

(d)

LIMITING DOME HEIGHT (mm)

LIMITING DOME HEIGHT (mm)


57

61
60

56

59
58

55

57
56

54

55
54

118

120

122

124

126

128

130

132

134

136

138

120

122

124

128

130

132

134

136

LIMITING DOME HEIGHT (mm)

(f)

LIMITING DOME HEIGHT (mm)

(e)

126

BLANK WIDTH (mm)

BLANK WIDTH (mm)

38.0

63

STZC1
STZC2
STZC3
GI
GA

37.5

60
57
54

37.0

51
36.5

48
45

36.0
42
39

35.5

36
128

130

132

134

136

138

140

142

144

146

148

150

-14 -12 -10 -8

-6

-4

-2

10 12 14

BLANK WIDTH (mm)

BLANK WIDTH (mm)

Figure 4.17: LDH curves for the (a) STZC1, (b) STZC2, (c) STZC3, (d) GI and (e) GA steels
and (f) curves normalized by setting each plane strain condition as an origin

87

5. Discussion
The surface friction, adhesion properties and formability of five zinc coated steel
sheets were analyzed. The next step is to clarify what factors affected surface friction
and coating adhesion and how the formability is correlated with surface friction and
coating adhesion. First, the factors affecting surface friction and adhesion propensities
will be discussed so that they can be controlled properly in zinc coated sheet steels. Next,
the formability will be correlated with surface friction and coating adhesion. Finally, how
the formability can be optimized via control of surface friction and coating adhesion will
be discussed.
5.1 Factors Affecting Surface Friction
There are many factors affecting surface friction either directly or indirectly.
Surface morphology and strain state affect directly surface friction because the former is
related to real contact area and the latter is related to modification of surface morphology,
i.e. roughness and creation of new surface. The microstructure of a zinc coating affects
indirectly surface friction because the combination of the type of phase and the geometry
of a phase affects mechanical properties such as the yield strength in shear. A galvanizing
condition like the aluminum content in the zinc bath and galvannealing conditions like
temperature and time affect indirectly the surface friction through their influence on the
microstructure.
5.1.1 Surface Morphology
The five zinc coated steel sheets experienced a modification to their surface
morphology via a skin pass process that imprinted their surface using a working roller
88

that had been textured by an EDT (Electrical Discharge Texturing) as described in


section 3.1.1. The morphology was investigated using SEM and the five zinc coatings
showed uniquely shaped asperities like a droplet or a mesa. The sum of the area of

0.25

Coefficient of friction

0.24
GI

STZC3

0.23
STZC2

0.22
0.21
STZC1

0.20
0.19
GA

0.18
20

25

30

35

40

45

Real area of contact,%


Figure 5.1: Effect of surface morphology on surface friction
an asperity, i.e. its real area of contact at zero normal load, was quantified by a point
count method as described in section 3.2.1. The real area of contact as a property
associated with a given surface morphology was compared with the surface friction
measured at a specific displacement under a biaxial stretching condition. The relationship
is shown in Figure 5.1. It is interesting to note that the surface friction coefficients are
proportional to the real area of contact for the STZC specimens and the GI specimen with
89

their phase asperities and that trend does not hold for the GA specimen. It implies that
surface friction could depend on not only the real area of contact as discussed by Bowden
and Tabor, but also other parameters such as the lubricant effect. The surface frictional
characteristics of the GA specimens coating with its higher hardness and yield strength
in shear and a surface morphology apparently composed of valleys and mesas, i.e. high
roughness, could be more strongly influenced by other parameters like the lubricant
effect rather than the real area of contact at a specific displacement. The frictional
characteristics of the zinc coatings with a low hardness and yield strength in shear and
low roughness including the STZC samples and the GI sample were more strongly
influenced by surface morphology at a specific displacement, showing a direct
relationship between coefficient of friction and the real area of contact.
5.1.2 Strain State
The surface friction was measured under different strain states like drawing and
stretching. Generally, the drawing condition and the stretching condition should be
considered differently. In a drawing operation as described in Fig.3.3, the regions
experiencing a frictional force are the flange, the die radius and the punch profile radius.
Throughout the flange region, in which the sheet is held with a normal force exerted by a
blank holder, sliding friction is predominant, and a pure drawing strain is present. In the
biaxial stretching condition as described in Fig. 3.5, metal sliding does not occur due to
constraints imposed by the draw beads. Plastic deformation continues until metal thinning
and ultimately failure take place. Due to stretching a large new area which has
intrinsically higher frictional characteristics is created during the deformation. That is one
90

of the reasons why the biaxial case has much higher surface friction than the drawing
case. According to several studies from Fox et al., Siegert et al., and Hira et al.
[39,40,41], surface friction under uniaxial deformation is higher than that under a
drawing strain. However, the increase of surface friction under uniaxial deformation is
lower than that from biaxial stretching because the new surface area created under the
uniaxial deformation is smaller than that under biaxial stretching. Surface friction as a
function of strain state is displayed in Figure 5.2 based upon values in the literature and
the present study.

0.26
0.24

GI[39]
GA[40]
P-GA[24,41]
Bare[25,40]

0.22
0.20

0.18

GI
GA

0.16
0.14
0.12
0.10
0.08

Drawing

Uniaxial

Biaxial

Strain State

Figure 5.2: Effect of different strain state on surface friction. Note GI: galvanized ,GA:
galvannealed, P-GA: galvannealed and phosphate treated, Bare: non-coated and open symbols:
present study (Source:Paik [24], Ghosh [25], Fox et al.[39], Siegert et al. [40], Hira et al. [41] )

Figure 5.2 shows that the coefficient of surface friction increases in going from a drawing
strain condition to a biaxial stretching condition for GI, GA, P-GA and bare steel surfaces.

91

The STZC steels which are not included in Figure 5.2 also exhibited a much higher
surface coefficient under biaxial stretching than under a drawing strain condition.
5.1.3 Microstructure
The microstructure of a zinc coating affects its surface friction in two ways. First,
it modifies the surface roughness and real area of contact for example when an Fe-Zn
intermetallic phases grow to the outer surface. A modification in the phase found at the
surface gives rise to a change of surface friction because the accompanying alteration to
the surface roughness and real area of contact are related directly to surface friction. L.G.
Garza et al. [42] showed that the surface friction of a galvannealed coated steel composed
of all phase is lower than that of a galvannealed coated steel with some amount of
phase on the surface.
Secondly, a change in the phase found at the surface modifies surface friction by
altering the yield strength in shear of the zinc coating. This factor applied to all five types
of steel coatings that were evaluated. The results of depth profiles of the five zinc
coatings [Fig. 4.4] and cross sectional analysis by SEM [Fig. 4.5] showed that the Fe-Zn
intermetallic phases such as the phase did not grow to the outer surface. According to
Bowden and Tabors friction theory, the surface friction is proportional to the yield
strength in shear of the surface layer. Thus, once the microstructure causes a
modification to the yield strength in shear of the coating, the surface friction changes.
The relationship between microstructure and yield strength in shear can be expressed
using a rule of mixture for composites and the Tresca yield criterion as follows.

92

=
y (T )
=
y (T )

y (T )

=
2

y (T )

=
2

y1V1 + y 2V2 + + yiVi


2

(5.1)

H1V1 + H 2V2 + + H iVi

(5.2)

where y (T ) = overall yield strength in shear in MPa, y (T ) = overall tensile yield strength
in MPa, y = tensile yield strength of phase i , Vi = volume fraction of phase i , = a
proportional factor, and H = Vickers hardness number of phase i . Equation (5.1) shows
that the greater the volume of a phase with a higher tensile yield strength that is contained
in a coating layer, the higher the overall yield strength in shear of the coating layer will
be. Because the tensile yield strength is proportional to hardness [43], equation 5.1 can be
expressed in the form of equation (5.2).
As a result, assumed the phase grows uniformly and continuously like Figure.5.3, the
magnitude of the yield strength in shear of the zinc coatings can be approximated as
follows.

=
y (T )

y (T )

=
2

t
tc

t2
ti
+ + Hi
tc
tc
2

H1 1 + H 2

(5.3)

where ti = thickness of phase i and tc = thickness of the overall coating layer. The yield
strength in shear of the five zinc coatings can be easily compared using equation 5.3 and
measured values for the hardness and thickness terms. The GI steel has the lowest yield
strength in shear because the zinc rich phase has the lowest hardness. The yield strength
in shear of the STZC steels depends on the volume fraction of phase with its higher
hardness than the phase. The yield strength in shear increases from the STZC1 steel to
93

the STZC3 steel because the thickness of the phase increases from the STZC1 steel to
the STZC3 steel. The GA steel has the highest yield strength in shear because the coating
is composed of the phase which is harder than the phase.
Figure 5.4 displays the coefficients of surface friction measured under a drawing
strain condition as a function of the average vertical height of the Fe-Zn intermetallic
phase present in each coating. The correlation between higher yield strength in shear (or
hardness) and increased coefficient of surface friction complies with the Bowden and
Tabors theory. The nonlinearity seems to be attributed to an imperfect match between
the average vertical height of the Fe-Zn intermetallic phase and the yield strength in shear.
Figure 5.4 implies that at a specific displacement, the predominant factor affecting
surface friction under a drawing condition could be due to the microstructure rather than
the lubricant or the surface morphology. Increasing surface friction is observed as the
average vertical height of the Fe-Zn intermetallic phase increases. Table 5.1 summarizes
the predominant factors affecting surface friction under specific conditions of biaxial
stretching and drawing strains.

Figure 5.3: A microstructure of zinc coating where an Fe-Zn intermetallic phase was grown
ideally
94

5.1.4 Galvanizing and Galvannealing Conditions


The thickness of the Fe-Zn intermetallic phases affects surface friction as
discussed in the previous section. The thickness can be controlled by modifying the
galvanizing conditions such as the aluminum content in the zinc bath or the
galvannealing conditions, e.g., annealing temperature and time. The aluminum is used to
control the formation of the Fe-Zn intermetallic phases and the galvannealing
temperature and time determine the thickness of the Fe-Zn intermetallic phases. Varying
the galvannealing conditions can affect indirectly the surface friction by modifying the
thickness of the Fe-Zn intermetallic phases. The thickness of a Fe-Zn intermetallic phase
affects a coatings yield strength in shear which in turn modifies the surface friction.
Assumed an overall coating layer is composed of two planar phases and one phase grows
at the expense of another phase as shown in Figure 5.3, the relationship between the
galvannealing temperature and the yield strength in shear of a zinc coating can be
obtained from equation 5.3 and Ficks 2nd law equation for diffusion as follows.

Q
Q
D0 exp
tc D0 exp
t
t
RT
RT

+ H2
H1

tc
tc

y (total )

=
=
y ( total )
2
2

(5.4)

where = a multiplier constant, D0 = the frequency factor, Q = the activation energy,


R =the gas constant, T = the galvannealing temperature and t = the galvannealing time.

Once several parameters corresponding to growth of a specific phase are determined, a


plot of the yield strength in shear as a function of galvannealing parameters can be
95

produced using equation 5.4. Figure 5.5 lists the parameters used and the variation of the
yield strength in shear for galvannealing temperatures ranging from 700 to 800K,
corresponding to phase growth. The yield strength in shear of a zinc coating increases
as the galvannealing temperature increases. Figure 5.5 indicates that the surface friction
should increase as the galvannealing temperature increases. The galvannealing
temperature can modify indirectly the surface friction of a coating by increasing the
coatings yield strength in shear through increasing the amount of phase in the coating.

96

0.18
0.17

Coefficient of Friction

0.16
0.15
STZC3

0.14
0.13
GA

0.12
0.11
STZC1

0.10
GI

0.09
STZC2

0.08
0

10

11

12

Average vertical height of Fe-Zn intermetallic Phase (m)


Shear Yield Strength of coating
Figure 5.4: Effect of average vertical height of Fe-Zn intermetallic phase on surface friction

Table 5.1 Predominant factors affecting surface friction under specific conditions
Steel Type

Shear Yield Strength

GI

Very Low

STZCs

STZC1
STZC2
STZC3

GA

Low
Intermediate
High

Very High

Strain Condition
Biaxial Stretching*
Drawing**
Microstructure
Surface Morphology
(Shear Yield
(Real Area of Contact)
Strength)
Microstructure
Surface Morphology
(Shear Yield
(Real Area of Contact)
Strength)
Forming
Condition
Forming Condition
(Lubricant)
(Lubricant)
Microstructure
(Shear Yield
Strength)

* punch displacement=25mm
** punch displacement=18mm

97

98

Figure 5.5: Effect of galvannealing temperature on shear yield strength of zinc coating [43,

44, 45]

H2= Vickers hardness number of =70 , H3= Vickers hardness


number of = 285

T=700K~800K, H1=Vickers hardness number of =112

Q=19.9kcal/mol , t=2sec, tc=15m, R=1.987 cal mol-1K-1

= 3.06 , = 2.7, D0=2.28 *10--6m2/sec

y (total) = 24675.84 exp(-1/T)5008 + 107.1

5.2 Factors affecting Coating Adhesion


There are many factors affecting the adhesion of a zinc coating either directly, or
indirectly. This study revealed that coating adhesion depends directly on the
microstructure of the zinc coating and the strain state. In addition, coating adhesion
depends indirectly on the galvannealing conditions, the steel substrate and surface friction.
In this section, how these factors affect coating adhesion will be discussed in detail.
5.2.1 Microstructure
The microstructure of a zinc coating directly affects coating adhesion because it
modifies the mechanical properties of the coating such as its fracture strength and strain
energy release rate. The brittle phase acts as an easy crack nucleation site which
decreases the fracture stress. The phase can nucleate a crack in the ductile matrix of the
phase after a sufficient load is applied. This can be expressed both in terms of fracture
stress and critical strain energy release rate as follows.

F =

=
G

2 E ( S + P )
t1

(5.5)

Y 2 2 t1
GIC
E

(5.6)

where F = fracture stress of the system, E = Youngs modulus of matrix, s = true


surface energy of matrix, p = plastic deformation energy of matrix, GIC = critical strain
energy release rate of system, Y = geometrical factor, = far field stress, and t1 = size of
pre-existing crack. Equation (5.5) corresponds to a fracture criterion for a non- ideally
brittle material called the Griffith- Orowan equation. The fracture stress of a system
99

decreases as the thickness of its brittle phase containing through thickness cracks
increases. Equation (5.6) presents this in terms of an energy release rate. Once the energy
release rate decreases due to the brittle phase and in turn exceeds the critical energy, the
system will fracture. Figure 5.6(a) shows powdering weight as a function of average
vertical height of phase in the STZC samples. The amount of powdering increases as
the average vertical height of phase, a brittle phase in + system increases. Figure
5.6(b) shows a comparison among the five zinc coatings. One can see that the powdering
weight of the GA is much larger than that of the other four steels. The coating of GA is
composed of only phase known as one of the most brittle phase in Fe-Zn intermetallic
phases and the coating itself acts as a crack initiation site. Figure 4.5 has already
represented that the GA involves many initial cracks of coating thickness. The effects of
Fe-Zn intermetallic phase on coating adhesion are illustrated schematically in Figure 5.7.
As a result, the coating adhesion depends on both the average vertical height and type of
Fe-Zn intermetallic phases and it is deteriorated as the average vertical height increases
and the brittleness of phase increases, i.e. K ICFeZn , GICFeZn .

100

(a)

Powdering Weight (g/m2)

0.0044
0.0040
0.0036

STZC3

0.0032
0.0028
0.0024
STZC1

0.0020
STZC2

0.0016
0.0012
0.0008
3

10

11

Average vertical height of phase (m)


(b)

Powdering Weight (g/m2)

0.10

GA

GI & STZC samples : 0.005g/m2


GA samples : 0.004g/m2

0.08

0.06

0.04

0.02
GI
STZC1

STZC2

STZC3

0.00
0

10

12

Average vertical height of Fe-Zn intermetallic phases (m)


Figure 5.6: Effect of average vertical height (a) and type (b) of Fe-Zn intermetallic phase
on coating adhesion for a drawing strain

101

GI steel

STZC steels

GA steel

Figure 5.7: The Fe-Zn intermetallic phases and surface oxides acting as initial crack sites

102

5.2.2 Strain State


The fracture of a zinc coating depends directly on its strain state because a
specific strain state makes a particular fracture mode operate. The five zinc coatings can
be divided to two categories in terms of their microstructure. The GI and the STZC steels
had an phase based coating and the GA steel had a phase based coating. Assuming
that both categories failed by a wedging mechanism, the fracture of an phase based
coatings is controlled by crack growth normal to the coating because crack propagation is
limited by crack tip blunting as a crack attempts to pass through a ductile phase matrix.
A crack oriented normally to a coating needs to grow until it approaches the interfacial
layer. The phase based coatings can fail under a specific strain state that induces both
severe plastic deformation and crack propagation to the interfacial layer in the system.
The fracture of the phase based coating is controlled by crack growth parallel to the
coating because of large preexisting cracks that lower the coatings fracture stress. As a
result, the phase based coatings can fail under a specific strain that has a strong
component of compression that induces crack propagation parallel to the coating/
substrate interface.
Figure 5.8 shows powdering weight as a function of strain states generated during
cup testing and LDH testing. In the case of the phase based coatings [Fig.5.8(a)], the
highest rates occurred under plane strain and stretching conditions. A study by R. Parisot
et al.[38] showed that the plane strain and stretching conditions meet the fracture
requirement for an phase based coating very well. In the case of the phase based
coating [Fig. 5.7(b)], the highest powdering rate occurred under a drawing condition. The
103

drawing condition meets the fracture requirement for the phase based coating. In fact,
the drawing condition involves lateral contraction that induces a strong compressive
strain so that crack propagation parallel to the direction of the coating /substrate interface
occurs more easily. Coating adhesion of the phase based coating appears to decrease
under a drawing condition. Figure 5.9 summarizes the effects of strain state on crack
propagation for the phase based and phase based coatings.

104

Powdering Weight (g/m2)

0.012

STZC1
STZC2
STZC3
GI

0.010

0.008

0.006

0.004

0.002

0.000
Pure Drawing

Plane Strain

Slightly Stretching

Strain State

0.12

Powdering Weight (g/m2)

b
0.10

0.08

0.06

0.04

0.02

0.00
Pure Drawing

Plane Strain

Slightly Stretching

Strain State
Figure 5.8: Effect of strain state on coating adhesion in phase based (a) and phase based
coating (b) based upon strain states generated during cup testing and LDH testing

105

Figure 5.9: Schematic illustration of crack propagation of phase based- and phase based
coating as fracture driving strains are subjected (Red line: crack initiation sites, blue line: crack
propagation)

106

5.2.3 Galvannealing Conditions


Galvannealing conditions like the galvannealing temperature and time affect
coating adhesion in two ways. First, the conditions are related to the growth of the Fe-Zn
intermetallic phases. An excessive growth of an Fe-Zn intermetallic phase increases the
brittleness of a coating which reduces coating adhesion. Secondly, the galvannealing
conditions are related to the growth of a surface oxide layer like ZnO. Figure 4.4 shows
the effect of galvannealing temperature on the formation of surface oxide. The STZC3
steel annealed at the highest temperature among the STZC steels has the largest
integrated area of oxygen among the STZC steels. The thickness of the zinc oxide
increases as the galvannealing temperature increases. The ZnO layer formed on the outer
surface can act as a crack initiation site because zinc oxide ( E 140GPa) is has a larger
elastic modulus than the phase ( E 75GPa). The STZC3 steel annealed at the highest
temperature experienced crack initiation at both outer and inner coating sites as and
correspondingly poorer coating adhesion. The powdering rate of the STZC3 steel was the
highest among the STZC steels.

107

5.2.4. Steel Substrate


The steel substrate affects coating adhesion in three ways. First, it modifies the
interfacial shear strength by modifying the microstructure in the interfacial layer.
According to a TEM study by C.S. Lin et al. [46], an IF grade had the phase and the
phase as an interfacial layer whereas a IFP grade had the phase and 1 phase as an
interfacial layer. The segregation of P to the surface of the steel substrate promotes the
nucleation of the 1 phase and prevents formation. Modification of the microstructure
in the interfacial layer affects its interfacial shear strength. The difference in lattice
parameter between the phase and the or 1 phase gives rise to a high lattice misfit,
and the interphase coherency depends on the misfit. Usually, if the misfit is larger than
0.25, an incoherent interface is formed. Unlike coherent or semicoherent interfaces, an
incoherent interface implies a lack of lattice continuity and poor bonding across an
interfacial plane. As a result, the interface possesses lower bonding strength and
corresponding lower interfacial shear strength. Table 5.2 lists calculated lattice misfit for
a / interface and a /1 interface. A /1 interface is close to a semicoherent interface
whereas a / interface is close to a truly incoherent interface.
Secondly, the steel substrate can modify the stress distribution through the
thickness of a zinc coating. According to microhardness measurements of Fe-Zn
intermetallic phases and their steel substrate (IF) from J. Foct et al. [45], the hardness
increases as the Fe content in an Fe-Zn intermetallic phase increases and the hardness of
the steel substrate (IF) is lower than that of Fe-Zn intermetallic phases with high Fe
contents like the phase and the phase. If the microstructure of a zinc coating has a
108

variety of Fe-Zn intermetallic phases and the coating is deposited on an IF grade steel
substrate, microhardness values for a region with Fe-Zn intermetallic phases located
between the phase and the soft steel substrate will be relatively high. The high hardness
of the , and 1 Fe-Zn intermetallic phases and their concave downward distribution
allow cracks to easily initiate and propagate. The STZC1 and the STZC2 steels are
examples of this effect. If a zinc coating is deposited on an IFP grade steel substrate, the
shape of the distribution of the hardness of Fe-Zn intermetallic phases is transformed
from concave downward to roughly linear. The embrittling effect by Fe-Zn intermetallic
phases can be reduced by a modification of the steel substrate. The STZC3 steel
corresponds to this situation. One of the reasons why the STZC3 steel shows relatively
good coating adhesion in spite of having a relatively large amount of phase is that its
IFP steel substrate reduces the stress raising effect of that Fe-Zn intermetallic phase.
Figure 5.10 illustrates the effect of the steel substrates on crack formation in the STZC
steels.
Finally, the steel substrate can affect coating adhesion through the influence of
interdiffusion of Fe and Zn between a zinc coating and its steel substrate. W. Zhong et al.
[49] reported that Zn diffusion along ferrite grain boundaries can occur during
galvannealing and these Zn atoms can lead grain boundary embrittlement in IF grade
steel substrates. The P solute segregated to ferrite grain boundaries in IFP grade steels
can retard Zn diffusion and reduce the potential for embrittlement of IFP grade steels by
Zn. Zn embrittlement can give rise to a flaking coating failure mode. A flaking coating
failure involves both the coating layer and a part of the steel substrate unlike powdering
109

where only a failure of the coating layer occurs. Zn embrittlement of an IF steel substrate
can reduce coating adhesion and the overall formability of the steel sheet.

Table 5.2 Calculated lattice misfit of interfacial layers [33, 47, 48]

Steel
Substrate

Interfacial
Layer

Interfacial
Shear
Strength
(MPa)

IF

14.7~15.2

IFP

/1

20

f* =

Lattice Misfit
( f *)

Lattice Parameter
( a , nm)

1.282
0.893
1.282
1.796

0.435
0.286

| as a f |
as

STZC1, STZC2

STZC3

Figure 5.10: Effect of steel substrate on coating adhesion in the STZC steels
110

5.2.5 Surface Friction


Surface friction which generates a tangential stress can affect coating adhesion in
two different ways. First, it can affect crack initiation from the extreme outer surface that
is composed of zinc oxide and aluminum oxide. This can be a key issue for the phase
based coatings like the GI and STZC steels because their coating adhesion depends on
how easily a fracture oriented normal to the direction of the coating can initiate and
propagate.
Secondly, surface friction can act as a source of shear stress that contributes to the
coating failure mechanism for the GA steel. The coating adhesion of the GA samples
depends on how easily fracture oriented along the coating/substrate interface is initiated
and propagates. Surface friction acts as an additional source of shear stress and assists
crack propagation and causes an increased level of powdering in the GA coating.
Surface friction can increase the level of powdering. The question to consider is
how much does it affect coating adhesion when compared to other factors like
microstructure, strain state and interfacial shear strength. If it is assumed that Amontons
law holds, i.e. = p , for a forming condition similar to that of cup testing, the
calculated friction stress of a pure zinc coating is only 0.2% of the shear yield strength of
the coating and only 2% of its interfacial shear strength [Appendix 2]. This implies that
friction-induced powdering is insignificant under drawing conditions.

111

5.3 Formability
It has been known that factors such as anisotropy and strain rate sensitivity
interact in a complex way to establish the formability of zinc coated steel sheets.
However, surface friction and coating adhesion also are very important parameters that
need to be taken into account. In this section, how surface friction and coating adhesion
affect formability in the various zinc coatings will be discussed.
5.3.1 Surface Friction
Surface friction can affect formability by modifying the strain distribution and the
strain state. When the relationship between surface friction and formability is considered,
it is important to consider only a constant strain state because surface friction is very
sensitive to strain state as discussed previously. A specific LDH value under a stretching
condition corresponding to the positive X- direction in Fig. 4.17(f) was matched with a
surface friction coefficient under the same condition in all five zinc coated steel sheets.
The result shown in Figure 5.11 indicates two different relationships between surface
friction and formability. The formability is inversely proportional to the surface friction
for the STZC1 and STZC2 steels whereas the formability is proportional to the surface
friction for the GI and GA steels. The conflicting result implies that the formability of
zinc coated steel sheets is not only a function of surface friction. Other factors like
coating adhesion must be taken into account. The steel types showing proportionality
between surface friction and LDH value (dashed circles in Figure 5.11) have either
different coating adhesion properties or a different fracture mechanism associated with
their coating. The GA steel has a much higher powdering value than the GI steel.
112

LDH
Surface Friction Coefficient

0.24

50

0.23

LDH(mm)

a
0.22
40
0.21

30

0.20

Surface Friction Coefficient

60

0.19
20
STZC1

STZC2

STZC3

GI

GA

Figure 5.11: Effect of surface friction on formability in various zinc coated steel sheets

The STZC2 steel with an IF grade steel substrate can exhibit damage to its steel substrate
by zincembrittlement whereas the STZC3 steel with an IFP grade steel substrate should
resist such damage. The steel types will not show a consistent relationship between
surface friction and formability if other factors are variable. R. Stevenson [50] also noted
in his work that changes in the punch-sheet coefficient of friction cannot account for the
changes in formability of galvanized steel sheet when the sheets have different levels of
coating adhesion. The STZC3 and GI steels (solid circle b in Figure 5.11) also do not
show an inverse proportionality between surface friction and formability. Even though
the coating adhesion properties and the fracture mechanism of these two types of steel are
very similar, the difference in surface friction between the two steels is too slight to show
113

an effect. The STZC1 and STZC2 steels (solid circle a in Figure 5.11) show an
inversely proportional relationship between surface friction and formability. These two
steels have not only a very similar coating adhesion and fracture mechanism due to
similar Fe-Zn intermetallic phase thicknesses and the same type of steel substrates so that
the effect of surface friction on formability is isolated, but also a significant difference in
surface friction. It can be concluded that high surface friction reduces the formability of
zinc coated steel sheets. Amit K. Ghosh [25] also showed that a reduction of friction
gives rise to an improvement in dome height at failure and that result holds universally
for a variety of materials.
The negative effects of surface friction on formability can be explained in two
ways. First, surface friction can modify the strain distribution during the forming of a
zinc coated steel sheet by acting as an additional constraint. The modification to the strain
distribution is defined in the following equation by Ghosh and Hecker [51].
ln

(2 z0 ) sin 2 ( s / ) r
= r + + f
P

(5.7)

where = punch radius, z0 = initial material thickness, s = arc distance between any
material point and the pole of the punch semi-arc length of contact, r = radial stress,
P = punch load which is the sum of components of pressure and friction, r = major

strain, =minor strain, f = strain modification by frictional stress. According to


equation (5.7), the term representing the strain modification by frictional stress affects
either the major strain or the minor strain producing an increasingly non-uniform strain
distribution, i.e. strain localization. Localized necking induced by frictional constraint
114

LIMITING DOME HEIGHT (mm)


63

STZC1
STZC2
STZC3
GI
GA

60
57
54
51
48
45
42
39
36

-14 -12 -10

-8

-6

-4

-2

10

12

14

BLANK WIDTH (mm)

Figure 5.12: Effect of a modification of strain state by frictional constraint on formability

contributes to premature fracture and decreased formability. Secondly, surface friction


can modify the strain state that develops during the forming of a zinc coated steel sheet.
When a zinc coated steel is formed under a stretching condition, the redistribution of
strain due to frictional constraint can give rise to a plane strain condition which causes a
decrease in formability. Figure 5.12 shows schematically the effect of a modification of
strain state by frictional constraint on formability. A relationship between a LDH value
and the coefficient of friction can be deduced from Ghoshs equations [25, 27] as follows.
LDH0 ( = 0 ) = LDH0 ( 0 ) +

C (sin c cos c c )
sin c {sin 2 c + ( c sin c cos c )}
2

(Appendix 3)

where LDH0 = Limiting dome height value at plane strain, = coefficient of surface
friction, c = the value of at the interface boundary defined in Fig.3.6 and C =
proportionality factor
115

5.3.2 Coating Adhesion


A coating layer can affect formability in two ways. First, it modifies a distribution
of strain during the forming of a zinc coated steel sheet. Some of the energy that drives
fracture is absorbed in the ductile zinc layer which lowers the energy driving fracture of
the steel substrate compared with an uncoated substrate and this improves tensile
properties like elongation. According to Stevensons equation [52], the formability of
sheet metal is roughly proportional to its total elongation. A ductile zinc coating can
improve the formability of a steel sheet by improving its elongation. However, the
change of elongation caused by a thin coating layer (~2%) does not modify the
formability appreciably. Both the strain effect caused by a thin coating layer and an effect
due to a loss of coating adhesion typically produce a very minor changes in the
formability of a steel sheet.
Secondly, a coating layer can affect formability due to its influence on the fracture
mechanism of the sheet steel. It is possible that some part of the coating layer, especially
an Fe-Zn intermetallic phase, can act as a stress raiser and assist in crack initiation in the
steel substrate. This could especially be true in an IF grade steel that experienced zinc
embrittlement. The formability of a zinc coated sheet steel strongly depends on the
average vertical height of its Fe-Zn intermetallic phase. Its formability decreases as the
average vertical height of its Fe-Zn intermetallic phase increases regardless of the
magnitude of its surface friction [Fig.5.13(a)]. This implies that the average vertical
height of an Fe-Zn intermetallic phase adds to the total potential crack length which then
lowers the fracture strength of the steel. The decrease of fracture strength of the sheet
116

steel gives rise to premature fracture lowering the steels formability. However, if crack
propagation is prevented at the steel substrate like an IFP grade which has no zinc
embrittlement, this can give rise to a huge improvement in formability regardless of the
average vertical height of the Fe-Zn intermetallic phases [Fig. 5.13 (b)]. A loss of coating
adhesion by a combination of fracturing of the coating layer and the steel substrate
(flaking) can be a much better indicator of formability than simple powdering weight.
Finally, a relationship between the LDH value and coating adhesion is presented in
equation 5.8.
LDH0

1
(FW)C

(5.8)

where FW= a loss of coating by a combination of fracturing of the coating layer and the
steel substrate, i.e. flaking weight, and C= frequency factor of flaking 1/{the at.%
phosphorus at steel substrate}.

55

55

50

LDH0

LDH0

50

45

45

40

40

35

35
0

10

12

Average vertical height of Fe-Zn intermetallic Phase (m)

10

12

Average vertical height of Fe-Zn intermetallic phase (m)

Figure 5.13: Effects of coating layer and steel substrate on formability; a) effect of average
vertical height of Fe-Zn intermetallic phase, b) effect of steel substrate (IF and IFP *)

117

6. Conclusions
There were three objectives to this study : 1.) to evaluate surface friction 2.) to
assess coating adhesion of differently treated zinc coated steel sheet subjected to different
strain states and 3.) to analyze factors correlated with surface friction and coating
adhesion. The conclusions of this study are listed below:
(I)

Five types of zinc coated sheet steels were evaluated for their surface friction and

coating adhesion properties. Three of the steels (STZC1, STZC2, STZC3) were designed
to have zinc coatings with iron and phase contents intermediate to those of -phase
galvanneal (GA) and - phase galvanized (GI) coatings. The STZC steels experienced an
intermediate galvannealing process and had microstructure containing + phases. The
phase with 6~9 at% iron grew at the expense of the phase. The STZC1 steel that
experienced a galvannealing process at the lowest temperature had a thin phase whereas
the STZC3 steel that experienced a galvannealing process at the highest temperature had
a thick phase. The GI and GA steels used for comparison purpose had pure phase and
pure phase microstructures, respectively. The aluminum used for suppressing Fe-Zn
intermetallic phases concentrated in two specific areas, the extreme outer surface and the
interface between a zinc coating and its steel substrate. The GI and STZC1 steels with a
thin Fe-Zn intermetallic phase exhibited a high aluminum deposition toward their
extreme outer surface whereas the STZC3 and GA steels with thick Fe-Zn intermetallic
phases exhibited lower concentration of aluminum both at their extreme outer surface and
the interface between their zinc coatings and their substrates. After being subjected to an
electric discharge texturing process, the surface of five zinc coatings showed specific
118

morphologies, affecting the shape of the asperities and real area of contact of their zinc
coatings.
( II )

The surface friction coefficient was measured under different strain conditions

such as drawing and stretching. The coefficient of friction of the five types of zinc
coated steel sheets observed during drawing strain state cup tests exhibited a dynamic
characteristic where surface friction varied as a function of punch displacement due to
deformation of surface asperities and lubrication effects. The measured coefficient of
friction from the STZC1, STZC2, and GI steels with a small amounts of Fe-Zn
intermetallic phase ranged from 0.09 to 0.11 whereas that from the STZC3 and GA steels
with larger volume fractions of Fe-Zn intermetallic phase ranged from 0.13 to 0.14.
These values were obtained at a punch displacement of 18mm. This result indicates that
modification of the mechanical properties of a zinc coating by a change of microstructure,
especially the shear yield strength affects surface friction, agreeing with Bowden &
Tabors adhesion theory. Under a stretching condition, the coefficient of frictions ranged
from 0.19 to 0.24, higher values than those obtained under a drawing condition.
Stretching produces a significant amount of new surface which has intrinsically higher
frictional characteristics. Under stretching condition, the coefficients from the GI and
STZC steels were proportional to the real area of contact whereas the GA steel had a
coefficient of surface friction that depended inversely upon real contact area. The
predominant frictional factor was the real area of contact for the GI and STZC steels with
their low shear yield strength whereas a lubrication effect was the primary frictional
factor for the GA steel with its high shear yield strength coating.
119

( III )

The powdering weight was measured under different strain states such as

drawing, modest drawing, plane strain, and modest stretching. Under a drawing strain
condition, the values of the powder weight of the STZC, GI and GA steels were
0.002~0.0035g/m2, 0.006g/m2, 0.1g/m2, respectively. The powdering weight increased as
the volume fraction of phase increased in the coating layer among STZC steel sheets. A
high iron content in a zinc coating ( 12.2 at% ) increases the level of powdering based
upon the results from the GA steels. The GA steel showed the highest powdering weight
among the five steel types regardless of the strain state. The iron content in the zinc
coating appears to be the strongest factor affecting coating adhesion. The STZC steels
and the GI steel exhibited higher powdering weights under plane strain and stretching
conditions than under a drawing condition whereas the GA steel exhibited higher
powdering weight under a drawing condition than under the plane strain and stretching
conditions. The fracture mechanism could depend on microstructure so that the soft
phase based coatings fractured by a tension-controlled mode whereas the less ductile
phase base coatings fractured by a compression- controlled mode.
( IV )

The surface morphology, strain state, and shear yield strength of a zinc coating

are direct factors affecting surface friction whereas the microstructure of a zinc coating
and galvannealing condition are indirect factors affecting surface friction. Stretching
deformation was associated with increased surface friction regardless of the displacement
and type of zinc coating due to the continuous creation of new surface. An increase of
real area of contact and volume fraction of an Fe-Zn intermetallic phase increased surface
friction at a specific displacement where lubrication effects were negligible.
120

(V)

Coating adhesion depends directly on the microstructure of the zinc coating and

the strain state. It depends indirectly on the galvannealing condition, the steel substrate
and surface friction. The effect of each factor on coating adhesion can be summarized as
follows. The coating adhesion deteriorated as the average vertical height and brittleness
of the Fe-Zn intermetallic phase increased. phase based coating adhesion deteriorated
under a drawing condition. phase based coating adhesion deteriorated under plane
strain or stretching conditions. Coating adhesion deteriorated due to excessive
galvannealing processing which increased the volume fraction of the Fe-Zn intermetallic
phases and thickened the brittle oxide layers on the extreme outer surface of the coatings.
Coating adhesion to the IFP grade steel substrate could be improved by forming a /1
semicoherent interface so that a much higher interfacial shear strength is achieved and
zinc embrittlement in steel substrate is suppressed.
( VI )

Surface friction could affect the formability of zinc coated steel sheet by

affecting the strain state. Strain modification by a frictional stress affects either the major
strain or minor strain producing an increasingly non-uniform strain distribution, i.e. strain
localization. Localized necking by frictional constraint contributes to premature fracture
and reduced formability. Even though a zinc coated steel is formed under a stretching
condition, the redistribution of strain by frictional constraint can give rise to a plane strain
condition resulting in reduced formability.
( VII )

Coating adhesion can affect formability by influencing the fracture mechanism

of the sheet steel. An Fe-Zn intermetallic phase that degrades coating adhesion can assist
crack initiation in a steel substrate due to a zinc embrittlement phenomenon that
121

decreases the fracture strength of the zinc coated steel sheet. It was suggested that a
flaking weight, i.e. an evaluation of the significance of crack initiation in the steel
substrate, should be taken into account to establish a relationship between coating
adhesion and the formability of zinc coated steels.

122

7. Appendix 1

Table A.1 Prediction of the number of fields (n) to be observed as a function of the desired
relative accuracy and of the estimated magnitude of the areal fraction of the constituent

Amount of
areal
fraction, Af
(%)
2
5
10
20

20% Relative Accuracy

10% Relative Accuracy

Number of fields (n) for a grid of


PT =

Number of fields (n) for a grid of


PT =

16
310
125
65
30

25
200
80
40
20

49
105
40
20
10

100
50
20
10
5

16
1250
500
250
125

25
800
320
160
80

49
410
165
85
40

Table A.2 95% Confidence interval multipliers


No. of Fields n
5
6
7
8
9
10
11
12
13

t
2.776
2.571
2.447
2.365
2.306
2.262
2.228
2.201
2.179

No. of Fields n
19
20
21
22
23
24
25
26
27

123

t
2.101
2.093
2.086
2.080
2.074
2.069
2.064
2.060
2.056

100
200
80
40
20

7. Appendix 2

Table A.3 Data used for calculation of ratio of a shear stress due to friction to a shear yield
strength of pure zinc coating and an interfacial shear strength of pure zinc coating

Tensile yield
strength[44,45]
( y )

Shear yield
strength

Coefficient of
friction () at
drawing strain

Hold down load


( FN )

0.10

10kN

74mm

214MPa

107MPa

Interfacial
shear strength
( i )

Area of blank

Shear stress
applied due to
friction
( f = p )

Ratio

D
(A= )
2

Normal stress
applied
F
(p= N )
A

14MPa

4.3 10-3m2

2.3MPa

0.233MPa

Diameter of
blank ( D )

124

( y =

f
y

y
2

f
i

0.002 0.017

7. Appendix 3

Ghosh [25] expressed the normal pressure ( p ) exerted by hemispherical punch as


p=

P
{sin c + ( c sin c cos c )}
2

where P = punch load, = punch radius, c = the value of at the interface boundary
defined in Fig.3.6 and = coefficient of surface friction
Now represent both the normal pressure under free friction, i.e. p ( =0), and the normal
pressure under friction, i.e. p ( 0). Then
P
{sin c + ( c sin c cos c )}

p ( 0) = 2

and
p ( = 0)
=

P
(sin 2 c )
2

p= p ( 0) p ( = 0)

=
where C is

C (sin c cos c c )
sin 2 c {(sin 2 c + ( c sin c cos c )}

Assuming then this pressure difference ( p ) is proportional to the change in LDH, one
obtains
LDH0( = 0 ) =LDH0( 0 ) +

C (sin c cos c c )
sin 2 c {sin 2 c + ( c sin c cos c )}

where LDH0= Limiting dome height at plane strain and C= Proportionality factor

125

8. References
[1] Burghardt, A.J. et al. , Galvatech 92 Conference Proceedings, 189, 1992.
[2] Irie, T, Zinc Based Steel Coating Systems: Metallurgy and Performance, TMS, 143,
1990
[3] Bowden, F. P. and Tabor, D. , Proceedings of the Royal Society A, 169 , 391 , 1939.
[4] Tabor, D. , Surface and Colloid Science, 5, 245-312, 1972.
[5] Orowan, E., Proceedings of the Institution of Mechanical Engineers, 150, 140, 1943.
[6] Shaw, M. C. and Mamin, P. A., Journal of Basic Engineering, 82, 342, 1960.
[7] Rabinowicz, E., Journal of Applied Physics, 22, 1373-48, 1951.
[8] Dalton, G. , SAE Technical Paper No. 2001-01-0081, Warrendale, PA, USA. ,pp.1-6,
2001.
[9] Stribeck, H. , Zeitschrift Des Vereines Seutscher Ingenieure, 46, 1342-48, 1902.
[10] Hess, D. and Soom, A., Journal of Tribology, 112, 147-152, 1990.
[11] Bay and Wanheim, T. , Wear, 43, 45, 1977.
[12] Alpas, A. T. and Inagaki, J. , ISIJ International, 40, 172-181, 2000.
[13] Amey, S. L. , Metallurgical and Materials Transactions A, 25A, 723-731, 1994.
[14] Janavicius, P. V., Galvatech 95 Conference Proceedings, Sept. ,17-21, 1995.
[15] Jordan, C.E. , Metallurgical and Material Transactions A, 25A, 2101-2110, 1994.
[16] Takechi, H. , ISIJ International, 34(1), 1-8, 1994.
[17] Rege, J. et al. , IF steels 2000 Proceedings, Pittsburgh, PA, 327-338, 2000.

126

[18] Karduck, P. , Fresenius Journal of Analytical Chemistry, 358, 135-140, 1997.


[19] American Society of Testing Materials (ASTM), E 562-05.
[20] Kumpulainen, J.O. et al., Journal of Engineering Materials and Technology,105,
119-127, 1983.
[21] Doege, E. , TMS-AIME, Warrendale, PA, 209-224, 1986.
[22] Lange, K. , Handbook of Metal Forming, McGraw Hill, New York, pp.209, 1985.
[23] Swift, H. and Chung, S., Proceedings of the institution of Mechanical Engineers, 165,
199-228, 1951.
[24] Paik, D, Ph.D Thesis in Case Western Reserve University, pp.33, 2002.
[25] Ghosh, A. , International Journal of Mechanical Science, 19, 457-470, 1977.
[26] Wang, N. , Journal of Applied Mechanics, 37, 431, 1970.
[27] Ghosh, A. , Metals Engineering Quarterly, 15, 53, 1975.
[28] Matsuda, H. et al. , Yosetsu Gakkai Ronbunshu, 14, 47-54, 1996.
[29] Wanheim, T. et al. , Annals of the CIRP. 27, 189-94, 1978.
[30] Feliu Jr., S and Barranco, V. , Acta Materialia, 51, 5413-5424, 2003.
[31] Raghavan, V., Journal of Phase Equilibria, 24(No.6), 544-545, 2003.
[32] Ghosh, G., Materials Science International Team (MSIT), 11, 21-41, 2005.
[33] Rolland, P. et al., Microscopy and Microanalysis, Savannah, GA, Aug 1-5, 2004.
[34] Keum, Y. T. et al. , AIP conference proceedings, 712, 989-994, 2004.
[35] McMahon, C. J. and Cohen, M. , Acta Metallurgica , 13, 591-604,1965.

127

[36] Inagaki, J. et al. , SAE Technical Paper Series 890349, Warrendale, 1989.
[37] Nakamori, T. et al. , In Corrosion Resistant Automotive Sheet Steels, ASM, Metals
Park, 139, 1988.
[38] Parisot, R. et al., Metallurgical and Materials Transactions A : Physical Metallurgy
and Materials Science, 35A, 813-823, 2003.
[39] Fox, R. T. et al., Metallurgical Transactions A: Physical Metallurgy and Materials
Science, 20A, 2179-82, 1989.
[40] Siegert, K., et al., Society of Automotive Engineers [Special Publication]SP, SP1132,1996.
[41] Hira, T. et al., Sosei to Kako, 34, 1141-6,1993.
[42] Garza, L. G. and Van Tyne C. J., Journal of Materials Processing Technology, 187,
164-168, 2007.
[43] Busby, T. J. et al., Journal of Nuclear Materials, 336, 267-278, 2005.
[44] Wakamatsu, Y. et al., J. Jap. Inst. Met., 39, 903, 1975.
[45] Foct, J., Scripta Metall. Mater. 28, 127,1993.
[46] Lin, C. S. and Meshii, M., Mechanical Working and Steel Processing Conference
Proceedings, 31, 99-105, 1994.
[47] Wen, B. et al., Physical Review Letters, 101, 2008.
[48] Tu, King-Ning et al., Electronic Thin Film Science, Macmillan Publishing Company,
1992.
[49] Zhong, W. et al., Zinc-Based Steel Coating Systems: Production and Performance
Conference Proceedings, 185-194, 1998.
[50] Stevenson, R., Research Publication, GMR-4283, 1983.

128

[51] Ghosh, A. and Hecker, S., Metallurgical Transactions A, 6A, 1065, 1975.
[52] Stevenson, R., Journal of Applied Metal Working, 3, 1984.

129

Anda mungkin juga menyukai