You are on page 1of 180

© 2007

MOHAMMAD FAIZAN

ALL RIGHTS RESERVED


EXPERIMENTAL STUDY AND MODELING OF METAL DISSOLUTION AND

INTERMETALLIC COMPOUND GROWTH DURING SOLDERING

A Dissertation

Presented to

The Graduate Faculty of The University of Akron

In Partial Fulfillment

of the Requirements for the Degree

Doctor of Philosophy

Mohammad Faizan

December, 2007
EXPERIMENTAL STUDY AND MODELING OF METAL DISSOLUTION AND

INTERMETALLIC COMPOUND GROWTH DURING SOLDERING

Mohammad Faizan

Dissertation

Approved: Accepted:

______________________________ ______________________________
Advisor Department Chair
Dr. Guo-Xiang Wang Dr. Celal Batur

______________________________ ______________________________
Committee Member Dean of the College
Dr. T. S. Srivatsan Dr. George K. Haritos

______________________________ ______________________________
Committee Member Dean of the Graduate School
Dr. Alex Povitsky Dr. George R. Newkome

______________________________ ______________________________
Committee Member Date
Dr. Ernian Pan

______________________________
Committee Member
Dr. Thein Kyu

ii
ABSTRACT

Soldering has become the predominant and established technique in electronic

packaging industry for joining electronic components. The industry is aiming for the use of

environment friendly lead-free solders. All the lead-free solders are high tin-containing

alloys. During the soldering process, an intense interaction of metallization on PCB and tin

from the solder occurs at the joint interface. Intermetallic compound (IMC) is formed at the

interface and subsequently PCB bond-metal (substrate) is dissolved into the molten solder.

The formation of the IMC is desirable for the joint to occur but excessive metal dissolution

and uncontrolled growth of the IMC is detrimental to quality of joint. In the present study the

terms bond-metal and substrate will be used interchangeably and the term ‘substrate’ refers to

the top layer of the PCB which comes in contact with the molten solder during soldering

reaction. During the wet phase of soldering process, the IMC exhibits scalloped

morphology. The growth of scalloped IMC during the solder/substrate interaction entails

complicated physics especially during early stages of the soldering process.

Understanding of the physics and the kinetics involved in the formation of IMC phase

and metal dissolution is important in achieving the desired control of the process.

This study presents an in-depth analysis of the physics involved in the formation

of the scalloped intermetallic phase. To substantiate the physics of the process and to

verify the kinetics involved, a mathematical model is setup. The actual physics involved

iii
in the dissolution of metal layer and growth of IMC during soldering process is translated

into mathematical model based on the fundamental equations and known material

properties and thermodynamic constraints. The model was applied to study the

interaction of Cu with Sn3.5%Ag solder as a representative example however, the model

can be applied to any substrate/solder system by incorporating the relevant physical

properties of material. The model served as a valuable tool to closely understand the

physics involved in the process of diffusion of various species across the joint, phase

change and kinetics effects during soldering process. Extensive parametric study was

conducted to explore the relative effect of various parameters like interface reaction

kinetics and physical properties such as the diffusivity of solder and IMC. The effect of

these parameters on the dissolution behavior of metallization layer and the growth of

IMC with time during soldering process was also studied.

The results from the model were also compared with experimental data for IMC

thickness and metal dissolution available in literature. Effects of non-isothermal

processing conditions were also considered in the model. A constant value of IMC

diffusivity was able to accurately predict the IMC growth for the cases involving planar

IMC layer, however, the single value of IMC diffusivity did not work very well for

scalloped IMC layer. Due to the complicated evolution of the scalloped IMC growth in

wet phase of soldering, it was suggested that the diffusivity in the IMC phase changes

with time and a varying value of IMC diffusivity be used in the model. The varying IMC

diffusivity gave satisfactory predictions for the IMC thickness and metallization layer lost

during reflow process. The model was also applied to study the dissolution behavior of

the micro-size copper particles in lead-free composite solders during reflow process.

iv
DEDICATION

Dedicated to my

Parents

and

Family

v
ACKNOWLEDGEMENTS

All praise to Allah (swt) who taught the mankind what they did not know. My

prayers for my deceased parents who gave countless sacrifice and did every effort in

reach to nurture me and provided the highest moral values. My sincere regards to my

advisor in the present work, Dr. Guo-Xiang Wang who always helped and supported me

in the time when I needed him. I am also thankful to the committee members for their

valuable suggestions in bringing out the present work. I feel obliged to express my

gratitude towards my brother, Dr. Shakir Husain who launched my academic career while

I was just ten year old. His continuous support, encouragement and valuable advice on

every occasion of my life helped me in achieving high goals in life. This letter would

remain incomplete if I did not mention the support, sacrifice and affection of my wife and

two lovely daughters, Zahra and Maria. Their presence always gave me strength and

courage to overcome every hurdle I faced in these years.

Mohammad Faizan

December 14, 2007

vi
TABLE OF CONTENTS
Page

LIST OF FIGURES...........................................................................................................xii

CHAPTER

I. INTRODUCTION............................................................................................................1

1.1 Kinetics of IMC Formation and Substrate Dissolution......................................3

1.2 Objective of the Study.......................................................................................5

II. LITERATURE REVIEW................................................................................................7

2.1 Global Efforts for Lead-Free Solders................................................................7

2.2 Significance of Solder Materials and their Selection Criterion.......................10

2.2.1 Environmental Concern....................................................................11

2.2.2 Process Conditions of Electronic Assembly.....................................13

2.2.3 Wetting Properties of Solder.............................................................13

2.2.4 Reliability and Cost...........................................................................15

2.3 Soldering Techniques in Electronic Packaging Industry.................................17

2.3.1 Reflow Soldering..............................................................................17

2.3.2 Wave soldering.................................................................................18

2.4 Anatomy of Solder Joint and Intermetallic Compound Layer.........................19

2.5 Challenges in Processing of Lead-Free Solder Joints......................................20

2.5.1 Controlled Growth of Intermetallic Phase........................................21

vii
2.5.2 Need to Conserve Metallization Layer.............................................22

2.6 Growth Morphology and Kinetics of IMC Layer............................................23

2.6.1 During Early Stages of Soldering Process........................................24

2.6.2 Diffusion of Elements across Solder Joint........................................26

2.6.3 Evolution of Scalloped Morphology of IMC....................................26

2.6.4 Interfacial Energy of IMC Grains.....................................................28

2.6.5 Heating/Cooling Rate of Assembly..................................................29

2.7 Factors Governing the IMC Thickness and Substrate Dissolution


during Soldering...............................................................................................31

2.7.1 Composition of Base Material..........................................................31

2.7.2 Process Temperature/Time...............................................................32

2.7.3 Volume of Solder..............................................................................33

2.7.4 Composition of the Solder Alloy......................................................34

2.7.5 Morphology of Intermetallic Phase..................................................34

2.8 Recent Progress in Understanding of IMC Growth Kinetics..........................36

2.9 Composite Solders...........................................................................................38

2.9.1 Recent Developments in Composite Solder Research......................39

III. MODELING HISTORY OF IMC GROWTH AND METAL DISSOLUTION.........43

3.1 Power-Law Relation........................................................................................43

3.2 Modifications Made in the Power Law............................................................47

3.3 Understanding the Apparent Diffusivity..........................................................49

3.4 Substrate Consumption Rate during Soldering................................................50

3.5 Kinetics of Scalloped Morphology of IMC.....................................................52

viii
3.6 Size Distribution of IMC Scallops...................................................................55

3.7 Soldering Process as Stefan Problem...............................................................57

IV. EXPERIMENTAL STUDY TO INVESTIGATE IMC GROWTH AND COPPER


DISSOLUTION...........................................................................................................60

4.1 Reflow Experiments.........................................................................................60

4.2 Dipping Experiments.......................................................................................62

4.3 Measurement of Copper Dissolution and IMC Layer Thickness....................64

4.4 General Behavior of the IMC Growth and Copper Dissolution......................66

4.5 Effect of Solder Volume on IMC Growth and Cu Dissolution.......................67

4.6 Summary of Experimental Study....................................................................68

V. PHYSICAL MODELING AND MATHEMATICAL FORMULATION OF IMC


GROWTH AND METAL DISSOLUTION DURING SOLDERING........................70

5.1 Kinetics of Substrate Dissolution and IMC Growth........................................72

5.2 IMC Growth as Moving Boundary Problem...................................................75

5.3 Mathematical Formulation of Substrate Dissolution and IMC Growth...........76

5.3.1 Assumptions made in the Mathematical Model................................78

5.3.2 Governing Equations and Boundary Conditions..............................78

5.4 Solution of Mathematical Model.....................................................................82

5.5 Numerical Scheme...........................................................................................83

VI. MODEL VALIDATION AND PARAMETRIC ANALYSIS....................................85

6.1 Comparison of Analytical and Numerical Solutions.......................................85

6.2 Parametric Analysis of Interface Kinetics and Diffusivity..............................87

6.2.1 Effect of Interface Reaction Kinetics................................................88

6.2.2 Effect of Substrate (copper) Diffusivity in Solder............................93

ix
6.2.3 Effect of IMC Diffusivity.................................................................97

VII. NUMERICAL RESULTS FOR IMC THICKNESS AND METAL


DISSOLUTION DURING SOLDERING .............................................................105

7.1 Dissolution of Planar Substrate without IMC Formation..............................105

7.2 Numerical Results of IMC Thickness for Planar IMC Growth.....................108

7.3 Model Predictions for IMC Thickness and Copper Dissolution for
Scalloped IMC Growth..................................................................................110

7.4 Variable IMC Diffusivity...............................................................................112

VIII. MODEL APPLICATION TO STUDY THE DISSOLUTION BEHAVIOR


OF MICRO-SIZE METAL PARTICLES IN COMPOSITE SOLDERS...............121

8.1 Mathematical Model for Micro-Particles/Solder Interaction.........................122

8.2 Effect of Inter-Particle Distance on Particle Dissolution...............................124

8.3 Effect of Interface Kinetics on Particle Dissolution......................................126

8.4 Effect of Initial Copper Content in Sn-Ag-Cu (SAC) Alloy.........................129

8.5 Growth of Intermetallic Compound...............................................................131

8.6 Effect of Non-isothermal Reflow Cycle on Particle Dissolution and


IMC Growth....... ...........................................................................................133

8.6.1 Effect of Heating Rate....................................................................135

8.6.2 Effect of Cooling Rate....................................................................137

8.6.3 Effect of Holding Time...................................................................139

8.6.4 Effect of Initial Copper Content in Sn-Ag-Cu Alloy......................140

8.7 Summary of Dissolution Behavior of Micro-Particles in


composite solders…………………………………………………………...143

IX. CONCLUSIONS AND FUTURE WORK................................................................144

9.1 Model Development and Parametric Study...................................................145

x
9.2 Application of Numerical Model...................................................................149

9.3 Future Work...................................................................................................150

BIBLIOGRAPHY............................................................................................................152

APPENDICES………………………………………………………………………….158

APPENDIX A. COORDINATE TRANSFORMATION...............................................159

APPENDIX B. DISCRETIZATION OF DIFFERENTIAL EQUATIONS...................161

xi
LIST OF FIGURES

Figure Page

1.1 Micrographs showing the section view and top view of the IMC...……..……......4

2.1 Typical setup for Plastic Ball Grid Array (PBGA)………………………………..9

2.2 Representation of the degree of wetting in terms of the contact angle…………..14

2.3 Schematic diagram showing the principle of reflow soldering………………….17

2.4 Schematic diagram showing the sequence of wave soldering…………………...18

2.5 Micrograph showing a typical IMC layer in a soldered joint……………………19

2.6 Schematic representation of the wetting angle and force balance at the
IMC/solder interface……………………………………………………………..29

2.7 Cu-Sn Phase diagram…………………………………………………………….35

2.8 Proposed schematic showing inter-phase diffusion of Cu in Sn…………………37

3.1 Arrhenius plot to estimate activation energy and apparent diffusivity..................45

4.1 Schematic showing experimental setup for reflow experiments………………...61

4.2 Real time-temperature curve of a typical reflow process………………………..62

4.3 Schematic showing experimental setup for dipping experiments………………..63

4.4 Optical micrograph showing the IMC layer and its trace used for the
measurement of average IMC thickness…………………………………………64

4.5 Micrograph showing thickness reduction in the copper substrate when


dipped in molten solder…………………………………………………………..65

4.6 Micrograph showing copper thickness reduction as a measure of copper lost


during reflow process…………………………………………………………….66

xii
5.1 Schematic and micrographs showing various stages of formation and
growth of IMC during soldering…………………………………........................74

5.2 Schematic showing various regions and interfaces inside a typical solder joint...75

5.3 Physical formulation, geometric representation and concentration


distribution of solute (Cu) in the substrate, IMC and solder layers...........………77

6.1 Comparison of analytical and numerical solutions to show the effect of finite
domain on interface position and interface velocity………………………….….86

6.2 Plots showing effect of interface kinetics on interface concentration as


function of time..............…………………………………………………………89

6.3 Plots showing effect of interface kinetics on interface location as function


of time……………………………………………………………………………90

6.4 Plot showing effect of interface kinetics on interface velocity as function


of time……………………………………………………………………………91

6.5 Plots showing effect of interface kinetics on solute distribution in the solder…..92

6.6 Plots showing effect of solder diffusivity on interface concentration as


function of time…………………………………………………………………..94

6.7 Plots showing effect of solder diffusivity on interface position as function


of time……………………………………………………………………………95

6.8 Plots showing effect of solder diffusivity on interface velocity as function


of time……………………………………………………………………………96

6.9 Plots showing effect of solder diffusivity on solute distribution in the solder…..96

6.10 Plots showing substrate/IMC interface velocity as function of time…………….98

6.11 Plots showing IMC/Solder interface velocity as function of time……………….99

6.12 Plots showing location of substrate/IMC interface as function of time………..100

6.13 Plots showing location of IMC/solder interface as function of time…………...101

6.14 Plots showing difference of velocity of the two moving interfaces as


function of time……………………………………………………………........102

xiii
6.15 Plots showing copper thickness consumption with time at various values
of IMC diffusivity................................................................................................103

6.16 Plots showing growth of IMC thickness with time for various values
of IMC diffusivity................................................................................................104

7.1 Schematic showing transformation of solder cap to an equivalent cylinder.......105

7.2 Plots showing comparison of numerical results with experimental data


for copper substrate thickness consumption at different temperatures................106

7.3 Plots showing comparison of numerically predicted IMC thickness with


experimental data during thermal aging process.................................................109

7.4 Plots showing comparison of numerical prediction with experimental data for
growth of intermetallic phase in the Pd-Sn binary couple...................................110

7.5 Plots showing comparison of numerical prediction with experimental data


for growth of scalloped intermetallic phase.........................................................111

7.6 Plots showing comparison of numerical prediction with experimental data


for copper consumption thickness.......................................................................111

7.7 Schematic showing various diffusion mechanisms in IMC layer during


Cu-molten solder interaction................................................................................113

7.8 Plot showing variation of IMC diffusivity with time used in present
calculations..........................................................................................................114

7.9 Plots showing comparison of numerical prediction with experimental data for
average thickness of scalloped IMC....................................................................115

7.10 Plots showing comparison of numerical prediction with experimental data for
consumed thickness of copper.............................................................................115

7.11 Plots showing comparison of numerical predictions with experimental data for
consumed thickness of copper after adding 1.5 µm to numerically predicted
values...................................................................................................................118

7.12 Comparison of numerical results with experimental data for average IMC
thickness and consumed copper thickness...........................................................119

8.1 Coordinate system and distribution of copper concentration in particle, IMC


and solder domains..............................................................................................123

xiv
8.2 Plots showing effect of inter-particle distance on particle dissolution in
Sn-Ag-0.9Cu solder alloy………………………………………........................125

8.3 Plots showing effect of interface kinetics on particle dissolution in


Sn-Ag-0.9Cu………………………………………………................................128

8.4 Plots showing effect of initial copper concentration in Sn-Ag-Cu alloy


on particle dissolution……………………………………..................................130

8.5 Plots showing effect of initial particle size on IMC thickness during
dissolution of copper particles in Sn-3.5Ag-0.9Cu solder alloy………..............132

8.6 Plots showing effect of initial copper content in SAC solder on IMC
thickness during dissolution of copper particles..................................................133

8.7 Typical thermal reflow cycle considered in the present numerical study............134

8.8 Plots showing effect of heating rate on the dissolution of copper particle
and IMC growth in Sn-Ag-0.9Cu solder alloy.....................................................136

8.9 Plots showing effect of cooling rates on the dissolution of copper particle
and IMC growth in Sn-Ag-0.9Cu solder alloy.....................................................138

8.10 Plots showing effect of holding time interval on the dissolution of copper
particle and IMC growth in Sn-Ag-0.9Cu solder alloy........................................139

8.11 Plots showing effect of initial copper concentration in Sn-Ag-Cu alloy on Cu


particle dissolution and IMC growth during non-isothermal reflow cycle..........141

xv
CHAPTER I

INTRODUCTION

Soldering is one of the oldest techniques of joining two pieces of metal together.

Historically, the method dates back more than two thousand years. In a most simple

soldering process, the gap between the metals to be joined is filled with an alloy, usually

a mixture of two or more pure metals, which has lower melting temperature than the

members to be joined. Heat is applied so that the filler material melts around the

members to be joined and, upon solidification, forms a permanent joint between them.

The filler material, called the solder, plays a vital role in determining the quality

of the resulting joint. In the past, the soldering industry usually employed lead-based

solders for most of the applications. Traditionally, the lead-based solders are tin-lead

alloys of eutectic composition. These solders have been in use in the soldering industry

for several decades and proved to be an excellent soldering material to satisfy most of the

industrial needs. Due to the fact that lead-based solders have a very high content of toxic

lead (37.0 %wt), they are considered harmful for the environment and must, therefore, be

replaced with eco-friendly materials. Due to the environmental and health concerns,

alternate lead-free solders have been intensively researched in the last two decades.

In the current industrial world, soldering has become indispensable for the

interconnection and packaging of virtually all electronic devices and circuits. The fast

changing technology and increasing miniaturization of electronic devices place challenge


1
for obtaining reliable and successful component joints. This is another reason which has

pressed the researchers to investigate and develop the novel solder materials to meet the

requirements and a goal of the modern day solder interconnects.

Almost all the potential lead-free solder materials are high Sn-containing alloys,

including the most promising Sn-Ag-Cu (SAC) alloy of eutectic composition. During the

soldering process, the metallization layer of the printed circuit board (PCB) or the

component pins react with Sn in the solder to form an intermetallic compound (IMC)

layer at the metallization/solder interface. Although the formation of the IMC layer is

desirable for good wetting and necessary bonding, an excessively thick layer is

detrimental to the health and reliability of solder joints. The problem arises due to the

brittle nature of the IMC which makes it prone to mechanical failure even at low loads. A

thicker IMC layer would also increase the heterogeneity in the physical properties of

material across the joint. Therefore, the growth rate of the IMC layer and subsequent

dissolution of the metallization (substrate) must be under control during soldering.

A quantitative assessment of the amount of substrate material lost and the

thickness of the resulting IMC is vital for achieving a better control of the process. The

information can also help in selecting the right solder for a particular application. A

proper mathematical modeling of the process is critical to the understanding and analysis

of various factors affecting the dissolution of substrate and growth of IMC during

soldering process. Researchers have addressed the problem of the IMC thickness

prediction by proposing several empirical and phenomenological models based on the

experimental observations. But these models do not reflect the actual physics involved in

the process. Also, these empirical relations need extensive experimentation and are valid

2
only for the particular material for which they are conducted. Most often these empirical

models are applicable within the parametric premise of the experiments.

1.1 Kinetics of IMC Formation and Substrate Dissolution

In terms of the actual physics involved, the process of soldering is essentially a

more complicated case of mass diffusion process. Any solder in question has a very

limited solubility of the substrate material which acts as a solute in this case. The early

stage of the process when the solder is in the molten state is most critical to the resulting

joint. Researchers have made many assumptions and approximations to deal with the

evolution of the morphologically complicated intermetallic compound and its effect on

the dissolution behavior of the substrate. Many things still need to be addressed and a

more detailed understanding of the process phenomenon is required.

As soon as the molten solder comes in contact with the substrate, an intense

chemical reaction takes place at the substrate/solder interface. The main reactants of the

interface reaction are the substrate material and tin in the solder. The product of the

soldering reaction is the IMC phase formed at the interface. Almost all the commercial

substrate-solder combinations result in the IMC formation at the interface. During wet

phase of the soldering process, the IMC grows in irregular morphology. Figure 1.1a

shows the end view of these undulated humps after the substrate/solder interface is cut

perpendicular to the joint interface, while Fig 1.1b shows the top view of the

morphologically irregular IMC phase when the solder is etched away from the interface.

3
Figure 1.1: Micrographs showing the (a) the section view and (b) top view of the IMC [6]

Once the IMC is formed, it acts as a barrier for further contact of the elements of

the substrate and solder. However, the reacting species continue to diffuse through the

narrow channels formed between the IMC scallops. Initially these channels are deep

enough to provide a direct contact of the solder with the substrate. Tin from the solder,

continues to diffuse through these channels to the substrate/IMC interface. The

availability of tin results in the formation of the IMC at the bottom of these scallops. The

atoms of the substrate elements also continue to travel to the front of the IMC through

these open channels and contribute to the formation of intermetallic phase. The

dissociation of the intermetallic phase into the constituent elements also takes place

simultaneously at the IMC/solder interface and the substrate atoms (solute) diffuse into

the bulk solder. The process of IMC formation and diffusion of substrate element in the

bulk matrix is fast at the beginning but slows down as the solder starts to saturate with

solute. The mechanism continues until the given volume of solder in contact with the

substrate is saturated with solute.

Micrographs in Fig 1.1 and the above explanation reveal that the process of metal

layer dissolution and the IMC growth during soldering is a complicated process. The

4
problem involves two interfaces moving simultaneously. The substrate/IMC interface

will always move towards the substrate side, while the IMC/solder interface will move to

the substrate side as long the dissolution of IMC takes place. Once the dissolution of IMC

stops, the IMC/solder interface will grow into the solder. The scallops of the IMC grow

larger in size but fewer in number with time, resulting in a net decrease in the channels

area for the element diffusion. The scalloped morphology of the IMC is dictated by the

minimization of surface energy and the Gibb’s free energy concept and these scallops can

grow to a certain size as long there is a gain in free energy.

1.2 Objective of the Present Study

The intent of this study is to translate the actual physics involved in the

dissolution of metal layer and growth of IMC during soldering process into mathematical

model based on the fundamental equations and known material properties. In the present

study, the process of soldering will be investigated by developing a one-dimensional

mathematical model based on diffusion and kinetic equations. Physical properties of

materials and thermodynamic constraints will also be incorporated in the mode. The

differential equations will be integrated using numerical techniques. The model will be

used to analyze the relative effect of various kinetic parameters and physical properties of

material on the dissolution behavior of substrate and the evolution of IMC with time

during soldering process. The results from the model will also be compared with

experimental data. Although the IMC grows in irregular morphology during wet phase of

soldering process, and thus the one-dimensional model will have some limitations for

representing the irregular shape of IMC layer, however, the model will accurately

5
translate the basic physics involved in the process of the diffusion of various species

across the joint, phase change and kinetics effects during soldering process. Extensive

parametric study will be conducted to analyze the effects of process parameters and

conditions on the dissolution and growth behavior of IMC and substrate. Variations in

industrial processing conditions for obtaining these joints including the temperature

change with time and the solder volume will also be simulated by setting the appropriate

geometry variables in the model and the effect of these processing conditions will be

analyzed.

6
CHAPTER II

LITERATURE REVIEW

Soldering is one of the oldest techniques for joining two similar or dissimilar

metals. Historically, the practice dates back to several thousand years before Christ. The

basic process of soldering is simple, which involves the melting of a soft filler metal

around the components to be joined and allowing it to cool. A permanent joint is obtained

upon solidification. The soft filler metal, called the solder, is usually an alloy of two or

more pure metals. Traditional solders are essentially an alloy of tin and lead with the

eutectic composition of Sn-63.0 wt% and Pb 37.0 wt%. The lead-based solders served

excellently for almost all the soldering applications and still continue to be a vital

soldering material in many areas. However, the potential health hazards and toxicity of

lead have become more widely understood and appreciated in recent years. The global

movement to eradicate lead from soldering industry has encouraged the researchers to

explore alternate lead-free solders. The electronic packaging industry, being the largest

consumer of solder material has pressed for eradication of lead from traditional solders.

2.1 Global Efforts for Lead-Free Solders

There is a persistent world-wide environmental movement to steer away the

industry from the use of lead and towards ‘non-toxic’ products. Various alternatives have

replaced the traditional use of lead in casting alloys for toys, as well as solders for certain
7
plumbing applications. Once considered a joke, lead-free ammunition (bullets and shot)

is now available and experiencing a significant growth in demand, particularly in the

USA where the possibility of litigation against environmental contamination or employee

exposure to hazardous materials is high. Added to this, there are now a series of

initiatives worldwide that outline targets for electronic equipment re-use and recycling. In

such initiatives, the use of hazardous materials such as lead is often limited in order to

improve the ease of recycling.

Legislation potentially directly affecting the solder and electronic assembly

industries has been passed by the European Commission in the waste from electronic and

electrical equipment (WEEE) and restriction of hazardous substances (RoHS) directives

outlining targets for electronic equipment re-use and recycling. This legislation also

limits the use of hazardous materials to improve the ease of recycling. For Europe, July 1

2006 has been set as the deadline for the transition. It will impact not just on solder alloys

but component finishes and temperature ratings and board finishes issues. Lead in solders

for automotive applications have a temporary exemption from the lead ban.

The Japanese Ministry of Trade (MITI) passed a recycling law for electrical

appliances with effect from April 2001. This suggests but does not include lead phase-out

timetable. OEM's are removing lead from electronics mainly due to market pressures.

Although there is no federal legislation yet in the US, there are a number of State

electronics recycling initiatives to consider. In September 2003 California enacted an e-

waste recycling Bill. In addition, the EPA has proposed a crack-down on lead emissions

from plants that may impact the soldering industry. (Source: www.lead-free.org)

8
Another reason that presses for the search of alternate soldering material is the

increasing demand of miniaturization. An electronic interconnection is very small in size.

A bonding wire is typically of the order of 0.001 inch (25 µm) in diameter [16]. In

general, a very high density of these interconnections is required on a substrate. In case of

a solder interconnection between two copper pads, it is very important to ensure that the

pad is not totally dissolved by the tin in the solder.

The present day aluminum or copper wiring on a very-large-scale-integration

(VLSI) Si chip is 0.5 µm or less. Assuming that the spacing between two wires is also the

same as diameter, the pitch becomes about 1.0 µm. Therefore, on a 1 cm2 area of chip,

one can have approximately 104 wires. To provide the electrical leads to all these wires

on the chip, several thousands of input/output (I/O) pads are needed on the chip. The only

feasible way to provide such a high density of I/O pads is to use area array of tiny solder

balls. Assuming the diameter of a solder ball as 50 µm with a 50 µm spacing in between,

there will be 10, 000 solder balls on a 1 cm2 chip surface. A typical cross-section of a

Plastic Ball Grid Array (PBGA) is presented in Fig 2.1. The available copper is around

25-30 µm while each ball of the solder is around 100 µm.

Figure 2.1: Typical setup for Plastic Ball Grid Array (PBGA)
(Courtesy: National Semiconductors Inc., USA)
9
For these miniature parts, properties like strength, reliability, thermal and

mechanical fatigue of solder joints become critical for the overall integrity of the device.

Traditional lead-base solders are sometimes not able to meet the demands of the modern

day industry. Therefore, there is a quest to improve the lead-free solders also to improve

upon the industry standards.

2.2 Significance of Solder Materials and their Selection Criterion

Solder joints are not merely the electric interconnections between the components

but they also serve as the mechanical support and heat dissipater for the device. Choice of

solder is a very important factor and is critical both to the soldering process and

reliability of the joint during service.

The quality of solder joints depends strongly on the combination of filler and

component materials, including the processing conditions. It is for this reason that a

sound understanding of the metallurgical changes accompanying the sequence of events

that occurs in making solder joint is so vital for the development of a reliable joint.

Theoretical principles have helped to furnish insights, guidelines and qualitative

explanation for the soldering technology, but have rarely provided reliable data for use in

the design of joining processes. In reality the soldering process is extremely complex,

because it brings into play a large number of variables, some of which may not be easy to

recognize and understand. Among the relevant factors are the condition of the solid

surfaces (i.e. nature of other coatings, oxides, sulfides, surface roughness etc.), the

temperature gradient that developed during the joining operation, the metallurgical

10
reaction involving the filler and parent material, and the chemical reactions with fluxes if

they are used [14].

The traditional tin-lead solders have a long history in soldering process. This

solder and the alloys developed with it provide many benefits such as ease of handling,

low melting temperature, good workability, ductility and excellent wetting on copper and

its alloys [9]. Lead bearing solders, and especially the eutectic or near-eutectic Sn-Pb

alloys, have been used extensively in the assembly of modern electronic circuits. [9].

Although, several commercial and experimental lead-free, Sn-based solder alloys

exist, none meets the required standards very well. The essentially desired material

characteristics include such as low melting temperature, wettability, mechanical integrity,

good manufacturability, and affordable cost. Current processing conditions and

equipment (involving fluxes) are optimized for Sn-Pb solder alloys over the past 30

years. The development of proper assembly processes for lead-free solders is also needed

[9].

A number of alloy systems are available to serve as solders for industrial use.

Each alloy is unique with regard to its properties and must be chosen to meet the

requirements of the assembly, such as ductility, resistance to low-cycle fatigue, tensile

and shear strength, and rate of consumption of the base metal in solder and/or substrate

[16].

2.2.1 Environmental Concern

Increasing environmental and health concerns related to the toxicity of traditional

tin-lead solders and the legislations that limit the usage of lead bearing solders, have

11
provided the much desired impetus for intensive research efforts for developing alternate

lead-free solders [6,9]. The first step in finding suitable alloy candidates is, therefore, to

search for some nontoxic, low melting temperature alloys which can replace the lead in

the traditional tin-lead solders [9].

Almost all the potential candidate materials, including the most promising Sn-Ag

alloy of eutectic composition, are high-tin alloys [6]. The candidate alloy components

involve Sn as the base element, Ag, Bi, Cu, and Zn as the major alloying elements, and In

and Sb as some other minor additions. Among these elements In is known to be the

precious metal with limited world production. Given its scarceness, In cannot become a

major alloying element [9].

Besides environmental issues, Pb-based solders may contain a minute amount of


210
Pb isotope. The isotope decays to 206Pb. During the decay, it transforms to Bi and emits

α-particles. When these particles pass though a Si device, they generate electrons and

holes. Before these charge carriers recombine, they may affect the charges stored in the

capacitors (memory units) of the system and lead to what is called a “soft error failure”

[30]. To prevent this problem, refined Pb or Pb-free solders are preferred for critical

applications.

Conventional tin-lead solders are also prone to inferior fatigue properties during

thermal excursions and/or cycling. A harder alloy with high melting point and enhanced

mechanical properties is needed to obtain durable solder joints [4]. The Sn-(3-3.9)%Ag

alloy offers improved mechanical properties and good interfacial bonding [9].

12
2.2.2 Process Conditions of Electronic Assembly

The current processing equipment and conditions for electronic assembly are

optimized for Sn-Pb solders. Any new conditions for lead-free alloys must ensure both

productivity and reliability at least equivalent to present level of Sn-Pb solders. One of

the most sensitive parameter for the quality of solder joints is soldering temperature. The

melting temperature of Sn-Pb eutectic alloy is 183oC, and the typical soldering

temperatures are 230oC and 250oC for reflow and wave soldering respectively. The

temperature margin between the melting point of the solder and processing temperature is

around 50oC for reflow soldering. In contrast, melting temperature for typical lead-free

solders are higher than Sn-Pb eutectic solders by about 30oC, which makes the process

window narrower. Because some of the electronic components cannot withstand an

increase in reflow temperature, processing conditions need to be developed to incorporate

heat-resistant components [9].

2.2.3 Wetting Properties of Solder

Wetting is one of the most important features of solders, because reliable

interconnection requires good wetting. Good wetting results in well formed solder fillets.

The liquid solder is sometimes constrained from spreading away by surrounding the

surface around the confines of the joint with a non-wetting material (a solder mask). The

ability of liquid to wet a solid surface is measured in terms of the contact angle. The

contact angle is the angle that the liquid solder makes with the solid base metal surface.

Figure 2.2 illustrates the relationship between the contact angle, θ and the degree of

wetting. A contact angle of more than 90 degrees indicates lack of wetting affinity. To

13
ensure good wetting, the molten solder should wet the contact surface with a contact

angle of less than 10 degrees [16].

θ
θ

(a) Total non-wetting (θ = 180o) (b) Very good wetting θ → 0o

Figure 2.2: Representation of the degree of wetting in terms of the contact angle, θ

Most lead-free solders seem to exhibit inferior wetting on Cu than Sn-Pb near

eutectic alloy solders. It is well known fact that small amounts of certain impurities in the

solder alloy influence the degree of wetting of the molten solder and the mechanical

properties of an interconnection [16]. While addition of Ag slightly promotes wetting on

Cu, addition of Bi improves wetting significantly. Improved fluxes have also

demonstrated enhanced wetting behavior of solders [9,14].

Wetting behavior of Sn-Pb solder on Ni substrate during a reflow process was

studied as a function of reflow time and the reflow temperature range of 200-240oC by

Kim et al. [26]. The wetting angle for a short reflow time (1 min) was observed to be in

the range from 20o-25o. Regarding the temperature dependency of the wetting angle, the

observations revealed that for the reflow time of less than 10 minutes, the wetting angle

was slightly lower at higher temperatures. However, the wetting angle was not

significantly changed for any temperature when the reflow time was extended beyond 10

minutes. During the same experiment, Kim et al. also studied the wetting behavior of

lead-based solder on Ti substrate using Ni/Ti thin films. It was observed that as the solder

14
came in contact with the Ni layer, the limited thickness of Ni was consumed to form

Ni3Sn4 after 13-15 min. For the extended reflow time when the solder came in contact

with the Ti layer, the dewetting of solder started. Once the dewetting started, the bulk part

of the molten solder began to ball up while a small part of it was left at the periphery

which formed a ring surrounding the ball at the center. The driving force for the

dewetting process was attributed to the minimization of surface energy. The key

parameter for the dewetting process is the high interfacial energy between the molten

solder and the Ti film. A similar phenomenon of dewetting by the Sn-Pb eutectic and

pure Sn solder was also observed on the Cr thin film.

2.2.4 Reliability and Cost

Reliability in electronic soldering involves various factors such as strength,

ductility, thermal and mechanical fatigue, creep, and shock resistance. Since most of the

solders have high melting temperature, it is difficult to understand various behaviors

around room and service temperature. Due to the reasons that diffusion of elements is

quite active and creep can occur even at room temperature, the reliability of the solder

joints is a critical issue [9]. Cost is another important factor in selecting a solder for

practical electronic applications. In general, taking cost of raw metals into consideration,

most lead-free solders cost about two to three times more than Sn-Pb solders. In contrast,

the cost of Sn-0.7wt% Cu eutectic solder is only about 1.3 times higher than Sn-Pb

solder, which explains why Sn-Cu has been successfully transferred to practical

production of consumer products. Table 1 summarizes the candidate lead-free solder

alloys with their respective features [9].

15
Table 2.1: Lead-free solder candidates with typical processing
temperature and features* [9]

Process Temperature
Candidate (oC)
Solder Benefits Drawbacks
Material Wave Reflow
Soldering Soldering

Sn-(3-3.9)% 250-260 235-250 High soldering


Ag Excellent
temperature
Mechanical Lift off** with Bi and
properties Pb
[(0.5-
0.7)%Cu] Soldering Partial melting reaction
temperature can at 139o C with much Bi
or [(1-3)% Bi]
be lowered by Bi
Poor compatibility with
various alloys and Sn-
Pb plating

Sn-57% Bi Very brittle but can be


Low soldering improved by adding Ag
[(0.5-1.0)% 180-200
temperature
Ag] Poor heat resistance

Severe oxidation but


Same soldering can be improved by
Sn-(8-9)% Zn
~250 220-230 temperature as for flux
[3% Bi] eutectic Sn-Pb Poor heat resistance of
interface with Cu

High soldering
Sn-0.7% Cu
250-260 Cheaper cost temperature
[Ag, Ni, Au]
Lift-off** with Pb
*Figures in [ ] are the third alloying elements and their typical compositions.
**Lift-off is type of failure when upon cooling of the joint, the fillet is peeled from the
Cu land located on the printed circuit board.

16
2.3 Soldering Techniques in Electronic Packaging Industry

There are numerous soldering practices employed in the electronic packaging

industry. These techniques are specific to particular need and application. Variations in

soldering methods arise due to different schemes in applying heat, flux and solder

material to the joining components. However, all the soldering techniques can broadly be

categorized into two basic methods viz. reflow soldering and wave soldering. Currently,

these methods are predominantly used by the electronic packaging industry.

2.3.1 Reflow Soldering

In reflow soldering, a mixture of solder and flux called the solder paste is applied

to every joint to be made in the assembly. Heat is applied by means of radiation,

conduction or convection in a controlled environment. The whole assembly is heated and

held at temperature above the melting temperature of solder followed by cooling. During

the process the solder melts and fills the gaps around, forming the joint. An added benefit

of this approach of soldering components to the printed circuit boards and cards is that

there is no geometry dependence, nor is there any limitation to types of components on a

board such as surface mount. The operational sequence of a reflow soldering process is:

1. Apply the solder paste (solder and flux) to the joint


2. Apply heat
3. Cool to the room temperature
Heat

Solder Components
paste to be joined

Figure 2.3: Schematic diagram showing the principle of reflow soldering

17
2.3.2 Wave soldering

In wave soldering, the printed circuit board populated with the components to be

joined is passed across the crest of a molten, standing solder wave. Only the bottom of

the board is exposed to the molten solder. The molten solder serves both as the source of

heat to the board and components, as well as the solder source for the joints. The

operational sequence of a wave soldering machine is:

1. Applying the flux

2. Applying heat

3. Applying solder and heat

4. Cooling to the room temperature

Printed Circuit
Board

Components
pins
Components
to be joined

Finished Joints

Standing
Solder Wave
Apply Heat

Apply Heat
and Solder
Apply Flux

Figure 2.4: Schematic diagram showing the sequence of wave soldering

18
2.4 Anatomy of Solder Joint and Intermetallic Compound Layer

The process of soldering is essentially based on a surface reaction between the

component to be soldered (the substrate) and the molten solder. This reaction at the

substrate/solder interface is of fundamental importance to create the joint. If the reaction

does not occur, there will be no bonding between the substrate and the solder and

consequently no joint will be formed. The control of the interface reaction is crucial both

for the process of soldering and the reliability of the resulting solder joint. The reaction

products, also called the intermetallic compounds (IMC), are formed at the interface of

the substrate and the solder. The formation and growth of the intermetallic layer has a

profound effect on the mechanical properties of the joint and its behavior during its

service life [14]. Any non-metallic surface layer, such as an oxide or sulfide on the

substrate or the solder inhibits the soldering reaction at the interface. Flux is added to

remove the oxide layer and prevent it from forming during soldering. It is important to

note that the flux only facilitates in the reaction to take place but it does not take part in

the actual soldering reaction [16].

During a soldering process, the substrate dissolves in the molten solder and reacts

with the solder. As a result a layer of the intermetallic compound is formed (Fig 2.5) at

the interface of the solid substrate and the molten solder. Most of the lead-free solders

contain high amount of Sn. During the wet phase of most soldering processes, the Sn in

the solder predominantly takes part in the reaction to form the intermetallic compound.
Solder

IMC
Copper

Figure 2.5: Micrograph showing a typical IMC layer in a soldered joint


19
Intermetallic phases formed during soldering in electronic packaging are mostly

the stoichiometric binary compounds containing Sn and the substrate material. The most

common intermetallic compounds formed during soldering are from the Cu-Sn, Au-Sn,

Ni-Sn binary systems [16]. Elemental Pb or Bi does not react with the commonly used

base metals like Cu, Au, Ni or Pd. Thus, in Sn-Ag, Sn-Pb or Sn-Bi solders, only Sn will

react with the base material (most often Cu) to form an intermetallic phase [16]. The

occurrence of Cu-Sn phase upon soldering becomes apparent when viewing the Cu-Sn

phase diagram (Figure 2.5), which reveals the presence of two stoichiometric compounds

corresponding to the compositions of Cu3Sn and Cu6Sn5 [16]. The phase near the copper

substrate has been observed to be Cu3Sn and is normally termed as ε-phase, while the one

next to it is observed to be Cu6Sn5 and is termed as η phase [12]. X-ray diffraction was

used to identify these phases by Lee and Duh [13]. Compositions of IMC layer were

further quantitatively measured. The molar ratio of the Cu/Sn was about 6:5, which is

close to stoichiometric composition of Cu6Sn5 in the compound layer on the solder side.

In the intermetallic layer on the copper side, the molar ratio of Cu/Sn was around 3:1,

approaching the stoichiometric composition of Cu3Sn [13]. In the reflow experiments of

Sn-Pb solder on Ni substrate, Kim et al. [26] identified the IMC layer the Ni3Sn4

compound. In the soldering reaction of molten pure Sn and Pd substrate, PdSn4 was

identified to form at the Pd/Sn interface [27].

2.5 Challenges in Processing of Lead-Free Solder Joints

In the on-going trend of miniaturization in electronic packaging industry the

available thickness of metallization is sometimes only a few microns. In order to get a


20
successful and reliable joint the process of soldering must be highly controlled to

conserve the metallization layer and keep the growth of IMC at minimum.

2.5.1 Controlled Growth of Intermetallic Phase

Although formation and presence of the intermetallic compound layer is desirable

for good wetting and bonding, an excessively thick layer is detrimental due to its intrinsic

brittleness which makes it prone to mechanical failures even at low loads [12]. The brittle

nature of an intermetallic phase is in part a result of the ordered crystal structure of the

compound phase [16]. Also, an excessively thick intermetallic compound layer results in a

joint having non-uniform physical and electrical properties, like coefficient of thermal

expansion and elasticity [9]. The solderability of pre-tinned circuit boards may be degraded

by the intermetallic compound penetrating the pre-tinned surface and deteriorating its

wettability in subsequent soldering operations [14,15].

Uncontrolled growth of IMC poses the weakness in the solder joint and causes

micro-strains due to the mismatch of the thermal expansion coefficient between the solder

and IMC [7]. As the micro-strains exceed a critical value, micro-cracks occur at the

interface of solder/IMC and might propagate, which would result in failure of the joint and

hence influence the reliability of solder joints [7]. Further, the film could continue to grow

during service if the joint attains high temperatures due to internal heat generated by the

chip or heat dissipated from external environment [12]. The growth of IMC in the solid

state is thought to depend, to some extent, on the initial film that formed during reflow

process [12]. It is argued that the degradation of the solder joint is related to the thickness

of IMC and to the kinetics of IMC growth during soldering [7]. In order to prevent the

21
excessive growth of thick intermetallic layers, a barrier layer is often deposited on the

copper base metal contacting the molten solder alloy. Nickel is most often used as a barrier

layer between solder and copper. Ni reacts with the Sn to produce Ni-Sn intermetallic

phases, which grow much more slowly than Cu-Sn intermetallic phases at high

temperatures [16].

2.5.2 Need to Conserve Metallization Layer

During a soldering process, the substrate or base material dissolves into molten

solder. Dissolution of base metals (Cu, Ni, Au etc.) in lead-free solders is observed to be

much higher than in lead-based solders. The increased thickness of IMC during the

reaction involving lead-free solders is due their high content of Sn [8]. In addition,

reaction between the high tin-containing solder and copper is intense during soldering,

which favors quick depletion of copper. Rapid depletion of copper is conducive for

dewetting and failure of the joint. As the thickness of copper is limited and the rework

and/or repair of a solder joint requires a layer of unreacted copper, the loss of copper

during soldering process must be kept under control [6]. Therefore, the dissolution of

copper at the solder/copper interface is of particular importance to the integrity of solder

joints and the overall reliability of electronic devices [2].

The dissolution of base metal during soldering reaction depends upon the relative

amounts of solder and substrate material. When there is a small amount of solder

available compared to the substrate, the solder saturates quickly with solvent, after which

dissolution decreases and fast intermetallic growth can occur. On the other hand, if there

is a large amount of solder available, dissolution of substrate is much greater and

22
resulting IMC thickness is low [8]. Complete dewetting of the scallop shape IMC on Ti

and Cr substrate was observed (contact angle 180o) after the base metal was completely

consumed by the solder [31]. Due to dewetting the scallops took the shape of spheroids.

The phenomenon is known as the “spalling” of the IMC, which sometimes results in the

failure of solder joint. It was further reported that the IMC spheroids were lifted up from

the substrate interface due to gravity effect.

2.6 Growth Morphology and Kinetics of IMC Layer

A core issue in soldering process is the formation of intermetallic compounds

between the solder and the substrate. In solid-state aging planar layer growth is typical

for the interfacial intermetallic formation [13,21,33]. However, in liquid-state soldering,

the scalloped growth is dominant, which is accompanied by a ripening process

[1,6,20,21]. While the temperature difference between these two reactions is less than

100oC, the resultant intermetallic interfacial morphologies differ dramatically [20].

However, an interesting exception is reported by Tu et al. [27] regarding the planar

morphology of IMC during the wet phase of soldering reaction. In these experiments,

pure Sn and Sn-Pd solders were reflowed over Pd substrate. A thick planar IMC,

identified as PdSn4, was observed at the Pd/Sn interface. It was reported that this is

probably the fastest growing intermetallic compound in the binary diffusion couple.

However, most of the basic soldering operations are performed at the temperature

higher than the melting point temperature of solders. In most of the base metal-solder

interaction, when the solder is in molten state, the morphology of the IMC has scalloped

shape. The scalloped shape morphology has generally been observed in Cu/Sn, Ni/Sn and

23
Zn/Sn systems [1,2,4]. Once the IMC is formed, the rate of substrate dissolution greatly

decreases, because it is then governed by the rate at which atoms of the base material can

diffuse through the solder intermetallic compound [15]. The IMC formed at the interface

acts as a barrier for further diffusion of solute in the bulk solder.

The kinetics of copper dissolution and subsequent growth of intermetallic

compound is widely studied and addressed in literature. The actual difficulty in dealing

with the subject arises due to the complex nature of the solder-substrate interaction

during soldering process. The soldering reaction, particularly during the wet phase of

soldering process is very intense. The reaction product of such soldering reaction is the

IMC phase, which forms at the substrate/solder interface. The process of intermetallic

compound formation involves a complicated physics.

The sequence of operations and more importantly the flow of the substrate and/or

solder material in the wet phase of soldering has always been a topic of debate. The

complex physics behind the scallop formation is not fully understood yet. The subject of

discussion is the flow of the material. Is it the molten Sn flowing towards substrate and

forming the IMC at the substrate/IMC interface or the substrate material flows towards

the IMC/solder interface and adds to the IMC surface?

2.6.1 During Early Stages of Soldering Process

As soon the molten solder wets the substrate, the soldering reaction starts rapidly.

The product of the reaction is the IMC phase formed at the substrate/solder interface. The

intermetallic phase layer attains a scalloped shape due to minimization of surface energy.

With time scallops grow larger but fewer, indicating that the coarsening process takes

24
place [6,34]. The mechanism of coarsening is not fully understood yet. To describe this

process Kim and Tu [6] suggested a two-flux non-conservative Ostwald ripening model,

which is based on the assumption that the rapid growth of the Cu6Sn5 compounds on the

solder side is actually a result of a ripening process. Under this assumption, the scallops

bigger than a certain size grow while the smaller ones shrink. The scallops increase in

size at the expense of the smaller scallops. No solid proof for the mechanism was

provided.

The theory of Ostwald ripening seems to be valid in the condition when the solder

is saturated with solute. However, the theory does not seem to satisfy the extremely fast

growth and dissolution of IMC scallops prior to the saturation stage. During the early

stage of soldering reaction the formation and dissolution process of IMC is very fast.

Also, the diffusion of solute in the molten solder is quite rapid prior to the solder

saturation. It is true that the scallops are formed in various sizes and follow some kind of

size distribution, yet the smaller scallops will dissociate into the constituent elements and

the substrate material thus detached will diffuse in the unsaturated solder matrix.

However, the ripening process dictates a kinetics which involves detaching of the

elements from small scallops and attaching them to the bigger ones. In the early stage of

a diffusion dominated process this process of slow interface kinetics will not be possible.

Schaefer et al. [11] considered a problem of soldering with a saturated solder and

introduced a model for the growth kinetics of IMC layer, which assumes that the grain-

boundary diffusion is the predominant transport mechanism through the layer.

25
2.6.2 Diffusion of Elements across Solder Joint

As pointed out by Tu et al. [6,34] during a reflow of Sn-Pb solder on substrate,

that Cu diffuses through the channels between the Cu6Sn5 scallops to reach the molten

solder. Due to high diffusivity of Cu in the molten solder, the Cu atoms diffuse rapidly to

the growth front of the IMC scallops, where they can react with Sn in the solder to add to

the IMC phase. Diffusion of substrate and solder elements through a continuous layer of

Cu6Sn5 was studied by Mei et al. [47]. They reported the intrinsic diffusivities of Cu6Sn5

and Cu3Sn compounds in a Cu-Sn diffusion couple. Based on the diffusivity values of

these compounds, it can be inferred that the diffusion of substrate material or solder

through the IMC phase is rather a slow process. Also, the quantitative results of the

substrate dissolution reflect that the substrate thickness reduction is much higher and

bound to follow a faster mechanism. A faster diffusion path must be arising through the

channels formed in between the scallops. These channels are sometime referred as the

short circuit paths to bring the solder directly in contact with the substrate.

2.6.3 Evolution of the Scalloped Morphology of IMC

Correct description of the IMC grain coarsening is not possible without a clear

picture of the mechanism of scallop formation and growth. Hayashi et al. [50] suggested

that the Cu6Sn5 compound has a scalloped morphology probably due to the fact that the

compound dissolves faster along the grain boundary. However, the authors did not

present a solid proof of that and did not attempt to study the dissolution rate

quantitatively. The reason seems to be valid as the grain boundary is nothing but the

26
crystal defects. It has been reported in literature that the diffusion in grain boundary is

much faster than in the bulk of the grain itself [51].

Quantitative assessment of various material fluxes and their direction of flow are

critical to the understanding of the process and setting a mathematical model. The early

stage the process of soldering involves simultaneous growth and dissolution of IMC as

well as a continuous dissolution of the base metal. The solute from the base metal

continues to diffuse in the solder matrix as long the solder is not saturated. Once the

solder is saturated with solute, the base metal still continues to dissolve, but at a very

slow rate. The IMC phase formed at this stage starts to deposit in the open channels,

making it difficult for the molten solder to come in direct contact with the base metal.

However, the IMC still maintains its scalloped morphology at the IMC/solder front.

There are numerous reasons given by scholars to justify the scalloped-shaped

morphology of IMC during wet phase of soldering process. One of the reasons is based

on gain in maximum free energy. All chemical reactions at constant pressure and constant

temperature occur with a negative Gibbs free energy change. The interfacial reactions

forming IMC should also be governed by free energy change. However, in order to have

the largest negative free energy change in a short period of time, it is the rate of free

energy change that becomes more important. Therefore, the high reaction rate during wet

phase of soldering process can be a result of obtaining a large free energy change. The

reaction path that gives the largest rate of free energy change will be followed in the

course, which in turn depends on the morphology and microstructure of the product

phase. In this respect the scallop type morphology has a higher growth rate and therefore,

is favored during the initial stage of the soldering reaction [30]. Since the atomic

27
diffusivity in a molten solder is much higher than solid solder, a high rate of reaction is

achievable when the morphology of the product IMC phase has scalloped shape. The

scalloped shape morphology provides higher kinetics for the reaction to take place. Yet,

in the solid state of solder, when the high atomic diffusivity is no longer available, the

scalloped shape is no longer stable and thus the IMC transforms to layered-type

morphology.

Another explanation for the layered or scalloped morphology of the IMC can be

based on the minimization of surface energy. Scalloped-type grains of IMC have been

found to be stable in contact with the molten solder, while the layer-type grains are stable

in contact with the solid solder. The stability is due to minimization of surface and grain

boundary energy [5]. Since scallops have a large area than a flat surface, the scallop-type

morphology is unfavorable when the IMC is in contact with the solid solder. However, in

the wetting reaction, the rapid gain in compound formation energy may compensate the

surface energy spent in growing the scallops.

2.6.4 Interfacial Energy of IMC Grains

Transition of a planar IMC layer to scalloped shape was studied by Ma et al. [5].

In these experiments, a planar IMC was grown by solid state annealing of Cu/Sn-Pb

solder joint. Upon subsequent reflow cycles in the molten state of the solder, the IMC

partially dissolved in the solder and the planar morphology changed to scalloped shape.

The transformation of the planar to scallop shape of IMC was attributed to thermal

grooving and the minimization of interfacial energy at the IMC/solder interface. The

planar morphology in contact with solid solder and scalloped morphology of IMC in

28
contact with the liquid solder was justified by Ma et al [5] by relating the interfacial

energy and grain boundary energy (Figure 2.6) as:

γ gb = 2γ is cos ϕ is for solid solder (2.1)

and γ gb = 2γ il cos ϕ il for liquid solder (2.2)

Figure 2.6: Schematic representation of the wetting angle, φ and the force balance at the
IMC/solder interface [5]

The difference in the dihedral angle, φ arises due to the difference in interfacial

energy of IMC in contact with solid and liquid solder (γil < γis). According to the available

data for Sn-Pb solder, γil = 90 mJ/m2 and γis = 200 mJ/m2. If γgb = 164 mJ/m2 is

considered same for both solid and liquid phase solders, the deep grooves in case of

liquid solder can be justified on the basis of difference in the interfacial energies.

2.6.5 Heating/Cooling Rate of Assembly

Effect of heating and cooling rates on the IMC morphology around micron size Ni

particles in contact with eutectic Sn-Ag solder was studied by Lee et al. [32]. It was

29
observed that regardless the cooling rate, the morphology of IMC around Ni particles was

scalloped (what they called as “sunflower”) shape. It was deduced that the cooling

segment of the thermal reflow cycle does not significantly affect the morphological

development of the IMC around the Ni particles in the eutectic Sn-Ag solder. The IMC

morphology change was influenced particularly by the heating rate of the samples.

Depending on the heating rate, two distinctly different IMC morphologies developed

around the Ni particles in Sn-Ag solder. Higher heating rates resulted in scalloped

morphology of IMC while the slowest rate caused blocky and faceted type morphology.

It has been suggested by Tu et al. that a non-conservative ripening process

contributes to the scalloped structure formation, resulting in coarsening of the scallops

that decrease in number with time [5,6,20]. In ripening process, due to the Gibbs-

Thomson effect, the smaller intermetallic grains dissolve into the liquid solder, feeding to

further growth of their neighboring larger grains. Therefore, the dissolution of the

intermetallic compound plays a key role in the intermetallic growth [5].

The dissolution kinetics may be the limiting factor for the ripening process if the

dissolution kinetic rate (the rate of decomposition of intermetallic compound into its

component atoms) is smaller than (or comparable with) the solute diffusion rate in the

liquid solder. This may happen during high temperature soldering when the diffusivity of

solute is relatively high [5]. It is important to point out here that Tu’s study addressed the

stage when the molten solder is saturated with the base metal. However, prior to the

saturation of solder, a different kinetics might determine the morphology and growth rate

of IMC. It has also been reported that the dissolution of the intermetallic compound

reduces the average thickness of the IMC layer during reflow [1,22]

30
2.7 Factors Governing the IMC Thickness and Substrate Dissolution during Soldering

The quantity of intermetallic present within the solder joint depends on a numbers

of factors. Of them, the composition of solder material, maximum temperature reached

during soldering and the length of time during which the substrate was exposed to the

molten solder (dwell time) are most important. Based on these factors, the quantity of

intermetallic within newly-formed joints can vary quite substantially.

The amount of the substrate material dissolved essentially depends on the volume

of solder, the contact area of substrate and solder, the operating temperature and the

solubility limit of the material in solder at that temperature. In most of the copper

substrates in contact with the tin-based solders, the copper dissolution is proportional to

the amount of tin present in the solder.

2.7.1 Composition of Base Material

The most influential factor determining the end IMC thickness is the nature of

base material. All of the common base materials form tin intermetallic compounds. The

most reactive base metal is probably palladium, while gold comes next. In a Pd-Sn binary

couple, PdSn4 phase was reported to be the fastest growing intermetallic compound [27].

Gold dissolution in tin is very high. It has been reported that at 250oC tin can

accommodate approximately 15%wt of gold. Silver also dissolves in tin-based solders

readily but quite less than gold (around 6%wt in pure tin). However, the solubility of base

metals in solid state is very low.

Copper is the most widely used base material. The rate of dissolution of copper in

tin-based solders under normal soldering condition is appreciable, but not to the extent

31
that special precaution need to be taken as for gold and silver. The solubility of copper in

pure tin at 250oC is around 1.5wt%. The rate of dissolution of nickel and palladium are

slower than that of copper. Of the common base materials the slowest to be attacked by

tin are those based on iron alloys such as the well known “Alloy 42” (42Ni-58Fe) which

is widely used for IC leads.

2.7.2 Process Temperature/Time

The temperature of the system strongly influences the solubility of the base

materials in the molten alloy. For example, for copper and silver the solubility at 250oC

in pure tin is approximately twice that at their eutectic temperatures (227oC and 221oC

respectively). It has been found experimentally, that the IMC thickness is profoundly

affected by the soldering temperature and dwell time [14,28,29] and increases by

increasing the two parameters. Another important variable is dwell time. The maximum

soldering temperature reached and the dwell time vary widely between different

soldering methods. There is a remarkable difference in dwell time for the wave and

reflow soldering process. In wave soldering, a contact time of approximately two seconds

is typical while in reflow process, this same may be an order magnitude higher.

During the thermal cycle of the soldering process, another important factor

governing the final thickness and quality of the intermetallic compound is the rate at

which the solder joint is cooled. It has been observed that faster the rate of solidification,

lesser the growth of the intermetallic layer and finer the grain structure of solder in the

joint.

32
2.7.3 Volume of Solder

Volume of solder is important, because as the solder starts to approach saturation

the dissolution rate decreases. From this stage on the dissolution of the base material

contributes solely for the build-up of the intermetallic compound at the joint interface.

The effect of solder volume on the amount of copper dissolution and the thickness

of IMC was studied by Sharif et al. [46]. The experiments were conducted by reflow of

solder balls of two different volumes made of eutectic Sn-Pb solder on the copper

substrate. The contact area of solder to copper was kept same in all the experiments. The

experiments were conducted at three different temperatures and for various time periods.

Reduction in original copper thickness was calculated as an estimate for the amount of

copper dissolution. It was noted that for the solder balls with bigger volume, resulted in

greater reduction in copper thickness. The phenomenon was observed at all temperatures.

The rate of copper dissolution was also higher for the bigger solder volume. The reason is

simple, as the solder volume or the operating temperature increases, more solute can be

dissolved in the solder before saturation. The only source of solute in this case is the

copper substrate. Consequently, an increase in the substrate thickness was noted.

However, the thickness of the IMC showed an opposite behavior. Greater

thickness was observed for smaller solder volumes at all temperatures. The reason for

greater thickness in lesser volume was attributed to the early saturation of the solder with

solute. As the copper continues to dissolve through the open channels even after

saturation, it forms the intermetallic compound and contributes to the increase in IMC

thickness. A similar phenomenon was observed in another study [27,28,47] but the

reason was the solder materials of different compositions.

33
2.7.4 Composition of the Solder Alloy

Perhaps the most obvious factor determining the final quality and thickness of the

intermetallic compound during soldering is the composition of the solder alloy itself.

Many alloys are intentionally pre-doped with the element which forms tin intermetallic,

such as copper and silver. The alloys with pre-doped base material will result in lower

base material dissolution and slightly higher amount of intermetallic compound in the

joint. In their reflow experiments, Alam et al. [47] observed a greater thickness of the

IMC when eutectic Sn-Ag solder was doped with 0.5 wt% copper. The same solder but

without and copper added yield in smaller IMC thickness. The increase in IMC thickness

was justified as due to early saturation of copper-doped solder as opposed to the ordinary

solder.

2.7.5 Morphology of Intermetallic Phase

The intermetallic at the substrate-solder interface generally starts to form as a

planar layer, but in most of the wet phase of soldering processes, it takes the shape of

hemispherical scallops. A planar IMC acts as a protective barrier for further diffusion of

base material into the solder and hence results in reduced thickness of IMC. This

protective layer becomes less effective if it grows in a non-planar morphology, resulting

in an increased dissolution of base material in the solder and consequently a greater

thickness of IMC at the substrate/solder interface.

34
Figure 2.7: Cu-Sn Phase diagram [16]

35
2.8 Recent Progress in Understanding of IMC Growth Kinetics

The exact kinetics of the scalloped morphology during very early stage of the

soldering process is not reported in the literature. The transformation of IMC phase into

scalloped specially takes when the substrate comes in contact with wet solder and

involves a complicated physics. In an attempt to understand the physical mechanism that

controls the formation of the IMC Lord et al. [52] conducted some experiments recently.

In these experiments pure copper was dipped in molten pure Sn solder for very short

periods of time. The morphology and growth rate of the IMC scallops were studied. The

size of the first nuclei was observed to be 0.4 to 0.8 microns in diameter depending on the

solder bath temperature.

Based on the experimental observations, Lord et al. [52] also suggested the

kinetics of the IMC scallop formation during wet phase of soldering process. The

formation of IMC occurs at the IMC/copper interface, through solid-state transformation.

The IMC grains are separated by the grain boundaries. The IMC/solder interface will be

dissolving most of the just created IMC phase replacing the crystallites with melt and the

grain boundaries with channels. The grain boundaries will be wetted by the solder and

dissolved faster than the grains themselves, because grain boundaries are defects of

crystalline structure. The fast dissolution along grain boundary leads to creating scalloped

morphology of the layer. In this regard the authors agreed with the conclusion made by

Hayashi et al. [50] that not the equilibrium process of grain grooving but the fast

dissolution of the grain boundaries is the primary cause of the formation of undulations of

the IMC layer in the form of scallops. It was concluded that the IMC formation is a

simultaneous process of growth and dissolution.

36
In another study [53], Tu et al. addressed the same topic and shed light on the

kinetics of the scallop formation during the wet phase of the soldering process. The

solder reaction of Cu and molten pure Sn was studied by scanning and transmission

electron microscopy. Size distribution and shape of the scallops were determined

experimentally. In light of the very high resolution TEM images the authors revealed a

detailed morphology of the IMC phases. A definite triple point separating the grain

boundary of the IMC scallops were identified and concluded as independent of the reflow

time. The observed wetting angle (Φ in Fig 2.8) was suggested as an equilibrium feature

at the reaction temperature instead of being formed at quenching. The channels formed in

between the IMC (Cu6Sn5) scallops were identified as the ordinary grain boundary.

Between the scallops and the Cu substrate a polycrystalline Cu3Sn layer (around 300 nm

thick) was observed. It was suggested that the Cu3Sn layer formed at very early stage of

the reaction. However, owing to the high density of the grain boundaries it was not

probably the rate limiting factor for the Cu supply to the reactive solid/liquid interface. A

schematic of the solid copper and liquid solder was suggested as shown in Figure 2.8.

Figure 2.8: Proposed schematic showing inter-phase diffusion of Cu in Sn [53]

It was further suggested that the small scallops dissolved faster and the left over

was seen as a thin disk attached to the copper substrate. The rapid dissolution of the small

scallop leaves behind voids seen as a channel in between the scallops.

37
2.9 Composite Solders

The fast changing technology and increasing miniaturization of electronic devices

place a challenge for obtaining reliable and successful component joints. These reasons

have pressed the researchers to investigate and develop the novel solder materials, such

as composite solders to meet the requirements of the modern day solder interconnects.

Composite approaches have been developed in lead-free solder research in an effort to

improve the service temperature capabilities and thermal stability of the solder joints.

One way to obtain composite solders is by embedding the solder matrix with

micro and sometimes nano-size metal particles. These particles, after solidification, act as

reinforcement in the solder matrix and enhance the thermo-mechanical properties of the

joint during service. The terminal size of these particles is important in order to assess the

joint strength and reliability. Dissolution behavior of these tiny particles becomes more

complicated due to their high curvature and associated high interfacial energy at the

IMC/solder interface.

Most of the studies in literature aim at the coarsening and growth of the IMC

layer during service and only limited data is available for the dissolution behavior of

particles during wet phase of soldering process. This is particularly true for lead-free

solders, which is an important issue to eradicate lead from traditional lead-based solders

[1-2]. With the increasing miniaturization of the electronic equipment and the

applicability of these joints in severe environmental conditions (e.g. under the hood in

automobiles), the solder joint reliability has become a challenging issue in recent years. It

has been reported in literature that presence of these particles in solder increases the

strength of the resulting joints [9]. Selection of the proper particle composition for a

38
particular solder is very critical. In addition to the composition of the metal particles, the

size is another important factor to be considered. If the particle size is very small, it is

possible that the entire particle dissolves into the solder during wet soldering reaction. On

the other hand, excessively large particles can lead to heterogeneity in the joint. Due to

the intense soldering reaction at the beginning of the wet phase of soldering process,

particle material quickly depletes during early stage of the soldering. Therefore, the

dissolution behavior of the metal particles in liquid solders has been a critical issue in

electronic packaging industry. Researchers have tried to assess their dissolution in the

liquid solder by studying the dissolution behavior of a planar substrate in molten lead-

free solder [10]. Due to the fact that the micron-level size and spherical shape of these

particles contributes a very high curvature effect at the particle/solder interface, their

dissolution characteristic can not be approximated to that of a planar substrate. The high

curvature associated with the resulting interfacial energy might change the dissolution

kinetics of these particles.

2.9.1 Recent Developments in Composite Solder Research

Composite approach to lead-bearing solders had been applied primarily to

improve their comprehensive properties. Microstructural analysis and mechanical testing

of such composite solders have been reported in literature. Certain composite solders did

show improved mechanical properties sought by electronic/automobile industries.

Marshall et al. carried out studies in micro-characterization of composite solders in early

1990s [56–59]. The composite solders were prepared by mixing Cu6Sn5 (10, 20, 30

wt.%), Cu3Sn (10,20, 30 wt.%), Cu (7.6 wt.%), Ag (4 wt.%), or Ni (4 wt.%) particles

39
with the eutectic Sn-37Pb solder paste. The microstructure of these bulk composite solder

specimens showed that Cu-Sn, Ag-Sn, and Ni-Sn intermetallics were developed in the

composite solders around Cu, Ag, and Ni particles respectively. Layer of Cu6Sn5

compound was formed around Cu3Sn particles in the Cu3Sn reinforced composite solder,

while no new intermetallic compound was formed in the Cu6Sn5 particle reinforced

composite solder. The microstructural analysis showed good bonding of the particulate

reinforcements to the solder matrix suggesting that the resulting composite solders might

exhibit improved strength. Pinizzotto et al. [60] studied the above composite solder and

reported that intermetallic formation at the solder/copper interface was formed after

samples were aged at 140oC for 0–16 days.

Similar studies were carried out by Wu et al. with aging temperatures of 110oC-

160oC for 0–64 days [62]. Addition of these particles was seen to greatly affect the

intermetallic formation near the Cu substrate. Sn sink theory, i.e., the particles act as Sn

sinks which remove Sn from the solder and decrease the amount of Sn for reaction at the

interface, were proposed for the effects of Cu-containing particles and Ag particles on the

kinetics of intermetallic formation. Dispersion strengthened in-situ composite solders of

Sn-Pb-Ni and Sn-Pb-Cu alloys containing 0.1–1.0 µm dispersoids/reinforcements were

produced by induction melting and inert gas atomization, by Sastry et al. [62]. It was

found that, upon reflow of the solder specimens, the fine spherical dispersoids in rapidly

solidified Sn-Pb-Cu alloys coarsen to > 1 µm platelets, however, the dispersoids in Sn-

Pb-Ni alloys remain spherical and be stable with a size of < 1 µm. The difference in

stability of dispersoids in Cu- and Ni-containing solders was explained on the basis of the

difference in solubilities and diffusivities of Cu and Ni in Sn-Pb matrix. These composite

40
solders showed an increase of 25–180% in yield stress and 20–80% in the modulus

values compared to eutectic Sn-37Pb solder.

Another type of dispersion strengthened composite solder was formulated by

Betrabet et al. by adding 2.2 wt.% of Ni3Sn4 intermetallic particles into the Sn-40Pb

solder matrix [63]. Mechanical alloying, a solid state high-energy milling process

developed for superalloy manufacture, provided the means to process such dispersion

strengthened solders. The presence of Ni3Sn4 dispersoids resulted in a smaller grain size

in the as-cast microstructure and after aging at 100oC for 29 h.

The composite solder mixture was made by thoroughly mixing a pre-weighed

amount of copper particles with a commercial Sn–3.5% Ag solder paste was made by Lin

et al. [64] and effect of the copper powder addition on the microstructure was analyzed

after rapid cooling of the samples. It was concluded that copper was dissolved completely

in the solder when 0.5 wt% copper particles were added. The copper particles were not

completely dissolved in the eutectic solder when their addition was in excess of 0.5 wt.%

and sunflower morphology was found in the eutectic solder with porosity located at and

near the top region of the sample. It was also recommended that addition of copper

powder to the eutectic Sn–3.5% Ag solder should be kept in the range of less than 1.0%

for a better microstructural development.

In a previous study Lin et al. [65] presented the microstructure and hardness of

composite solders obtained by addition of nano-size copper powder to conventional Pb-

Sn solder. Copper powders-reinforced composite solders were prepared by thoroughly

blending nano-sized copper powders (average powder particle size 100 nm) with a

powder of a eutectic solder using water-soluble flux. It was observed that copper powder

41
precipitated as intermetallic compounds that were non-uniformly distributed through the

microstructure. Micro-hardness measurements revealed that 30–40% increase in hardness

of the composite solder was attained over the conventional un-reinforced eutectic

counterpart.

In another study, the influence of single-walled carbon nanotubes (SWCNTs) of

93% purity on the phase formation, microstructural characteristics and mechanical

behavior of conventional solder alloys (63Sn-37Pb and Sn-3.8Ag-0.7Cu) was studied by

Kumar et al. [66]. The composite solder specimens were prepared by mechanical mixing

of SWCNTs with solder powders. The composite solders reinforced with the SWCNTs

also exhibited reduction in melting temperature due to the higher surface free energy and

interfacial instability when compared to unreinforced solders. It was also observed that

the SWCNT addition resulted in finer grain size morphology of the solders and SWCNTs

were embedded deeply into the solder matrix. Tensile properties and micro-hardness of

the composite solders were improved with the addition of SWCNT as compared to the

unreinforced counterpart.

In general, composite solders tend to render improved properties. All the reported

investigations were basically exploratory in nature, and the extent of improvement must

be weighed against environmental and economic factors before widespread adoption can

be realized. However, studies on lead-free Sn-Ag based composite solders have received

attention only recently [55].

42
CHAPTER III

MODELING HISTORY OF IMC GROWTH AND METAL DISSOLUTION DURING

SOLDERING

Dissolution of base metal and the development of intermetallic compound (IMC)

are vital to the quality of a solder joint. Final thickness of the IMC in the terminal joint

during a soldering has been a critical concern in the soldering industry. The assessment of

base metal lost during the process is another important subject being investigated by the

researchers in the last two decades. The quantity of the base metal lost and the thickness

of the resulting IMC at the substrate/solder interface quantitatively on the process

parameters such as temperature cycle together with the type and quantity of solder

relative to the substrate thickness.

3.1 Power Law Relation

The kinetics of copper dissolution and growth of IMC can be quantified based on

their thickness measurement. The power-law relation, relating the average thickness of

the IMC formed to the dwell time and the diffusion coefficient, is widely cited in

literature [1,2,9,12,13,15]. The simple empirical relation is given in equation (3.1).

x(T , t ) = Dt n (3.1)

43
where the symbols have the following meanings

x thickness of the IMC layer formed

t dwell time (time for which the solder remains in contact with substrate at

the process temperature)

T absolute temperature of molten solder

D apparent diffusion coefficient, and

n experimentally determined constant

The diffusion coefficient D is related to the process temperature through the Arrhenius-

type relation as given in equation (3.2)

 −Q 
 
D = Do e  RT 
(3.2)

where:

Do a pre-exponential temperature-independent constant

Q the apparent activation energy of the solute (material from base metal),

R the universal gas constant

The activation energy for diffusion, Q, to a first approximation, is proportional to

the melting point of the metal through which the solute diffuses [15]. The rate-controlling

step for reaction between two metals is the diffusion of atoms between the reacting

phases [15]. In general, the concentration of dissolving metal in the molten filler

increases in an inverse exponential manner with respect to time. That is, the dissolution

rate is initially very fast, but then slows as the concentration of the solute tends toward its

saturation limit (i.e. equilibrium) [15].

44
The object of the analysis is to find the proper values of D and n in equation (3.1)

for the temperature range. Multivariable linear regression analysis of experimental data

average IMC thickness at various times is done to estimate the parameters. Experiments

are conducted (for example reflow), where the molten solder is brought in contact with

the substrate for a known time period at a set temperature. The temperature and time is

recorded and the samples are observed under microscope after proper mounting to

calculate the average thickness of IMC. A phenomenological model is assumed as given

by equation (3.1), the logarithmic of which gives the equation of a straight line.

ln( x ) = ln( D ) + n ln(t ) (3.3)

1.1
Pure Sn
Sn-3.5% Ag
1.0

0.9
y = -1.6641x + 3.9097
0.8

0.7
ln D

0.6

0.5

0.4

0.3
y = -1.7584x + 3.7781
0.2
1.7 1.8 1.8 1.9 1.9 2.0 2.0 2.1
1000/T (1/K)

Figure 3.1: Arrhenius plot to estimate the values of activation energy (Q) and apparent
diffusivity (Do)

45
The data is then fitted to a standard straight line equation and by comparing the

parameters of the line with equation (3.3), values of D and n are estimated for each

temperature. Once the value of diffusivity is known for each temperature, an Arrhenius-

type plot is drawn as shown in Fig 3.1. The factor of 1000 with temperature is introduced

to adjust the scaling. The plot is a linear fit for the logarithmic of equation (3.2) as given

in equation (3.4).

By comparing the parameters of the straight line of the Arrhenius plot with

equation (3.4), the apparent activation energy Q and the constant Do are obtained.

Q1
ln( D) = ln( Do ) −   (3.4)
R T 

These parameters are now valid for the given range of temperature. The apparent

activation energy, Q and the constants Do and n have been estimated for various solders

in contact with different substrates and are cited in literature extensively both for wet

soldering [1,2] and solid phase (thermal aging) [13,35] process.

However, for the solid phase of soldering reaction, since the process is mostly

diffusion controlled, the value of n is assumed as 0.5 and the form of the equation is

usually chosen as given by equation (3.5) [13,15].

x = Dt (3.5)

The value of D is estimated by regression analysis of experimental data for the IMC

thickness at various time periods.


46
An improvement in the above model was introduced by Schaefer et al. [1] in

1996. The study was conducted for the reflow of a lead-based solder alloy (62Sn-36Pb-

2Ag) % by wt. on copper substrate. The analysis was based on the same empirical

relation as given in equation (3.1). They considered the reflow process to be non-

isothermal and divided the temperature-time profile of the wet phase of soldering process

into small isothermal regions. Temperature was averaged for the region and the relevant

activation energy and other constants were determined numerically. Optimization

technique was adopted to estimate the overall values of these parameters, which was

valid for the given temperature range of non-isothermal reflow process.

3.2 Modifications Made in the Power Law

Chada et al. [2] introduced an enhancement in the model of Schaefer [1] in 1999.

The analysis was done for the reflow of eutectic Sn-Ag solder on a Cu substrate. The

improved method took into account the dissolution and precipitation of the intermetallic

phase during the thermal cycle of the reflow process. To take into consideration the effect

of non-isothermal behavior, like the previous model, they also divided the relevant

temperature-time profile of reflow cycle into practically small isothermal regions. Kinetic

parameters were calculated numerically for each isothermal portion of the cycle. A

simultaneous dissolution of the IMC was considered until the time when solder is

saturated with Cu. Net growth of the IMC in each time step was taken as the growth

minus dissolution prior to solder saturation and growth plus precipitation after the solder

was saturated with Cu. The model was primarily based on the power law relationship,

and required the experimentally determined data to calculate parameters for the equation.

47
The amount of copper dissolved during a numerical time interval was calculated

by using the Nernst-Brunner equation (3.6).

 C s − C o  KAt
ln s  =
 C −C  V
(3.6)

In equation (3.6) the symbols have the following meanings

C concentration of solute in the liquid solder

Co initial solute concentration

Cs saturation concentration of solute in the solder

t dwell time (time for which the solder remained in contact with Cu)

A contact area of solder with substrate

V volume of the solder

K a constant

The constant, K in equation (3.6) is an experimentally determined parameter which was

reported to follow an Arrhenius relationship of the following form

 − Qd 
 
K = K oe RT 
(3.7)

where Ko and Qd are experimentally determined Nernst-Brunner constants. Assuming that

the IMC is the only source of solute, the thickness of the IMC dissolved during a time

step was calculated using the above equations (3.6) and (3.7) together with a mass

48
balance equation which involved the weight fraction of Cu in the IMC and the density of

the IMC phase. The atomic concentration of Cu in the solder was measured with Energy

Dispersive Spectroscopy (EDS) analysis to calculate the parameters in equation (3.7).

The contribution of solute precipitation in increasing the thickness of IMC, ∆x, after the

solder was saturated with Cu, was calculated by making use of the diffusivity of solute in

solder Di, the thickness of the boundary diffusion layer of solute at the IMC/solder

interface, δ and the atomic fraction of solute in IMC and solder (Xs). The final relation

used to calculate the IMC growth due to precipitation was given as

DiI  X s − X is−1 
∆x ip =  ∆t
δ  X is − X η 
(3.8)

where Di again is given by an Arrhenius relation of the form of equation (3.7). The

numerical results were compared with experimental data. It was found that the present

model under-predicted the thickness of the IMC for all the specimens. The difference in

the results was attributed to the values of experimental data used to calculate the

parameters in Arrhenius relations.

3.3 Understanding the Apparent Diffusivity

The diffusivity parameter appearing in equation (3.1) is not the real diffusivity of

the solder, rather it has been cited in literature as the apparent diffusivity of solute. In an

attempt to crystallize the subject of diffusivity, Ghosh [22] analyzed the coarsening

kinetics of Ni3Sn4 scallops during the interfacial reaction between liquid solders and Cu,

49
Ni, Pd and Ni-Pd substrates. Eutectic Sn-Ag, Sn-Bi and Sn-Pb solders were reflowed

over the substrates for various times and at different temperatures. Ghosh suggested the

shape and kinetics of the IMC on Ni substrate similar to that of Cu6Sn5. A similar

analysis as done by the previous researchers [1,2] was done and the kinetic parameters

were estimated. Coarsening kinetics of NI3Sn4 scallops was rigorously investigated in this

study. It was demonstrated that in the radial growth of scallops the grain boundary

diffusion might be playing a dominant role. A relation, as given in equation (3.9) was

presented for connecting various diffusivity values involved in the coarsening process.

κδDgb
Dapp = Dbk +
d (3.9)

where Dapp, Dbk and Dgb are apparent, bulk and grain boundary diffusivity respectively.

Grain boundary width, δ and a geometric parameter, κ are also involved in the process.

With the assumption that the thickening of Ni3Sn4 scallops is controlled by the grain

boundary diffusivity only, equation (3.9) was simplified after dropping the first term on

the right hand side of equation.

3.4 Substrate Consumption Rate during Soldering

In the scientific community, H. K. Kim and K. N. Tu are actively involved in the

research of soldering process. They have published various papers focusing on the

experimental results, interface morphology, kinetics and modeling studies of soldering

process. In an attempt to estimate the rate of consumption of copper in a soldering

50
process [6], they analyzed the reflow process of eutectic Sn-Pb solder on Cu substrate.

With the assumption that the amount of copper lost from the substrate is the sum of the

copper present in the Cu-Sn compound and in the liquid solder, they presented a mass

balance equation as follows.

1  nV 
∆h =  ρ L + f Cu ρ cVc 
ρ Cu A  100  (3.10)

where

A total interfacial area between solder and Cu

n wt % of Cu in liquid solder

V total volume of liquid solder

fCu weight fraction of Cu in the Cu-Sn compound

Vc total volume of the Cu-Sn compound

ρCu density of Cu

ρL, density of liquid solder

ρc density of Cu-Sn compound

When the molten solder gets saturated with Cu, the consumption rate of Cu depends only

on the change of volume of Cu-Sn compounds, and the rate equation is given by

dh f Cu ρ c dVc
= (3.11)
dt A ρ Cu dt

Kim and Tu demonstrated that the experimental data for Cu consumption closely

followed equation (3.11). But their analysis is good for the case where the amount of

solder is very small and the results will be valid only after the solder is saturated with Cu.

51
3.5 Kinetics of Scalloped Morphology of IMC

In another study [34], Kim and Tu investigated the growth of IMC during the

reflow reaction of Sn-Pb alloy solder on Cu substrate. The focus of the study was to

assess the mechanism involved behind the growth of scallops of Cu6sn5 compound. They

found, like other researchers, that the scallops of the IMC compound grow larger but

fewer with time. The basic mechanisms involved behind the growth were proposed to be

the ripening reaction among the scallops and the interfacial reaction between the scallops

and the substrate. However, they suggested that the ripening is not a constant volume

process since it is also accompanied with the soldering reaction at the Cu/IMC interface.

The ripening flux was calculated following the Gibbs-Thomson effect, which the Cu

concentration in the molten solder at the surface of the IMC grain as follows

 2γΩ 
 
C r = Co e  rRT 
(3.12 a)

2γΩ
and if << 1 then
rRT

 2γΩ 
C r = C o 1 +  (3.12 b)
 rRT 

where

Co equilibrium concentration of Cu in the solder

52
γ interfacial energy per unit area between Cu6Sn5 compound and the molten
solder

Ω molar volume of Cu6Sn5

R the universal gas constant

T absolute temperature

Due to curvature difference, the concentration gradient of Cu is setup in the molten solder

between grains with different radii. The concentration difference causes the ripening flux

of Cu atoms going from the small grains to the larger grains, which permits the large

grains to grow and the small ones to shrink and disappear. Using the quasi-steady state

approximation, the flux of Cu atoms due the concentration gradient was given as

dC C − Ci
J = −D = −D (3.13)
dx δ

where

D the diffusivity of Cu in the molten solder

δ mean separation distance of the grains

The mean separation distance is difficult to estimate directly. Therefore, it was related to

the mean grain radius with a constant, L through equation (3.14). The parameter, L can

have a value from 0 to 1. It was assumed that all scalloped grains tend to achieve a mean

grain radius, r . Combining equations (12) and (13) and considering that

3
ri max = r and δ = Lr (3.14)
2

53
they presented the final ripening flux as

2γΩDCo 1
JR = (3.15)
3LRT r 2

Another flux, contributing to the growth of the IMC phase, called the interfacial reaction

flux was calculated with the assumption that the total Cu lost from substrate is equal to

the total amount of Cu dissolved through the open channels, and was given as

m
ρ A ∆h =
NA ∫ J IT dt
(3.16)

In equation (3.16) the symbols have the following meanings,

ρ the density of Cu

A total area of the solder/Cu interface

∆h consumed thickness of Cu

A∆h consumed volume of Cu

m the atomic mass of Cu

NA Avogadro’s number

JIT total flux of Cu atoms though channels

The total interfacial reaction flux then becomes

NA
J IT = ρ Aν
m (3.17)
54
where ν=dh/dt is the consumption rate of Cu in the reaction. With the assumption that JIT

is uniformly distributed over all the surfaces of hemispherical IMC grains, the atomic

flux of Cu diffusing into a surface element of grain of mean radius r is

J IT ρN A Aν (t ) 1
JI = =
2π (r ) N P 2πmN P (t ) (r ) 2
2
(3.18)

By making use of the Gauss’ theorem, the two fluxes in equation (3.15) and (3.18) were

combined together to give a final relation for the growth of IMC scallops with time as

 γΩ 2 DCo ρAΩν (t ) 
r 3 = ∫  + dt
 3 N A LRT 4πmN P (t )  (3.19)

The first term on the right hand side of equation (3.19) represents ripening and the second

term is due to the interfacial reaction. The above analysis is good only when the solder is

saturated with solute.

3.6 Size Distribution of IMC Scallops

The assumption of a constant mean radius of the IMC scallop as given in equation

(3.20) was relaxed in another study by Gusak and Tu in 2002.

3
ri max = r
2 (3.20)

55
In this analysis [36] of the growth of IMC scallops with time, they considered a bell-

shaped size distribution of the spheroids. The total flux of Cu contributing to the growth

of these scallops was again considered to be the sum of the ripening and interfacial

reaction. A mathematically tedious analytical solution was presented and it was

concluded that the ripening process in the IMC growth is a flux driven ripening (FDR)

process. However, the analysis was limited to the condition when the solder is saturated

with solute. The application of the study was stated to be marginal or nil prior to the

saturation stage.

For the dissolution of the solute in the molten solder, a relation by Dybkov [37] is

widely cited in literature [5]. The equation is given as

dc S
= k * (c s − c )
dt V (3.21)

where the symbols have the following meanings,

cs solute solubility in liquid solder at the reflow temperature

c concentration of solute in the solder at any time, t

k* dissolution rate constant

S surface area of the intermetallic compound in contact with the solder

V volume of the liquid solder

t time

It is important to note that the dissolution rate constant, k* is an experimentally

determined parameter and represents the overall rate of dissolution and does not

56
particularly reflect any reaction kinetics at the dissolving interface. It can easily be

demonstrated that equation (3.21) follows immediately from Fick’s equations on the

assumption of a quasi-stationary distribution of the concentration of solute components

within the diffusion boundary layer. This means that the solute distribution in the

diffusion boundary layer is close to linear. However, this assumption is not always true

especially at the beginning of the dissolution process and one should use the Fick’s law

of diffusion to address the problem in a concise manner.

3.7 Soldering Process as Stefan Problem

The class of Stefan’s problem was extensively studied by the research group of

Vermolen [38-44]. Most of their research work dealt with the dissolution of various

chemical phases into the multi-component metal alloy matrix. Both analytical solutions

[40,42,43] and numerical techniques [38,39,41,44] were used to solve the problems.

Numerical analysis of the dissolution of a spherical particle in binary alloys was done for

a one-dimensional geometry [38]. The diffusion of solute in the bulk of the matrix was

assumed to be governed by the Fick’s second law as follows.

∂c ( r , t )  ∂ 2 c ( r , t ) 2 ∂c ( r , t ) 
= D +  (3.22)
∂t  ∂r
2
r ∂r 

where c(r, t) is the solute concentration in the bulk of alloy matrix at any time, r is the

spatial dimension, t is time and D is the solute diffusivity in bulk alloy.

57
The effect of interface reaction was incorporated with the assumption that the numbers of

solute atoms detached from the dissolving sphere are finite and their rate depends on the

concentration difference of solute between the particle and the matrix. A general

representation of the solute atoms dissolution rate was suggested as follows

n =∞
dN (t )
= 4πR 2 (t ) ∑ K n (t )[c( R (t ), t ) − cα / β ]n (3.23)
dt n =0

Equation (3.23) could be an accurate representation of the interface reaction kinetic, but

the coefficients of the infinite series are unknown, and, therefore, it would complicate the

mathematical approach of the problem. A first order interface reaction at the moving

boundary was considered as follows

 ∂c( r , t ) 
K [c( R(t ), t ) − cα / β ] = D  (3.24)
 ∂r  r = R ( t )

In equations (23 and (24) the symbols have the following meanings

N number of solute atoms separated from the dissolving sphere

R(t) radius of the sphere at any time

Kn interface reaction rate

c(R(t),t) interface concentration at any time

cα/β solubility of solute in the alloy matrix

58
Another boundary condition at the moving interface was obtained by the mass

conservation as given in equation (3.25)

dR (t ) D ρα  ∂c(r , t ) 
=   (3.25)
dt  ρ β β / α ρα  M α  ∂r 
 c − c( R (t ), t ) 
M M 
 β α 

The additional symbols in equation (3.25) have the following meanings

ρα, ρβ average density of alloy matrix and particle respectively

Mα, Mβ average molecular mass of alloy matrix and particle respectively

Parametric study for the moving interface concentration was conducted and their

variation with respect to time and the reducing radius of the sphere was reported.

Dissolution analysis of stoichiometric second phase in ternary alloys was studied

by Vermolen et al. using a numerical approach [39,40]. The analysis was based on the

Fick’s second law of diffusion and equilibrium conditions at the moving interface were

assumed. A numerical approach was presented to solve the problem. The effect of the

diffusivities of various species in the compound on the dissolution behavior was also

studied. In conclusion of the study, it was shown quantitatively that the dissolution

kinetics of a stoichiometric phase in a ternary alloy is determined by physical parameters

like diffusion coefficient of both alloying elements, solid solubilities, particle

composition and stoichiometry. The compositional parameters, like the ratio of the cell

size and the initial particle radius, initial matrix composition of the matrix and particle

geometry (spherical, cylindrical or planar) also play important role in determining the

dissolution behavior of the precipitate.

59
CHAPTER IV

EXPERIMENTAL STUDY TO INVESTIGATE IMC GROWTH AND COPPER

DISSOLUTION

The substrate/solder interaction under varying conditions of solder volume,

composition and process temperature was also investigated experimentally by the author

in another study [54]. In the series of experiments two solders, the pure tin and eutectic

Sn-3.5wt%Ag were selected for investigation. The substrate material used was pure

copper. Two types of experiments were conducted to see the effect of the solder volume

on the kinetics of intermetallic phase formation and substrate dissolution. The reflow and

dipping experiments were conducted to imitate the actual industrial joining process.

4.1 Reflow Experiments

In one series of experiments, 99.9% pure copper substrates were light ground and

micro-polished to remove any layer of oxide and improve wettability. The substrate

samples were then rinsed with water and alcohol. Solder paste of two different

compositions pure Sn and Sn-3.5 wt %Ag was prepared separately by mixing the solder

powder in RMA flux. Two solder tablets each around 100 milligrams of same solder

paste were dispensed over the copper substrate. A fine gauge chromel-alumel

thermocouple was embedded in one of these balls, which acted as a dummy sample. The

dummy sample was used to record the temperature-time profile of the solder copper
60
assembly during soldering. A closed furnace was maintained at the desired experiment

temperature. The furnace temperature was monitored with another dummy sample of

solder copper assembly embedded with a fine gauge chromel-alumel thermocouple.

Copper samples with two solder balls were put in the furnace and the temperature of the

assembly was monitored in real time on computer with the help of a data acquisition

system. The start of the reflow time was calculated from the time when all solder attained

the melting temperature and was in liquid state. The samples were taken out of the

furnace after time periods ranging from 10 seconds to 10 minuets and immediately

quenched in cold water. These experiments were repeated for the two solder

compositions and various furnace temperatures. The four different temperatures selected

for these experiments were the melting points of the solders (221oC for Sn-3.5 wt %Ag

and 232oC for pure Sn), 250oC, 275oC and 300oC.

Real Time
Temp. Recorder
Furnace

Actual
Solder
Sample
Furnace
Temp.
250

Copper
Substrates

Figure 4.1: Schematic showing experimental setup for reflow experiments

61
300
Actual Reflow Time
250 Actual Reflow
Temperature

200
End of Melting
Tem p. (C)

150
Start of Melting

100
Room
50 Temperature

0
0 50 100 150 200 250 300
Time (s)

Figure 4.2: Real time-temperature curve of a typical reflow process

4.2 Dipping Experiments

In other series of experiments, around 400 grams each of Sn-3.5 wt %Ag and pure

Sn solders were taken and melted separately in a stainless steel crucible. The molten

solders were then transferred to glass beakers for the two different experiments. Copper

substrate samples of size 10 mm × 20 mm × 0.4 mm thick were cut from a 99.9 % pure

copper sheet. The sample substrates were mechanically ground and then finish polished

to remove any oxide layer on the surface and to enhance the wettability. The samples

were then rinsed in water and alcohol.

62
Glass containers having molten solders were put in an enclosed furnace

maintained at the desired experiment temperature. The temperature of the solder was

monitored by immersing a fine gauge chromel-alumel thermocouple in the middle of the

solder bath. Solder bath temperature was maintained in a close range of ±2oC of the

experiment temperature. The molten solder was maintained at temperatures of 221oC,

250oC, 275oC and 300oC for Sn-3.5 wt %Ag solder, and 232oC, 250oC, 275oC and 300oC

for pure tin. A thin coating of mildly activated rosin (RMA) flux was applied to the

polished samples. Copper samples were then dipped vertically in the solder bath and

taken out after different time intervals ranging from 5 seconds to 10 minutes.

Thermocouple

Copper Substrate
Temperature
Display
Furnace

250

Molten Solder

Figure 4.3: Schematic showing experimental setup for dipping experiments

63
The samples were cut along the length and perpendicular to the copper/solder

interface using a diamond-cutting wheel. All samples were cleaned with methanol to

remove any dirt and/or grease prior to cold mounting in epoxy at room temperature. The

mounted samples were then wet-ground using 320, 600, 800 and 1200 grit SiC

impregnated emery paper. After grinding, the samples were fine polished using 5 micron,

1 micron and 0.05 micron alumina based lubricant (aluminum oxide suspended in

distilled water). The polished samples were then etched for around 1 to 2 seconds using

an etchant (a solution mixture of 5 ml HNO3, 2 ml HCl and 93 ml Methanol).

4.3 Measurement of Copper Dissolution and IMC Layer Thickness

The polished and etched samples were examined in optical microscope to analyze

the formation and growth of the intermetallic compound. The images at the solder copper

interface were captured with a digital camera attached to the microscope and stored on a

computer. To calculate the average thickness both in case of dipping and reflow process,

the area of IMC, in each micrograph, was manually traced out using a high accuracy

digital planimeter. A thin layer of Cu3Sn was observed in case of reflow experiments at

higher temperatures and longer dwell time. Since the thickness of Cu3Sn layer was very

low (≤ 1 micron), it was included in the total IMC thickness growth.

(a)

(b)

Figure 4.4: Optical micrograph showing (a) the IMC layer and (b) its trace used for the
measurement of average IMC thickness by a digital planimeter
64
To calculate the total amount of copper dissolved in the solder during

copper/solder interaction, it was assumed that a reduction in the thickness of the copper

substrate is a good measure and directly proportional to total copper dissolution. For the

dipping experiments, the reduced thickness of the copper substrate (Fig 4.5) was

measured at three different locations top, middle and bottom of each sample. Their

average was subtracted from the original thickness of the sample to get the reduction in

total thickness. It was noted that in this experiment, the solder tends to attack the

substrate from both sides. Consequently, the measured value of decrease in thickness is

twice the value in an actual soldering process, and therefore, only half of the measured

value is plotted.

Original
Cu

Dipping
Direction

Cu thickness
after dipping

Figure 4.5: Micrograph showing thickness reduction in the copper substrate when dipped
in molten solder

To estimate the dissolution of copper during reflow process, images were taken at

the junction of the solder cap and the copper substrate as shown in Fig 4.6. The images

were taken at the right and left ends of the cap. The dissolution of copper was measured

65
as the depth to which the copper solder interface moved after. Measurements were taken

at both ends. Their average was taken as the thickness reduction during the reflow

process.

Solder

Thickness Copper
Reduction
Figure 4.6: Micrograph showing copper thickness reduction as a measure of copper lost
during reflow process

4.4 General Behavior of the IMC Growth and Copper Dissolution

Growth of the intermetallic compound layer at the copper/solder interface and the

subsequent dissolution of copper substrate were investigated for the two types of

experiments. In general, the average IMC thickness for both solders increased with

increasing the reflow or dipping time and process temperature. A similar trend was

observed for the dissolved thickness of the copper substrate.

The IMC growth rate at all temperatures was high during the initial stage of the

process and then gradually slowed down during the terminal stages for the solders under

question. A similar trend was observed for the dissolution of the substrate material. In

fact, the IMC growth and the substrate dissolution are closely interrelated to each other

and are part of a complicated mechanics of soldering process. The intermetallic

compound layer acts as the diffusion barrier for copper atoms to travel to the IMC/solder

interface. Therefore, the decrease in the substrate consumption rate is actually a result of

a thicker intermetallic layer formation at the copper/solder interface.


66
4.5 Effect of Solder Volume on IMC Growth and Copper Dissolution

In the present study, the two series of experiments differ from each other mainly

in terms of the amount of solder in contact with the copper substrate. In the reflow

experiments there is a limited amount of solder in contact with the substrate, while in the

dipping experiments, the amount of solder around the substrate is practically infinite. The

two experiments were devised to study the important kinetics of the intermetallic

compound formation and the subsequent dissolution behavior of the substrate. The effect

of solder volume and the substrate dissolution was also studied by Ahmed et al. [46] for

the reflow of Sn-Pb solder on copper and by Yu et al. [49] for the dipping of copper

substrate in eutectic Sn-Ag solder.

The average intermetallic compound thickness recorded in the present dipping

experiments was much lower than that for the reflow process. The observation was

consistent for both solders and at all temperatures. The relevant phase diagrams involving

copper and pure tin or Sn-Ag reveal that both the solders have a limited solubility of

substrate material (which acts as solute) at a given temperature. It can easily be inferred

that a larger volume of liquid solder (dipping process) can dissolve more copper as

compared to a smaller solder volume (reflow process). Moreover, the solute from the

substrate is transported in bulk of the solder through diffusion mechanism. Diffusion of

solute was fast at the beginning of the process and slowed down as the solder stated to

saturate with solute. In case of the dipping process, owing to the large amount of solder,

the copper atoms continued to diffuse in the solder even for longer periods of time. A fast

diffusion of copper from the IMC/solder interface would not allow the IMC to grow

beyond a certain thickness. Consequently, a thin intermetallic layer favored for more

67
copper to diffuse to the IMC/solder interface and eventually into the bulk solder. Due to

high volume of the solder, practically large dissolution of solute is possible. Since in the

dipping process, there is no restriction of the amount of solute to diffuse through the

solder, the overall process was limited by the interface kinetics.

On the other hand, in case of the reflow process, where the solder volume was

very limited, the intermetallic compound started to form as soon the molten solder came

in contact with the copper substrate. Copper started to diffuse into the solder, but after a

short period of time, the solder started to saturate with solute. The saturation of the solder

slowed down further diffusion of solute in the bulk of solder. A decrease in solute

diffusion resulted in the build-up of copper concentration at the copper/solder interface.

An increased concentration of copper enhanced the formation of the intermetallic

compound at the interface. As a result, the IMC grew thicker after short period of time.

Eventually, the thicker IMC, acted as a barrier for the diffusion of elements, thereby

reduced the copper dissolution at longer reflow periods. The early stage of process where

a sharp increase in the IMC formation and copper dissolution rate is observed is actually

a diffusion controlled mechanism which is followed by a kinetic controlled process at the

later stage of reflow process.

4.6 Summary of Experimental Study

The experimental study was aimed at finding the effect of solder volume, process

temperature, solder composition on the IMC thickness and dissolution of copper during

copper/lead-free solder interaction. It was observed that growth of intermetallic

compound and dissolved thickness of copper was a strong function of solder temperature

68
and dwell time. Copper dissolution and IMC growth both were favored by increasing

solder temperature. The rate of copper dissolution and IMC growth was observed to be

higher at the initial stage of the process but tend to be constant at the terminal stages. The

solder composition had a mild effect on the terminal IMC thickness and dissolved

thickness of copper where the dissolution of copper was observed more in pure Sn than

Sn-Ag solder. Although the solubility of copper in pure Sn and Sn-Ag solder is almost

same, the difference in IMC thickness for the two solder compositions can be attributed

to the difference in the morphology of the IMC exhibited by the two solders. Both solders

revealed scalloped morphology for IMC, but the scallops were round in pure Sn but more

jagged and needle-like in Sn-Ag. The round scallops in the case of pure Sn created deep

open channels in between them, which allowed the solder to come in direct contact with

the copper substrate and thereby resulted a net increase in copper dissolution.

The experiments were conducted for the reflow and dipping setting depicting the

industrial reflow and wave soldering processes. The volume of solder had significant

effect on the IMC thickness and copper dissolution. Experimental observations revealed

that the end IMC thickness was more in case of reflow while copper dissolution was

much greater in the dipping process. These experimental observations point to the

complicated physics and kinetics involved in the growth of the IMC and dissolution of

the metallization layer during soldering process. This is the intent of the present study to

substantiate the physics with the help of numerical modeling and parametric study using

the model.

69
CHAPTER V

PHYSICAL MODELING AND MATHEMATICAL FORMULATION OF IMC

GROWTH AND METAL DISSOLUTION DURING SOLDERING

In terms of the actual physics involved, the process of soldering can be viewed as

a more complicated case of mass diffusion process. In most cases, tin in the solder reacts

with the substrate to form the intermetallic phase. When molten solder comes in contact

with substrate metallization layer or the leads of electronic components to be joined, the

stage is known as the wet soldering stage. The early stage of the wet soldering is most

critical to the resulting joint quality. During this stage the substrate-solder reaction is very

intense and most of the substrate layer is consumed and IMC is formed. Researchers have

made persistent effort to analyze and evaluate the growth and evolution of the

morphologically complicated intermetallic compound and the dissolution behavior of the

substrate. Most of these studies are based on the experimental observations involving a

careful study of the microstructure of the IMC phase. Still many questions need to be

answered and a more detailed understanding of the process phenomenon is required.

The present study pertains to the post-nucleation stage where the evolution of

IMC phase takes place by a simultaneous process of growth and dissolution. The

nucleation of intermetallic during interfacial reaction with the liquid solder and the

substrate is observed to be extremely rapid and consequently, the entire area of the

70
substrate in contact with the liquid solder will be covered with the intermetallic

compound very quickly.

In case of the substrate/solder interaction, the development of the IMC and the

dissolution of substrate are the integrated processes of the same overall mechanism of

soldering reaction. Progression of each process affects the other. For an in-depth and

detailed analysis, the dissolution of substrate and formation of IMC should be analyzed

simultaneously. Although the experimental observations give a quantitative assessment of

the substrate dissolution and IMC thickness growth, they do not give any insight of the

actual physics involved in the process. A mathematical model which can translate the

actual physics of the process would be a valuable tool to gain an in-depth acquaintance of

the simultaneous process of substrate dissolution and IMC growth. A model based on the

first principal method which can take into account the kinetics of IMC growth and

dissolution can help understand the process more accurately. This is one of the intent of

the present study to develop a mathematical model based on the diffusion equations and

also involving the kinetics and thermodynamics factors. Physical understanding of the

basic physics involved in the process is vital for establishing the mathematical model of

the overall mechanism. In the following paragraphs a detailed analysis of various stages

involved in growth of IMC during soldering process will be presented. It is to be noted

that the wet soldering reaction and the resulting dissolution of substrate and growth of

IMC is a complicated process and researchers are still trying to investigate the process

through experimental observations and theoretical means.

71
5.1 Kinetics of Substrate Dissolution and IMC Growth

Formation and growth of a continuous layer of intermetallic is easy to translate in

terms of physics. The IMC covers the entire interface of the joint. The solder and

substrate atoms diffuse though the IMC and react to form the fresh compound at the two

moving interfaces. The solubility of most of the substrate elements is very low at low

temperatures. The substrate and solder diffusivities are also very low in solid state. Under

these conditions the rate controlling parameter is mainly the intrinsic diffusivity of the

intermetallic phase. Below the melting temperature of the solder, the IMC diffusivity is

also low. Consequently, the process of IMC growth in solid state solder is very slow.

Although, at elevated temperatures, when the solder is in liquid state, the intrinsic

diffusivity of the IMC doesn’t change much, the scalloped morphology of the

intermetallic phase entirely changes the course of growth and dissolution process. The

evolution of the scalloped morphology involves a complicated physics and therefore, the

subject has been a topic of debate in the scientific community. During wet phase of

soldering process, the rate of formation and dissolution of intermetallic phase is governed

both by the interface kinetics and the solute diffusivity in bulk solder and intermetallic

phase. The diffusivity of molten solder is several orders of magnitude higher than the

IMC diffusivity. Therefore, interface kinetics is the rate limiting parameter at the

beginning of the process followed by the solder diffusivity as the reaction proceeds.

On the basis of the analysis of our own experimental results [28,29] and recent

developments towards understanding of the process of IMC formation during early stage

of wet soldering process, the following mechanism of IMC growth and substrate

dissolution is suggested.

72
1. The very first traces of the IMC phase appear in the form of individual nucleated

grains separated by visible valleys. These grains grow rapidly and within a very

short period of time form a scalloped layer where the individual grains are

separated by channels. Initially these channels extend all the way to the substrate.

These open channels should not be confused with the grain boundary as

experimental observations have revealed that their width is much greater than the

grain boundary thickness.

2. Tin from the solder continues to diffuse through these open channels to the

substrate and further diffuses in the inter-phase boundary between IMC and

substrate where growth of IMC takes place.

3. The intermetallic compound (Cu6Sn5 in Cu-Sn reaction) dissociates into the

constituent elements at the IMC/solder interface and the substrate element (Cu)

dissolves into the bulk solder. Consequently a simultaneous process of growth at

the base and dissolution at the front of IMC is established.

4. Due to the soldering reaction at the base of the scallops (substrate/IMC interface)

a fresh layer of substrate is converted into the intermetallic phase. The scallops of

IMC grow bigger and the area of the open channels decreases.

5. The substrate/IMC interface moves towards the substrate side at velocity slightly

faster than the dissolving IMC/solder interface in the same direction. The small

difference in the velocities of the two fronts determines the net growth rate of the

IMC layer.

6. Once the solder is saturated with the solute (metal from the substrate), further

diffusion of solute in the solder is inhibited. However, tin from solder and the

73
substrate element from the metallization layer continue to diffuse through IMC

layer, consequently, the IMC/solder interface, starts to grow in the solder.

(a)
Solder

IMC Islands

Cu Substrate

(b)
Solder
Open
IMC Scallops Channels

Cu Substrate

(c)
Solder

Cu Cu Cu Cu Cu
Sn Sn Sn Sn

Cu Substrate

(d)
Solder

IMC
Cu Substrate

Figure 5.1 Schematic and micrographs showing various stages of formation and
growth of IMC during soldering
74
5.2 IMC Growth as Moving Boundary Problem

The progression of substrate dissolution and growth/dissolution of the IMC layer

during soldering process can also be viewed as a moving boundary problem involving

complicated boundary conditions at the moving interfaces. The problem also involves

phase change and reaction kinetics at the moving interfaces. As soon the solder comes in

contact with the substrate, an intense soldering reaction takes place and intermetallic

compound is formed. The two moving interfaces viz. the substrate/IMC and IMC/solder

are established very quickly. The two interfaces initially move at greater rate but slow

down as the process is advanced. The IMC is formed at the substrate/solder interface and

dissolved at the IMC/solder interface. Consequently, the two interfaces initially move

toward the substrate side.

IMC
Substrate/IMC
Interface Solder
IMC/Solder
Interface

Substrate

Figure 5.2 Schematic showing various regions and interfaces inside a typical
solder joint

The process of growth and dissolution continues until the solder is saturated with

the substrate elements (solute in this case). Once the solder becomes saturated with solute,

the dissolution of the IMC at the IMC/solder interface stops and IMC starts to grow into

the solder. As a result the substrate/IMC interface always moves towards the substrate

75
side while the IMC/solder front initially moves towards the substrate side but starts to

move into the solder once the solders is saturated with solute in the vicinity of IMC.

5.3 Mathematical Formulation of Substrate Dissolution and IMC Growth

From the standpoint of the physical metallurgy, one of the important aspect of the

soldering problem is the growth of intermetallic compounds (Cu6Sn5, Cu3Sn, Ni3Sn4) etc.,

between the solder and the substrate. As the wet soldering process advances, the IMC

grows thicker. The physical problem involves the diffusion of various species across the

joint and into the bulk solder, the kinetics at the moving interfaces and thermodynamics

constraints like solubility in solder at a certain temperature. Different mechanisms could

be dominating at various stages of the process. A mathematical model which can entail

the above mentioned factors will be helpful in understanding the basic physics of the

process.

Although during the wet phase of the soldering process, the IMC is formed in

irregular morphology however, the one-dimensional model will still be able to represent

the physics of the IMC growth and substrate dissolution. The solution of the diffusion

equation in one dimension will help to analyze the relative effect of various physical

properties and important parameters on IMC growth and substrate dissolution during

soldering. The problem involves two moving interfaces viz. the substrate/IMC interface

and IMC/solder interface. The interface concentrations C12, C21, C23 and C32 are assumed

to be fixed throughout the computation and can be obtained from the relevant phase

diagram at the process temperature. The substrate, IMC and solder are represented in Fig

5.3 as domains 1, 2 and 3 respectively.

76
Substrate IMC Solder

Substrate IMC Solder

1 2 3
C12
Substrate C IMC Solder
21 C23
C32

Subs/IMC IMC/Solder
Interface Interface

Substrate IMC Solder


(Cu)
Vi12 Vi23

x
x=0 x=x1 x=x2 x=x3

Figure 5.3: Physical formulation, geometric representation and concentration distribution


of solute (Cu) in the substrate, IMC and solder layers

77
5.3.1 Assumptions made in the Mathematical Model

The assumptions listed as follows are reasonable approximations to address the

problem of substrate dissolution and IMC growth during soldering process.

• The substrate and solder material are homogeneous and isotropic.

• The soldering process is considered an isothermal reaction.

• Diffusion of various elements occurs only due to the concentration gradient of

the specie.

• All other sources such as electrical, magnetic field etc are negligible.

• Due to the above assumptions, Fick’s law of diffusion is valid.

• The diffusion coefficient is constant i.e. independent of concentration,

position and time.

• Diffusion is the dominant mechanism of transport and convection is negligible.

Peclet number was calculated for the system and was found to be much

smaller than one.

• Principles of continuum mechanics hold.

5.3.2 Governing Equations and Boundary Conditions

Under isothermal process conditions, a mathematical model in its simplest form

can be written using the Fick’s second law of diffusion. In the one-dimensional diffusion

equation, (equation 5.1), the lame coefficient a, can be adjusted as 0, 1 or 2 to set the

coordinates as planar, cylindrical or spherical geometry respectively.

∂C j 1 ∂  a ∂C j 
= r Dj  (5.1a)
∂t r a ∂r  ∂r 

78
the lame coefficient, a can have the values 0, 1, or 2 respectively for linear, cylindrical or

spherical coordinate systems.

For one-dimensional planar coordinate system the above equation can be written as

∂C j ∂  ∂C j 
=  D j  (5.1b)
∂t ∂x  ∂x 

where subscript j represents the substrate (Cu), IMC or solder domain.

Suitable boundary conditions were obtained at the two moving interfaces by

writing the proper mass balance equations.

Mass balance equation at the substrate/IMC interface will be

 ∂C 2   ∂C 
Vi12 ( ρ1C12 − ρ 2 C 21 ) = ρ 2  D2  − ρ1  D1 1  (5.10)
 ∂x   ∂x 

and at the IMC/solder interface as

 ∂C 3   ∂C 2 
Vi 23 ( ρ 2 C 23 − ρ 3 C32 ) = ρ 3  D3  − ρ 2  D2  (5.11)
 ∂x   ∂x 
The outermost boundary of the solder and copper domain can be set as the “no flux”

boundary condition. In addition, at the IMC/solder interface, the following kinetic

relationship can also be written [17, 23].

Vi12 = µ (C32 − C eq ) (5.3)

Refer to the coordinate system in Fig 5.1, it is assumed that the boundaries at the

two ends of the domain are isolated with the surroundings, so the following boundary

conditions are valid


79
∂C
=0 at x = 0 and x = x3 (5.4)
∂x

It is also assumed that the solute concentration in bulk solder is zero prior to the onset of

dissolution process. Assuming that the substrate is made of pure metal, the following

initial conditions exist

C(x, t =0) =100 for x = 0 to x1 (5.5a)


C(x, t =0) = C23 for x = x1 to x2 (5.5b)
C(x, t =0) = 0 for x = x2 to x3 (5.5c)

The symbols have the following meanings

j index representing the substrate, IMC or solder domain


(1 = Substrate (Cu), 2 = IMC, 3 = Solder)

t time (s)

x spatial coordinate normal to substrate (m)

Cj solute (Cu) conc. in the substrate, IMC or solder domain (wt%)

Dj diffusivity in substrate, IMC or solder domain (m2/s)

ρj average density of substrate, IMC or solder (kg/m3)

C12 solute (Cu) conc. on the substrate side of substrate/IMC interface (wt%)

C21 solute (Cu) conc. on the IMC side of substrate/IMC interface (wt%)

C23 solute (Cu) conc. on the IMC side of IMC/solder interface (wt%)

C32 solute (Cu) conc. on the solder side of IMC/solder interface (wt%)

Vi12 velocity of the substrate/IMC interface (m/s)

Vi23 velocity of the IMC/solder interface (m/s)

µ interface reaction kinetics coefficient (m/s)

80
The one dimensional, three layers model is a reasonable simulation of the three

domains appearing in the physical problem of substrate dissolution and IMC formation

during soldering. The model is expected to effectively predict the IMC growth and

dissolution and the dissolution of substrate for a given process condition. It has been

widely reported in literature that when the diffusion process proceeds below the melting

point of the solder, a planar IMC is formed at the substrate/solder interface. However, in

most cases involving wet soldering conditions, the morphology of the IMC formed at the

substrate/solder interface is non-planar. Rather, the compound formed at the

substrate/solder interface takes the form of round scallops. It has been consistently

observed in the present study that this scalloped morphology of the IMC phase is a result

of the intricate physics involved in the process.

Although, the one-dimensional model has certain limitations to represent the

irregular morphology of the IMC phase, it will serve a useful tool to understand the basic

physics and kinetics involved in the substrate dissolution and growth/dissolution of the

IMC phase during soldering. In the preset study attempt will be made to utilize the one-

dimensional model to carry out the important parametric study of various rate governing

parameters involved in the basic governing equations. The one dimensional model will

also be utilized to analyze the dissolution behavior of the micron-size particles in solder

matrix. This study is important in the development of composite solders, where the

micron-size embedded particles exhibit a complicated dissolution behavior during

thermal processing of solder joint.

81
5.4 Solution of Mathematical Model

The moving boundary problem with mass balance at the moving interface falls in

the category of the classical Stefan problem. The class of Stefan problems has been of

practical importance for a long time. Analytical solution of the one-dimensional diffusion

equation with moving interface is available in scientific literature. Researchers have tried

both analytical and numerical approaches to solve the moving boundary problem. While

the analytical approach is easy to implement but has limited applications, a numerical

approach needs extensive programming but provides a better representation of the

physical behavior of the problem.

Analytical solution of the one-dimensional diffusion equation has been

successfully used to predict the position of the moving interface. The solution is more

involved in mathematical sense yet can accurately predict the results. The mass

conservation boundary condition at the moving front can easily be implemented in the

solution. The method has been applied to study the dissolution kinetics of metal

precipitates during metallurgical processes. However, the analytical solution suffers

through the assumption of a semi-infinite domain and in most case can be applied to one-

dimensional geometry only. In the present problem of soldering the solder domain is of

finite size, and therefore, the analytical solution will not provide the accurate results.

The problem in implementing the bounded or finite domain can easily be

overcome if the diffusion equation is solved numerically. Since the implementation of the

analytical solution ultimately needs some kind of computer programming assistance to

solve the problem, the use of a numerical technique for the entire solution becomes more

attractive to solve the diffusion equation over required domain. The real challenge in the

82
numerical technique is to track the moving interfaces. In the present study the one-

dimensional problem will be solved numerically by using the coordinate transformation

technique. In this method the physically moving interface is fixed in the computational

domain by suitable transformation of coordinates.

5.5 Numerical Scheme

A simple iteration scheme based on the interface velocity to deal with the

complicated interface conditions was used. The iteration procedure for the non-

equilibrium dissolution problem is described as follows

1. At a given time step, if the kth iteration for the interface velocity is Vi12
k
, the

substrate/solder melt is advanced by ∆εk as follows

∆ε k = Vi12
k
∆t

2. The interface concentration on the solder side is either set as the equilibrium

concentration or calculated through kinetic equation.

3. The solute diffusion equation is then solved for the solder domain.

4. If the interface velocity is correct, the interface mass conservation equation (6.2)

should be fully satisfied. Otherwise, a correction to the interface velocity is

k
needed. If the corrected velocity is Vi12 + δVi12
k
, then the following mass

conservation condition at the interface should be satisfied.

 ∂C 2   ∂C 
k
(Vi12 + δVi12
k
)( ρ1C12 − ρ 2 C 21 ) = ρ 2  D2  − ρ1  D1 1  (6.6)
 ∂x   ∂x 

83
where the concentration gradient is evaluated from the values of concentration

generated during the kth iteration.

5. The interface velocity correction factor, δVi12


k
can then be derived from equation

(6.6) as follows

 ∂C 2   ∂C 
ρ 2  D2  − ρ1  D1 1 
 ∂x   ∂x 
δVi12k = − Vi12
k
(6.7)
( ρ1C12 − ρ 2 C 21 )

6. A new interface velocity can then be determined using this interface velocity

correction as follows

k +1
Vi12 = Vi12
k
+ ωVi12
k
(6.8)

where a relaxation factor ω is applied to control large variations in interface

velocity, Vi12.

7. Using the new interface velocity, steps (1) through (6) are repeated until a

desired convergence is achieved for this time step. The convergence criterion

for the present model is taken as follows

δVi12k
k
<φ (6.9)
Vi12

A typical value of φ = 0.0001 gave satisfactory results for the present study.

8. Upon convergence, the solution is saved and steps (1) through (7) are repeated

for the next time step.

84
CHAPTER VI

MODEL VALIDATION AND PARAMETRIC ANALYSIS

The focus of the present study is to understand the physical process of substrate

dissolution and growth of the intermetallic phase at the substrate/solder interface during

the soldering process. The dissolution of the substrate material and formation and growth

of IMC phase was modeled as mass diffusion problem. In order to understand the

underlying physics of the process experiments were conducted and a mathematical model

based on the basic diffusion equation was developed. The one-dimensional diffusion

equations in planar coordinate system were solved numerically and the relative

contribution and effect of various parameters were analyzed.

6.1 Comparison of Analytical and Numerical Solutions

The problem of substrate dissolution was solved analytically and outcome was

compared with numerical results. Dissolution of a planar copper substrate into liquid tin

solder was considered as the case for making comparison using the following data

Initial copper thickness (x1) = 10 µm

Width of solder domain = 200 µm

Cu diffusivity in solder (D2) = 1.0 x 10-9 m2/s

Characteristic velocity (vo) = 1.0 x 10-6 m/s


85
1.0
Analytical Solution
Numerical Solution
0.9

0.8

xi(t)/x1 0.7

0.6

0.5

0.4
0 10 20 30 40

(a) t(s)

1.0
Analytical Solution
Numerical Solution

0.8

0.6
Vi(t)/Vo

0.4

0.2

0.0
0 10 20 30 40

(b) t(s)

Figure 6.1: Comparison of analytical and numerical solutions to show the effect of finite
domain on (a) interface position and (b) interface velocity

The effect of the underlying assumption of unbounded domain in analytical

solution in the present problem is evident from Fig 6.1. After a short time period of

around 10 sec, the two solutions started to deviate. The deviation in the interface position
86
comes from the fact that, as the dissolution of copper substrate progresses and the solder

volume is limited, the solder started to saturate with the dissolving copper, thereby

slowing down further dissolution of copper. However, since the analytical solution

considers the unbounded solder domain, the solder would never saturate with copper and

the interface position and velocity will not be affected over time due to solder saturation.

The present problem of soldering process involves a finite solder domain and therefore,

the solution to the problem may not be accurate using analytical methods. The finite

solder volume around the substrate in actual problem can be implemented using

numerical techniques.

6.2 Parametric Analysis of Interface Kinetics and Solder Diffusivity

In the present model, one can identify that there are two rate-controlling

parameters, which play critical role in the overall progression of the dissolution of

substrate material in the solder matrix and growth of intermetallic phase. These two

parameters are the interface reaction kinetics coefficient, µ and the diffusivity of substrate

(copper) element in the solder, D2. Parametric study of these physical parameters will be

helpful to understand their contribution and effectiveness in the process of soldering. The

study will also help in understanding the underlying physics of substrate dissolution and

IMC growth. The diffusion equations were solved for a planar geometry to avoid any

influence of curvature effect of cylindrical or spherical geometry on the results.

87
6.2.1 Effect of Interface Reaction Kinetics

To study the effect of the interface reaction (µ) on the overall dissolution behavior

of substrate, numerical experiments were conducted for Cu-Sn system. Due to the

prevailing interface reaction coefficient, the copper concentration at the moving interface

(C21) can not be assumed at equilibrium value rather the same will be affected with the

value of interface kinetics coefficient and interface velocity. Numerical results were

obtained for different values of µ and using constant value of copper diffusivity in tin

solder (D2 = 1.0 x 10-9 m2/s). Copper concentration and location of the moving interface

together with the dissolution rate were calculated and plotted against time. To simulate

and compare the results in a consistent and convenient manner, the following

dimensionless parameters were introduced

Vi (t ) xo C21 (t )
V* = C* = (6.1a)
DL Ceq

x1 (t ) DL t
x* = t* = 2
(6.1b)
xo xo

where V*, C*, x*, and t* are normalized interface velocity, solute concentration, interface

location and time, respectively.

The effect of interface kinetics coefficient (µ) on the dissolution behavior of the

copper substrate in the liquid tin solder was explored (Figs 6.2 - 6.4). The non-

equilibrium interface concentration of copper at the moving interface was observed to

increase rapidly with time (Fig 6.2) after the copper dissolution started and slowed at

later stage. The rate of increase of interface concentration was observed to be a strong

88
function of the interface reaction rate and increased with increasing value of µ. It was

inferred that for an infinitely high value of µ, the interface would attain the equilibrium

concentration very quickly. It was interesting to note that after the onset of dissolution

process, it takes a finite time for the interface concentration to reach a constant value

highlighting the fact that a non-equilibrium conditions exist at the moving inter especially

during early stages of wet soldering process.

Effect of interface reaction kinetics on the dissolution rate of the dissolving

copper substrate was studied (Fig 6.3 and 6.4). These plots also show the variation of the

normalized variables for the equilibrium interface condition for comparison purpose. It

was observed that for any value of µ, the copper dissolution rate was very high at the

beginning and slowed down as the process advanced.

Equilibrium Concentration
1.0

0.8

0.6
C*

0.4 µ increasing

µ = 1 x 10-7 m/s
0.2 µ = 2 x 10-7 m/s
µ = 5 x 10-7 m/s
µ = 10x 10-7 m/s
0.0
0 50 100 150 200

t*

Figure 6.2: Plots showing effect of interface kinetics on interface concentration as


function of time

89
The dissolution rate of copper was slow for low values of µ. The low dissolution

rate of copper was due to the extra time it took for the copper atoms to leave the substrate

due to slow kinetics. As the value of µ was increased, the dissolution rate approached

towards the equilibrium condition very quickly. A higher dissolution rate at the beginning

of the process was attributed to the strong interface reaction between copper substrate and

liquid tin. Due to the intense copper-solder reaction, copper concentration started to

buildup rapidly at the copper/solder interface. Since the copper diffusivity in tin is

limited, the dissolution rate, consequently, had to slow down after short period of time to

maintain the mass balance at the moving interface. These observations revealed that at

the beginning, the copper dissolution rate was controlled by interface kinetics and in the

later stages by diffusion mechanism. These results are consistent with the findings of

other researchers [4].

1.0
µ = 1 x 10-7 m/s
µ = 2 x 10-7 m/s
µ = 5 x 10-7 m/s
0.8 µ increasing µ = 10x 10-7 m/s
Equil. Int. Cond.

0.6
x*

0.4

0.2

0.0
0 200 400 600 800 1000

t*

Figure 6.3: Plots showing effect of interface kinetics on interface location as function of
time

90
0.005
µ = 1 x 10-7 m/s
µ = 2 x 10-7 m/s
µ = 5 x 10-7 m/s
0.004 µ = 10x 10-7 m/s
Equil. Int. Cond.

0.003
V* µ increasing

0.002

0.001

0.000
0 50 100 150 200

t*

Figure 6.4: Plot showing effect of interface kinetics on interface velocity as function of
time

The effect of interface kinetics on the copper (solute) concentration distribution in

the solder matrix was analyzed for the Cu-Sn interaction time interval of t* = 100. These

results are plotted in Fig 6.5. It was observed that the rate and amount of copper in the

solder matrix increased with increasing value of µ. It was interesting to note that a very

high value of interface reaction constant closely resembled the equilibrium interface

concentration conditions. The copper concentration near the interface was observed to be

highest and then decreased towards the outer boundary of solder domain. If the copper

and solder are allowed to remain in contact for substantially longer time, the copper

concentration in the entire solder matrix would reach the equilibrium concentration.

91
1.0 µ = 1 x 10-7 m/s
µ = 2 x 10-7 m/s
µ = 5 x 10-7 m/s
µ = 10x 10-7 m/s
0.8
Equil. Int. Cond.

0.6
C*
µ increasing
0.4

0.2

0.0
0.00 0.05 0.10 0.15 0.20

x1/X

Figure 6.5: Plots showing effect of interface kinetics on solute distribution in the solder

Any increase in the value of µ from 5 x 10-7 to 10 x 10-7 m/s did not result any

significant change in the dissolution behavior of the substrate. The little or no change in

the dissolution behavior when increasing µ to higher values was due to the limitations

imposed by long-range diffusion to balance the process and satisfy the mass balance at

the moving interface. The dissolution process is mutually controlled both by interface

reaction and long-range diffusion of copper in the bulk of solder. The two parameters

must balance their respective contributions to control the overall process. Any excessive

change in the value of one parameter will not cause significant effect in the overall

behavior, because the process will also be limited by the other controlling parameter.

92
6.2.2 Effect of Substrate (copper) Diffusivity in Solder

To study the effect of long-range diffusion of copper in liquid solder on the

dissolution behavior of copper substrate in liquid tin solder, numerical experiments were

conducted for different values of copper diffusivity in liquid tin, D2 and for a constant

moderate value of interface kinetics coefficient, µ = 4.0 x 10-7 m/s. The dissolution

behavior of a planar copper substrate in the liquid solder was explored and plotted in

Figures 6.6 - 6.8. Similar normalized variables as in the study of interface kinetics study

were introduced as follows for the sake of comparison and plotted against time.

Vi (t )Ceq C21 (t )
V* = − C* = (6.2a)
µ Ceq

x1 (t ) µCeq t
x* = t* = (6.2b)
xo xo

The effect of long-range diffusion of copper in solder was studied with respect to the

changing copper concentration at the Cu-Sn interface (Fig 6.6). As the dissolution of

copper started, copper concentration at the Cu-Sn interface also started to increase rapidly

at the early stage and slowed down at longer time period. The rate of increase of interface

concentration was observed to be a strong function of copper diffusivity (D2) in tin solder

and increased with decreasing value of D2.

It was inferred that for a significantly low value of D2 the interface attained the

equilibrium concentration very quickly while for higher values of D2 the interface copper

concentration kept a lower peak value.

93
1.0

0.8

DSn increasing
0.6
C*

0.4
DSn = 0.1x 10-9 m 2/s
DSn = 1 x 10-9 m2/s
DSn = 3 x 10-9 m2/s
0.2
DSn = 5 x 10-9 m2/s
DSn = 7 x 10-9 m2/s

0.0
0 2 4 6 8

t*

Figure 6.6: Plots showing effect of solder diffusivity on interface concentration as


function of time

It is worth mentioning here that increasing the solute (copper) diffusivity in solder

beyond a certain value didn’t significantly change the copper concentration at the

interface. The reason for the minor change in interface concentration at higher values of

diffusivity is due to the fact that at this stage the dissolution is bounded by the finite

interface kinetics rate. As the dissolution of copper started, there was a sharp rise in

copper concentration at the interface, thereafter the rate decreased and the concentration

remained almost constant. The plateau in the concentration curve at higher reflow time

was a result of the balance maintained between the diffusivity of copper atoms and the

interface kinetics. The phenomenon remains active prior to the saturation state of the

solder. After the solder would be saturated with solute (copper), further diffusion of

copper atoms in the solder will be negligible or significantly low.

94
1.0

0.8

0.6
x*
DSn increasing

0.4

DSn = 0.1x 10-9 m2/s


0.2 DSn = 1 x 10-9 m 2/s
DSn = 3 x 10-9 m 2/s
DSn = 5 x 10-9 m 2/s
DSn = 7 x 10-9 m 2/s
0.0
0 2 4 6 8

t*

Figure 6.7: Plots showing effect of solder diffusivity on interface position as function of
time

Effect of copper diffusivity in liquid tin solder on the dissolution rate of the

dissolving copper substrate was studied (Fig 6.7 and 6.8). The dissolution rate was slow

for low values of diffusivity. The low dissolution rate was due to the slow transport rate

of atoms from interface to the interior of solder matrix due to low diffusivity. For any

value of diffusivity, the dissolution rate was very high at the beginning and slowed down

after short period of time, as the process advanced. A higher dissolution rate at the

beginning of the process was due to the absence of any copper concentration at the

interface, as the copper concentration started to build-up rapidly at the interface and the

diffusion length started to increase, the dissolution rate slowed down. Also, since the

solute (copper) diffusivity was finite, the dissolution rate, consequently, had to slow

down after short period of time to maintain the mass balance at the interface. These

95
observations revealed that at the beginning, the dissolution rate was controlled by

interface kinetics and in the later stages of reflow by diffusion mechanism.

2.5
DSn = 0.1x 10-9 m2/s
DSn = 1 x 10-9 m2/s
DSn = 3 x 10-9 m2/s
2.0
DSn = 5 x 10-9 m2/s
DSn = 7 x 10-9 m2/s

1.5
V*

DSn increasing
1.0

0.5

0.0
0 2 4 6 8

t*

Figure 6.8: Plots showing effect of solder diffusivity on interface velocity as function of
time

1.0 DSn = 0.1x 10-9 m2/s


DSn = 1 x 10-9 m2/s
DSn = 3 x 10-9 m2/s
0.8 DSn = 5 x 10-9 m2/s
DSn = 7 x 10-9 m2/s

0.6
DSn increasing
C*

0.4

0.2

0.0
0.0 0.2 0.4 0.6 0.8 1.0

x1/X

Figure 6.9: Plots showing effect of solder diffusivity on solute distribution in the solder

96
The concentration distribution of the dissolving copper element in the solder

matrix for increasing values of diffusivity is plotted (Fig 6.9). It was observed that the

rate and amount of copper in the solder matrix increased with increasing value of D2. It

was interesting to note that a very high value of diffusivity did not make any significant

effect in the solute distribution in the solder matrix due to the limited interface kinetics.

6.2.3 Effect of IMC Diffusivity

One of the important parameters to determine the terminal IMC thickness and the

thickness of substrate dissolved is the intrinsic diffusivity of the intermetallic compound

(D2). The diffusivity of the IMC is a measure of the rate at which the substrate and solder

elements can diffuse through intermetallic phase. Effect of IMC diffusivity on the

dissolution behavior of the copper substrate and growth of the IMC was studied for the

wet phase of Cu-Sn interaction. The interface concentrations at the two interfaces were

assumed to be fixed during the entire calculation and were obtained from the Cu-Sn

phase diagram. A moderate value of interface kinetics (µ = 4 x 10-7 m/s) at the

IMC/solder interface and copper diffusivity in liquid tin solder (D3 = 1 x 10-9 m2/s) was

used in the calculation. These plots show the location and velocity of the two moving

interfaces with time for various values of IMC diffusivity. Similar normalized variables

as in the study of interface kinetics and solder diffusivity study were derived as follows

for the sake of comparison and plotted against time

97
Vi12 (t )Ceq Vi 23 (t )Ceq
*
Vi12 = Vi*23 = (6.3a)
µ µ

xi12 (t ) xi 23 (t ) µCeq t
xi*12 = xi*23 = t* = (6.3b)
x1 x2 x1

where xi12, xi23 are the interface locations and Vi12 and Vi23 are the velocity of the Cu/IMC

and IMC/solder interface respectively. Symbols with (*) are the derived normalized

variables. Other symbols have their usual meanings as defined in chapter 4.

The substrate dissolution rate for various IMC diffusivity values is shown in Fig

6.11. At any value of IMC diffusivity, the substrate dissolution rate was high at the

beginning of the process and then slowed down after short period of time. The negative

velocity indicates the dissolution of the dissolution of the substrate.

0.0

-0.1

-0.2
Vi12*

-0.3

-0.4 Dimc = 1x 10-12 m2/s


Dimc = 1x 10-13 m2/s
Dimc = 1x 10-14 m2/s
-0.5
0 2 4 6 8 10
t*

Figure 6.10: Plots showing substrate/IMC interface velocity as function of time

98
The IMC diffusivity did not have a significant effect of the dissolution rate of

substrate. The dissolution rate was slightly higher for low values of diffusivity. The

higher dissolution rate at low values of IMC diffusivity is attributed to the fact that at low

IMC diffusivity, the IMC thickness was less. Due to thin IMC layer, the copper atoms

quickly diffuse though the IMC and reach the IMC/solder front and eventually diffused in

the solder matrix. The enhanced copper atoms movement consequently, led to the higher

dissolution rate of copper substrate. On the other hand, at high IMC diffusivity the IMC

grew thicker with time, which caused the reduction in the supply of copper atoms to the

IMC/solder interface causing a slower rate of copper consumption.

0.05

0.00

-0.05

-0.10
Vi23*

-0.15

-0.20
Dimc = 1x 10-12 m2/s
-0.25 Dimc = 1x 10-13 m2/s
Dimc = 1x 10-14 m2/s
-0.30
0 2 4 6 8 10
t*

Figure 6.11: Plots showing IMC/Solder interface velocity as function of time

It was evident from Fig 6.11 that the IMC diffusivity (Dimc) had an opposite effect

on the IMC/solder interface movement. The IMC/solder interface also moved rapidly at

the very beginning of the process and slowed down at the later stage. With increasing

99
value of IMC diffusivity the velocity of the interface increased. A negative value of

interface velocity again indicates the movement of the IMC/solder front in the negative

direction (dissolution of IMC), while a positive sign of the interface velocity indicates the

growth of the IMC front in the solder matrix.

It was interesting to note that the substrate/IMC interface velocity (Vi12) was

always negative while the IMC/solder interface velocity (Vi23) was initially negative and

then changed sign at later stage of the process. A negative interface velocity of

substrate/IMC interface indicated that the substrate always dissolved at all stages of the

process. Part of the substrate is consumed in the formation of the IMC and part is

dissolved in the solder bulk. A change in the sign of IMC/solder interface velocity is an

indication of the fact that initially the IMC was dissolving and the front of IMC was

receding but at later stage, the positive interface velocity indicated that the IMC started to

grow at the IMC/solder interface into the solder matrix.

1.0
Dimc = 1x 10-12 m 2/s
Dimc = 1x 10-13 m 2/s
Dimc = 1x 10-14 m 2/s
0.9
xi12*

0.8

0.7

0.6
0 2 4 6 8 10
t*

Figure 6.12: Plots showing location of substrate/IMC interface as function of time

100
The growth of the IMC front in the solder happens when the solder in the vicinity

of the IMC becomes saturated with the dissolving substrate element (copper) and the

IMC can no longer dissolve in the solder. The same phenomenon is also evident from Fig

6.13 which shows the changing location of the IMC/solder interface with time.

The interface velocity changed sign earlier for higher values of IMC diffusivity

indicating that the IMC started to grow earlier for higher value of Dimc. The early onset of

IMC growth into the solder was related to the fact that at higher diffusivity, the transport

of atoms through the IMC was higher but the diffusion of solute copper atoms into the tin

matrix was limited and reduced with time. Consequently, the concentration of copper

started to build-up at the IMC front and the layer started to grow into the solder.

1.00
Dimc = 1x 10-12 m 2/s

0.95 Dimc = 1x 10-13 m 2/s


Dimc = 1x 10-14 m 2/s

0.90

0.85
xi23*

0.80

0.75

0.70

0.65
0 2 4 6 8 10
t*

Figure 6.13: Plots showing location of IMC/solder interface as function of time

The negative slope of the lines showing the changing locations of the two moving

interfaces (Figures 6.12-6.13) indicated the dissolution of the substrate and the IMC.

101
However, it was interesting to note that the slope of the IMC/solder interface location

became positive at later stage of the process, indicating that the IMC started to grow in

the solder matrix.

As the substrate comes in contact with the molten solder, the two moving fronts

are established very quickly. The difference in the velocity of the two interfaces

determines the net growth rate of the IMC at the substrate/solder interface.

0.10
Dimc = 1x 10-12 m2/s
Dimc = 1x 10-13 m2/s
0.08 Dimc = 1x 10-14 m2/s

0.06
abs(Vi12*-Vi23*)

0.04

0.02

0.00

-0.02
0 2 4 6 8 10
t*

Figure 6.14: Plots showing difference of velocity of the two moving interfaces as
function of time

Effect of the change in IMC diffusivity on the net dissolution of the substrate and

terminal thickness of the IMC was also analyzed in the present study (Figures 6.15-6.16).

Plots in Fig 6.14 indicate that initially the difference in the velocity of the two moving

interfaces was high but reduced rapidly with time. However, the difference started to

grow at later stage of the process and eventually became constant.

102
It was evident from numerical results that an increase in the IMC diffusivity

increased the IMC thickness and the consequential increase in the dissolved copper

thickness. The low IMC thickness at the lower diffusivity value was a result of the slow

diffusion of species through IMC. The net growth of the IMC thickness is governed by

the difference in the rates of the IMC formation and dissolution. At low diffusivity, the

IMC dissolved at almost the same rate as it was formed resulting in a very thin IMC

thickness. This was true particularly during the early stage of the process.

12
Dimc = 5x 10-12 m 2/s
Dimc = 1x 10-12 m 2/s
10 Dimc = 1x 10-13 m 2/s
Cu Thickness Consumed (µ m)

0
0 100 200 300 400

t (s)

Figure 6.15: Plots showing copper thickness consumption with time at various values of
IMC diffusivity

103
12
Dimc = 5x 10-12 m 2/s
Dimc = 1x 10-12 m 2/s
10 Dimc = 1x 10-13 m 2/s

IMC Thickness (µ m)
8

0
0 100 200 300 400

t (s)

Figure 6.16: Plots showing growth of IMC thickness with time for various values of IMC
diffusivity

It was interesting to note that a change in the IMC diffusivity value changed

significantly the IMC thickness but the corresponding change in the copper consumption

thickness was not that great. The small change in the copper consumption thickness can

be attributed to the matter that at high IMC diffusivity, although copper and tin elements

can diffuse through IMC at a much greater rate, the diffusivity of copper in tin is limited.

The addition copper would be consumed in the formation of the IMC at the IMC/Solder

interface consequently, the IMC layer grows thicker. When the solder became saturated

with substrate (copper) element, copper consumption rate decreased, then-after, any

consumption of copper elements mainly contributed in the formation of the IMC.

104
CHAPTER VII

NUMERICAL RESULTS FOR IMC THICKNESS AND METAL DISSOLUTION

DURING SOLDERING

7.1 Dissolution of Planar Substrate without IMC Formation

Considering the soldering as an isothermal process, the experimental data for

reduction in copper substrate thickness during the reflow process of pure molten Sn was

compared with the present model. The substrate thickness reduction was a result of

dissolution of the material during reflow process. Part of the material dissolved from

substrate is consumed in the formation of IMC and part is diffused in the bulk solder. In

the present comparison, the formation of the IMC is not considered for the sake of

simplicity. Volume of the solder in contact with the substrate together with the area of

contact is important in determining the reduction in the substrate thickness. Volume of

the solder cap was transformed to a cylinder for the same contact area, as shown in Fig

7.1. It was assumed that one dimensional model with the space axis perpendicular to the

copper/solder interface can be a good approximation to represent the diffusion problem.

The height of the transformed cylinder, h was taken as the width of the solder domain.

solder cylinder
solder cap substrate
h

Original Transformed
Figure 7.1: Schematic showing transformation of solder cap to an equivalent cylinder
105
Results from the model were compared with the available experimental data

[28,29] together with the power law equation. The comparison of thickness reduction of

pure copper substrate in contact with pure tin during a reflow process is shown for four

different temperatures (Fig. 7.2).

10 14
9
12
Thickness Red. (mic

Thickness Red. (mic


8
7 10
6 8
5
4 6

3 Experimental 4 Experimental
2 Power Law Power Law
2
1 Present Model Present Model
0 0
0 60 120 180 240 300 360 420 480 540 600 0 60 120 180 240 300 360 420 480 540 600
Time (s) Time (s)

(a) (b)

18 20
16 18
Thickness Red. (mic

Thickness Red. (mic

14 16
14
12
12
10
10
8
8
6
Experimental
6 Experimental
4 4
Power Law Power Law
2 Present Model 2 Present Model
0 0
0 60 120 180 240 300 360 420 480 540 600 0 60 120 180 240 300 360 420 480 540 600
Time (s) Time (s)

(c) (d)

Figure 7.2: Plots showing comparison of numerical results with experimental data for
copper substrate thickness consumption at (a) 232oC, (b) 250oC, (c) 275oC and (d) 300oC

106
The results from the two-layer (Cu and Sn) model were plotted for the

equilibrium condition at the interface. For this model, the IMC formation was not

considered and the problem was viewed as the diffusion of substrate element in the bulk

solder. However, in actual soldering process, most of the substrate element diffusing in

the solder comes from the dissociation of IMC phase, which might change the interface

kinetics. Since the present model deals with an isothermal case at this stage, the effect of

temperature is to only determine the solubility limit and other interface concentrations of

copper in the solder. The copper (solute) solubility in the solder can be obtained from the

relevant phase diagram. The dissolution results are sensitive to the solute solubility limit

and therefore, these limits should be picked very carefully. For the present comparison

the solubility limits of copper in pure Sn at various temperatures are taken from Cu-Sn

phase diagram as follows.

Table 7.1: Copper solubility in pure Sn solder

Solubility of Cu in pure Sn solder*


Temperature (C)
(at. % Cu) (wt. % Cu)
232 2.41 1.30
250 3.12 1.70
275 4.09 2.23
300 5.04 2.76

* Source: NIST (National Institute of Standards and Technology)

The plots in Figure 7.2 also show the results from the fitting of the data with the

power law equation using the parameters calculated for these experiments. The prediction

from power law was good for the overall reflow process at all temperatures. However, the

107
power law is not based on any physics involved in the process. The actual reflow process

is governed both by the interface kinetics and long-range diffusion of solute. The two

mechanisms control the process and are the rate limiting parameters at various stages of

the process. The model based on the diffusion equation takes into account both these

effects and reflects their respective influence at various stages.

Although, the results from the present simple mathematical model under-

predicted the consumed thickness of copper substrate, however, the present model takes

into account the process physics. The observed difference in the experimental results and

model prediction can be a result of not considering the formation of IMC at the

substrate/solder interface. A significant amount of the substrate material is consumed in

the formation of IMC phase. This would further increase the substrate consumption

during the process. The formation of the IMC would also change the kinetics of copper

dissolution. Once the IMC is formed, the dissolution of substrate is mainly due to the

mass diffusion through the open channels in between the scallops. These scallops slow

down the diffusion and hence the rate of substrate dissolution. At higher reflow time, it

has been observed that a thin layer of the Cu3Sn compound is formed on the substrate

side of the substrate/IMC interface. The Cu3Sn compound being rich in Cu, would

consume more copper than Cu6Sn5 compound consequently further consuming the copper

substrate.

7.2 Numerical Results of IMC Thickness for Planar IMC Growth

The present model was applied to simulate the planar intermetallic growth during

the solid state aging of the copper-solder binary couple. Plots in Fig 7.3 show the IMC

108
prediction through the model at various time periods and for different temperatures. The

growth of the intermetallic phase during solid state aging is a diffusion controlled process.

The present model accurately predicted the IMC growth at various temperatures. It can be

inferred from the binary phase diagram that the solute solubility in solder does not change

significantly as long the binary couple is held below the melting point temperature of the

solder. However, it is the intrinsic diffusivity of the intermetallic compound phase which

increases with temperature and consequently results in an increase in the overall

thickness of the intermetallic phase. The IMC diffusivity values used in the study of solid

state diffusion were taken from reference [47].

12
170 C
125 C
10 Experimental [13]
Intermetallic Thickness (microns)

0
0 2 4 6 8 10 12 14 16 18 20
t, sqrt(hrs)

Figure 7.3: Plots showing comparison of numerically predicted IMC thickness with
experimental data during thermal aging process

109
The present model was also applied to predict the growth of PdSn4 intermetallic

phase for the Pd-Sn binary couple at 250oC [27]. It was reported that PdSn4 is probably

the fastest growing intermetallic phase. The unique feature of the compound is its planar

morphology even during the wet phase of the process. Plots in Fig 7.4 show the evolution

of the IMC phase with time and its comparison with the predicted values from the present

mathematical model. The present model was able to predict the IMC thickness

reasonably well.

60
Present Numerical Model
Experimental Data [27]

50
IMC Thickness (µ m)

40

30

20

10

0
0 10 20 30 40

t (min)

Figure 7.4: Plots showing comparison of numerical prediction with experimental data for
growth of intermetallic phase (PdSn4) in the Pd-Sn binary couple at 250oC

7.3 Model Predictions for IMC Thickness and Copper Dissolution for Scalloped IMC
Growth

The present model was also applied to study the IMC growth during wet phase of

soldering process. When the solder is in molten state, the IMC does not grow as

continuous layer, rather the IMC morphology is observed to be highly irregular.

110
4
Present Numerical Model
Experimental Data [28, 29]

IMC Thickness (µ m) 2

0
0 100 200 300 400

t (s)

Figure 7.5: Plots showing comparison of numerical prediction with experimental data for
growth of scalloped intermetallic phase (Cu6Sn5) at 250oC

10
Present Numerical Model
Experimental Data [28, 29]
Cu Thickness Consumed (µ m)

0
0 100 200 300 400

t (s)

Figure 7.6: Plots showing comparison of numerical prediction with experimental data for
copper consumption thickness at 250oC

111
Plots in Fig 7.5 show the comparison of the results from the present model with

the experimental data of the IMC thickness during reflow soldering. The IMC and solder

diffusivity value used for this numerical experiment were 1 x 10-12 m2/s and 1 x 10-9 m2/s

respectively. The numerical predictions for IMC thickness were low during the early

stage of the process. The difference in the theoretical and experimental data is attributed

to the non-planar development of the IMC. During the wet phase of the copper-solder

interaction, the IMC grows in irregular morphology. The consumed thickness of copper

during reflow experiments at 250oC was also compared with the numerical predictions

from the present model. The consumed thickness of copper substrate was very well

predicted by the numerical model.

7.4 Variable IMC Diffusivity

The high disparity in the experimentally observed and numerically predicted IMC

thickness during reflow soldering process is a result of the highly complicated physics

involved in the growth of the IMC and metal dissolution during Substrate/Solder

interaction. During the wet phase of the soldering process, the IMC layer evolves as a

highly irregular structure. The morphologically irregular IMC layer is a result of a

complex kinetics and surface energy at the IMC/solder interface. The irregular

morphology also affects the dissolution behavior of the substrate metal.

Since the IMC layer is not continuous, the diffusion of substrate and tin elements

across the IMC layer is not straight-forward, rather, they follow a complicated course.

The scalloped structure of the IMC leaves the open channels in between the scallops.

These channels are much wider than the grain boundary and decrease with increasing

112
time as the IMC scallops grow bigger. These channels serve as short circuit paths for the

tin atoms to come in direct contact with the substrate and diffuse in between the IMC and

substrate and form fresh intermetallic compound. During the process, the diffusion rate of

various species is different for each mechanism. The fastest being the direct diffusion of

substrate elements in the bulk of solder while the slowest is the diffusion of copper and

tin elements in the IMC layer. Besides, these mechanisms, the diffusion of substrate and

tin elements in the inter-phase boundary is also very fast while the diffusion of the

elements across the IMC phase is slower than the inter-phase diffusion. As the IMC

scallops grow bigger with time, the area of open-channels decreases and the overall rate

of diffusion across the IMC layer region decreases.

Sn Solder

Cu Cu Cu Cu Cu
Inter-phase
Diffusion
Diffusion Sn Sn Sn Sn
through IMC

Cu

Direct diffusion
in Solder

Figure 7.7: Schematic showing various diffusion mechanisms in IMC layer during Cu-
molten solder interaction

Considering the IMC layer as a region in between the substrate and solder layers

where the diffusing elements cross the region through various mechanisms. The IMC

113
layer has an overall effective diffusivity. The effective diffusivity of the IMC layer is an

engrossment of the contribution of the diffusivities through various mechanisms working

in the IMC layer. The effective diffusivity of IMC layer was very high at the beginning

due to greater area of open channels but decreased with time. The effective diffusivity of

the IMC should not be thought as equivalent to the apparent diffusivity used in the Power

law equation [1]. The apparent diffusivity used in the Power law does not pertain to any

specific region (e.g. IMC or solder), rather it is a measure of the overall diffusivity of

substrate elements across the IMC and in the solder. In light of the parametric analysis

from the present model and the experimental observations, effective diffusivity of IMC

was estimated (equation 7.1) and its change with time is shown in Fig 7.8. The IMC

diffusivity decreased exponentially with time as follows.

Dimc = Do e −αt (7.1)

6
m /s)
2

5
-12
IMC Diffusivity ( x10

0
0 100 200 300 400 500 600

t (s)

Figure 7.8: Plot showing example of variation of IMC diffusivity with time for variable
IMC diffusivity used in this study (Do = 6x10-12 m2/s, α = 0.008)

114
6
Experimental [28, 29]

300oC
5
275oC

IMC Thickness (µ m)
4
250oC

221oC
3

0
0 100 200 300 400

t (s)

Figure 7.9: Plots showing comparison of numerical predictions with experimental data
for average thickness of scalloped IMC using variable IMC diffusivity

18
Experimental [28, 29]
16
Cu Thickness Reduction (µ m)

14
300oC
12

275oC
10

8 250oC

6 221oC

0
0 100 200 300 400

t (s)

Figure 7.10: Plots showing comparison of numerical predictions with experimental data
for consumed thickness of copper using variable IMC diffusivity

115
Numerical predictions from the present model for the average thickness of

scalloped IMC and copper consumption during reflow of Sn-Ag solder on copper

substrate were plotted (Figures 7.9-7.10) and compared with the experimental results

[28,29]. Application of varying IMC diffusivity in the model predicted the results very

well for the reflow temperature range of 221oC to 300oC. The corresponding results of

consumed thickness of copper were under-predicted when compared with the

experimental data from references [28,29]. The under-prediction of the copper

consumption could be a result of the local precipitation and formation of IMC compound

in the bulk of solder. It has been reported that chunks of intermetallic were observed in

the solder bulk [47]. These IMC could be the result of local precipitation in the solder.

The precipitation of intermetallic consumes additional copper and depletes the copper

concentration in the solder. To compensate for the reduced copper concentration,

additional copper comes from the substrate and eventually the net consumed thickness of

copper increases slightly. A similar phenomenon was observed by Alam et al. also during

their experiments of copper dissolution in Sn-Ag and Sn-Ag-Cu solders [47]. The values

of various important parameters used to obtain the plots in Fig 7.9 and 7.10 are listed in

Table 7.2. The equilibrium concentration of copper at various temperatures used in this

study was obtained from reference [68,69].

Table 7.2: Parameters used in the modeling of IMC growth and copper
dissolution during reflow process of Sn-3.5% Ag solder

T (oC) Ceq (wt%) DSn (m2/s) α Do (m2/s)


221 1.1 0.8 x 10-9 0.0055 2.5 x 10-12
250 1.5 1 x 10-9 0.006 4.0 x 10-12
275 2.0 1 x 10-9 0.006 6.2 x 10-12
300 2.5 1 x 10-9 0.006 9.0 x 10-12
116
It was observed that the solder diffusivity (DSn) slightly increased at higher

temperature but did not change at elevated temperature. The index, α remained almost

constant at all temperatures but Do was seen as a strong function of temperature and

increased with increasing temperature. Variation of equilibrium concentration and Do

with temperature was observed to follow a linear relationship and can be related as given

in equations 7.1 and 7.3.

Ceq = 0.0179 T – 2.9003 (7.2)

Do = (0.0823 T – 16.103)x10-12 (7.3)

Another reason of slight under-prediction in copper consumption at higher reflow

time could be due to formation of Cu3Sn compound at the copper/IMC interface. In the

present model the composition of the entire IMC is considered as Cu6Sn5. The Cu3Sn

compound, being rich in copper would consume more copper than Cu6Sn5 compound

resulting in slight increase in copper consumption. Assuming that the equivalent copper

thickness consumed in the formation of IMC in the bulk solder is represented as δ. When

the value of δ ≈ 1.5 µm was added to the consumed thickness of copper, the total copper

thickness consumed is plotted in Fig 7.11. Plots in Fig 7.11 reveal a very good fit with the

experimental data and further that approximately 1.5 µm copper thickness from the

metallization layer was consumed in the formation of Cu6Sn5 compound in the solder

bulk.

117
16
Experimental [28, 29]
300oC
14

Cu Thickness Reduction (µ m)
12 275oC

10
250oC

8
221oC

0
0 100 200 300 400

t (s)

Figure 7.11: Plots showing comparison of numerical predictions with experimental data
for consumed thickness of copper after adding 1.5 µm to numerically predicted values

Numerical results from the present model were also obtained to predict the

average IMC thickness obtained experimentally by Chadha et al. [1] for the Sn-Ag solder.

Plots in Fig 7.12 show the numerical predictions for the average IMC and their

comparison with experimental data [1]. The values of various parameters used in these

numerical experiments were obtained from the established values of these parameters as

given in Table 7.2. These plots reveal that the model was able to predict the average IMC

thickness very well during reflow process. The experimental data for copper consumption

for the study was not provided in the reference [1], however, the corresponding copper

thickness consumption from the present numerical model is also shown in the plots (Fig

7.12 b).

118
7
Experimental [1]

6 292 oC

5
IMC Thickness (µ m) 269oC

4 251oC
226oC

0
0 2 4 6 8

t (min)

14
292oC

12 269oC
µ m)
Cu Thickness Reduction (µ

10 251oC

8 226oC

0
0 2 4 6 8

t (min)

Figure 7.12: Comparison of numerical results with experimental data of Chadha et al. [1]
for (a) average IMC thickness and (b) consumed copper thickness

119
The introduction of the variable diffusivity not only predicted the IMC thickness

and copper thickness consumption well but also substantiated the complicated physics of

the diffusion of substrate and solder elements across the IMC layer. The numerical results

using variable IMC diffusivity also revealed that a single value can not represent the

diffusivity of the IMC especially during early stage of soldering process. It was also

revealed by the application of variable IMC diffusivity that at the beginning or reflow

process grain boundary diffusion and diffusion through open channels is dominated

which losses its influence with followed by the diffusion of elements through IMC layer

at later stage of reflow process.

120
CHAPTER VIII

MODEL APPLICATION TO STUDY DISSOLUTION BEHAVIOR OF MICRO-SIZE

METAL PARTICLES IN COMPOSITE SOLDERS

The ever increasing miniaturization of electronic devices has pressed the need to

find the alternate solders that can deliver the necessary strength and reliability of the

solder joints. In this view the development of composite solders has become the focus for

many researchers in recent years. One way to obtain the composite solders is by

embedding the solder matrix with micro and some times nano-size metal particles. These

particles, after solidification, act as reinforcement in the solder matrix and enhance the

thermo-mechanical properties of the joint during service. The terminal size of these

particles is important in order to assess the joint strength and reliability. Dissolution

behavior of these tiny particles becomes more complicated due to their high curvature

and associated high interfacial energy at the IMC/solder interface. The present

mathematical model was also utilized to simulate the dissolution behavior of metal

particles in composite solders during reflow process. Thermal effects and related

thermodynamic constraints together with the non-equilibrium interface kinetics of

dissolving micro-size metal particles were also considered. The growth of intermetallic

compound (IMC) and terminal size of embedded particles for various reflow conditions

were analyzed using the model. Dissolution behavior of micro-size Cu particles in lead-

121
free, Sn-Ag-Cu (SAC) alloy solder was studied. The end particle size and the thickness of

the IMC layer around the particles are presented for various initial particle sizes and

reflow conditions. Effects of initial copper content in Sn-Ag-Cu (SAC) solder alloy and

the interface reaction kinetics on the dissolution of copper particles and growth of

intermetallic compound are also investigated.

8.1 Mathematical Model for Micro-Particles/Solder Interaction

The model is based on the basic Ficks law of mass diffusion. The one-

dimensional diffusion equation in spherical coordinates can be written as follows

∂C k 1 ∂  2 ∂C k 
= 2  r Dk  (8.11)
∂t r ∂r  ∂r 

where the symbols have the usual meanings as given in chapter 6, the value of k as 1, 2

or 3 represents the value of various physical parameters in the substrate, IMC or solder

domains respectively. The copper (solute) concentration for the present problem in the

particle, IMC and solder is also shown in Fig 1. Referring Figure 1, the mass balance

equation at the particle/IMC interface (I12) and IMC/solder interface (I23) can be written

as equations (8.2) and (8.3) respectively

 ∂C 2   ∂C  (8.2)
Vi12 ( ρ 1C12 − ρ 2 C 21 ) = ρ 2  D2  − ρ1  D1 1 
 ∂r   ∂r 

 ∂C   ∂C 2  (8.3)
Vi 23 ( ρ 2 C 23 − ρ 3C32 ) = ρ 3  D3 3  − ρ 2  D2 
 ∂r   ∂r 

122
In addition, at the IMC/solder interface, the following kinetic relationship was also

obtained [17, 23]

Vi 23 = µ (C32 − Ceq )
(8.4)

Solder

IMC

Particle
0 r
r1(t)
r2(t)
R
I12 I23
C
C12

C21 C23

1 2 C32 3
r

Figure 8.1: Coordinate system and distribution of copper concentration in particle, IMC
and solder domains

The interface concentrations C12, C21, C23, and Ceq were obtained from the Cu-Sn

phase diagram at the operating temperature. The interface concentration values C12, C21

and C23 are fixed and considered to remain constant throughout the calculation.

The present model can be used to analyze the dissolution of metal particles by

inserting the relevant physical properties for the system composition of interest. The

model in the present study was tested by considering the dissolution of micro-size copper
123
particles in Sn-Ag-Cu (SAC) solder alloy. It can be identified that besides various rate-

controlling parameters, there are three important conditions which play a critical role in

the overall progression of the dissolution of particles in the solder. These three conditions

are the inter-particle distance, the initial copper content in the SAC solder and the

interface reaction kinetics. In the present study the reflow process was considered

isothermal at 250oC and the copper diffusivity in IMC and molten SAC solder were taken

as D2 = 10-13 m2/sec [18] and D3 = 10-9 m2/sec respectively.

8.2 Effect of Inter-Particle Distance on Particle Dissolution

The inter-particle distance represents the distance between the centers of any two

neighboring spherical particles and will be equal to twice the solder cell radius, (2R) in

Fig 1. For the present analysis it was assumed that the particles in the composite solder

are embedded in such a way that the concentration fields of dissolving particles do not

overlap. A non-dimensional parameter, r* was introduced for the sake of analysis as

follows

r1 (t )
r* = (8.5)
ro

where r1(t) is the radius of the dissolving particle at any time instant, t and ro is the initial

radius of the particle. Effect of inter-particle distance or the solder cell size was studied

(Fig 8.2). The plots show the dissolution of copper particles in Sn-Ag-0.9Cu solder alloy

124
for different initial copper particle radii. It was evident from the plots that the dissolution

of Cu particles in the molten solder was very rapid.

1.0

0.8

5 µm
r* = r1(t)/ro

0.6
3 4
2
ro=1 µm
0.4

0.2

(a) 0.0
10-3 10-2 10-1 100 101 102 103
t (sec)

1.0

0.8 12 µm

0.6 10 µm
r* = r1(t)/ro

9 µm
ro=1 µm 5 µm

0.4

0.2

(b) 0.0
10-3 10-2 10-1 100 101 102 103
t (sec)

Figure 8.2: Plots showing effect of inter-particle distance on particle dissolution in Sn-
Ag-0.9Cu solder alloy for (a) 2R = 50 µm, (b) 2R = 100 µm

125
Numerical results revealed that smaller particles dissolved in the molten solder

completely in a short time period however, the bigger particles first dissolved very

quickly but slowed down after short period of time. Rapid dissolution of the particles can

be attributed to their small size. The complete dissolution of the smaller particles

indicated that the solder cell around the dissolving particle was not saturated with copper

during the entire dissolution course.

A decrease in the dissolution kinetics of bigger particles at extended dwell time

can be a result of the saturation of the solder cell with copper. Once the solder around the

particle was saturated with dissolving copper, further diffusion was stopped which

slowed the dissolution kinetics. Another important but expected result was revealed that

the bigger solder cell size or greater inter-particle distance was able to dissolve larger

copper particles completely. The result can be justified again to the fact that the greater

volume of solder was able to dissolve a larger amount of copper in it before the saturation

stage.

8.3 Effect of Interface Kinetics on Particle Dissolution

During the wet phase of soldering the overall dissolution of the particle in the

molten solder is governed by the interface kinetics and bulk diffusivity of the solder. This

is particularly critical when the solute (copper) diffusivity in the solder is high. The

assumption of equilibrium concentration at the IMC/solder interface no longer remains

valid, especially at the early stage of the process. In the present model, the interface

kinetics was introduced through equation (8.4). In this equation the kinetics coefficient, µ

is a measure of the interface kinetics. A higher value of µ represents faster kinetics and in

126
the limiting case an infinite value of µ will result in the equilibrium concentration at the

interface. The parametric study related to the dissolution of a planar copper substrate in

the liquid solder in the present study suggested that a value of µ ≈ 4 x 10-7 m/s can be

used to represent a moderate kinetics at the IMC/solder interface.

1.0
µ = 2 E-7m/s

0.8 µ = 4 E-7m/s
µ = 6 E-7m/s

0.6
r* = r1(t)/ro

Equlbm Cond.

0.4

0.2

0.0
(a) 10-3 10-2 10-1 100 101
t (sec)

1.0

µ = 2 E-7m/s
0.8
µ = 4 E-7m/s

0.6 Equlbm Cond.


r* = r1(t)/ro

µ = 6 E-7m/s
0.4

0.2

0.0
(b) 10-2 10-1 100 101 102
t (sec)

127
Figure 8.3 continued…

1.0
µ = 2 E-7m/s
µ = 4 E-7m/s
0.8
µ = 6 E-7m/s
Equlbm Cond.

0.6
r* = r1(t)/ro

0.4

0.2

0.0
10-2 10-1 100 101 102
(c)
t (sec)

Figure 8.3: Plots showing effect of interface kinetics on particle dissolution in Sn-Ag-
0.9Cu for (a) ro = 1 µm, (b) ro = 3 µm, (c) ro = 5 µm

Kinetics effects at the particle/IMC interface were not considered in this study due

to the low diffusivity of Cu and IMC layer. Under these conditions, the assumption of

equilibrium concentration at the copper/IMC interface is valid.

In the present study the effect of interface kinetics at the IMC/solder interface was

studied (Fig 8.3) and compared with the assumption of equilibrium concentration. The

solder cell size was taken as R = 25 µm. It was evident from Fig 8.3 that the interface

kinetics had a significant effect on the dissolution of the copper particles. The model with

the equilibrium concentration at the interface (solid lines) showed very fast dissolution of

particle. The introduction of interface kinetics actually retarded the dissolution process

following that slower the interface kinetics, slower the dissolution. It was also evident

from the plots that interface kinetics had more profound effect on the smaller particles
128
than bigger ones during the entire dissolution process. However, for the large particle

(ro=5 µm, Fig 8.3 c), the convergence of these plots at longer reflow time indicated that

the kinetics had an important effect at the beginning but diffusion was the governing

factor at longer dwell time.

8.4 Effect of Initial Copper Content in Sn-Ag-Cu (SAC) Alloy

In the growing trend of lead-free solders, the Sn-Ag-Cu (SAC) alloy has now

become the standard solder in the electronic packaging industry. Other additives to the

SAC alloy are also researched to improve the performance of the solder. Besides the

eutectic composition of the Sn-Ag-Cu, these alloys are also researched for the effects of

varying amount of initial copper content in the solder.

1.0

0.8
r* = r1(t)/ro

0.6
Cu = 0.0 wt%
0.5%

0.9 %
0.4

0.2

0.0
10-1 100 101
(a)
t (sec)

129
Figure 8.4 continued…

1.0

0.8

Cu = 0.0 wt%
r* = r1(t)/ro
0.6 0.5%

0.9 %

0.4

0.2

0.0
(b) 10-1 100 101
t (sec)

1.0

0.8

Cu = 0.0 wt%
r* = r1(t)/ro

0.6
0.5%
0.9 %

0.4

0.2

0.0
(c) 10-1 100 101 102
t (sec)

Figure 8.4: Plots showing effect of initial copper concentration in Sn-Ag-Cu alloy on
particle dissolution (a) ro=1 µm (b) ro=3 µm (c) ro=5 µm

130
Since a given solder volume can only dissolve a limited amount of copper, any

predoping of the solder with copper will reduce this capacity. With the same notion, the

knowledge of the effect of the initial copper content in the SAC alloy on the dissolution

of copper particles in composite solder becomes an important issue. In this view the

effect of particle dissolution in Sn-Ag-Cu solder with different initial copper content was

studied. Moderate interface kinetics of µ23=4x10-7 m/s was considered at the IMC/solder

interface. It was evident from the plots in Fig 8.4 that although the smaller particles

completely dissolved in the solder, the dissolution was slower in the solders pre-doped

with copper. It was an obvious observation that higher the copper content of the SAC

alloy, slower was the dissolution process of the particle. The change in the dissolution

process with the change in the copper content of the solder was due to the same reason

that an increase in the copper content in the SAC alloy decreased its capacity to dissolve

the copper from the dissolving particles and hence retarded the dissolution process.

8.5 Growth of Intermetallic Compound

The present model can also predict the thickness of the IMC around the dissolving

particle. Growth of IMC around copper particles dissolving in the Sn-Ag-0.9Cu solder

was studied (Fig 8.5) for various initial particle sizes. For these plots equilibrium

concentration was considered at the IMC/solder interface. It was evident from the

numerical results that a very thin IMC layer was formed around the smaller particles

however, for the larger particles the IMC was relatively thicker. At the beginning of the

reflow process, the IMC was thin but starter to grow thicker at the longer dwell time and

was an indication of the saturated or near saturation stage of the solder.

131
101

100

IMC Thickness (µ m) 12 µm
10-1
10 µm

10-2
ro=9 µm

10-3

10-4
10-1 100 101 102 103
t (sec)

Figure 8.5: Plots showing effect of initial particle size on IMC thickness during
dissolution of copper particles in Sn-3.5Ag-0.9Cu (wt %) solder alloy

Effect of the initial copper concentration in the SAC alloy on the terminal

thickness of IMC was also analyzed (Fig 8.6). The initial particle radius was taken as ro =

5 µm. It was apparent from the plots that the solders with higher initial copper

concentration yielded a thicker IMC. Solders pre-doped with copper had lower capacity

to dissolve additional copper and therefore, saturated earlier. An early saturation of the

solder retarded the IMC dissolution and thus resulted in a thicker IMC layer during

particle dissolution.

132
100

µm)
10-1

IMC Thickness (µ
Cu = 0.9 %wt

Cu = 0.5 %wt

10-2
Cu = 0.0 %wt

10-3
10-1 100 101 102
t (sec)

Figure 8.6: Plots showing effect of initial copper content in SAC solder on IMC thickness
during dissolution of copper particles

8.6 Effect of Non-isothermal Reflow Cycle on Particle Dissolution and IMC Growth

In the previous sections, the isothermal reflow process was considered to proceed

at 250oC and the corresponding equilibrium concentrations at various interfaces were

taken from the phase diagram at the temperature. An underlying assumption in the study

was of instant heating/cooling which means that at the start of reflow process, the

particle-solder system achieved the process temperature instantly and after the desired

reflow time the system was cooled down to room temperature instantaneously. However,

in a real reflow process, it takes a finite time for the system to achieve the process

temperature. This is sometimes called the temperature ramping. The heating rates and

thereby the time it takes the solder to achieve the desired process temperature could be

different depending upon the actual furnace temperature and the volume of the solder etc.
133
The thermal cycle of a typical reflow soldering might involve a heating ramp, a holding

time and the cooling stage as shown in Fig 8.7.

245
Heating Holding Cooling

240

235
T (oC)

230

225

220

215
0 10 20 30 40 50
t (sec)

Figure 8.7: Typical thermal reflow cycle considered in the present numerical study

The present model was used to analyze the effect of the thermal history of the

actual reflow process on the dissolution of metal particles in composite solders. The

model was tested by considering the dissolution of micro-size copper particles in Sn-Ag-

Cu (SAC) solder alloy. It can be identified that besides various other rate-controlling

parameters, the thermal history of the soldering process also plays a critical role in the

overall progression of the dissolution of particles in the solder. A typical thermal cycle of

the process would involve the heating stage, holding stage and the cooling stage. The

actual heating and cooling rates may vary for different industrial processes. The process

was assumed to remain at the peak temperature of 240oC before cooling down.

134
In the present study, the inter-particle distance represented the distance between the

centers of any two neighboring spherical particles and would be equal to twice the solder

cell radius, (2R). For the present analysis it was assumed that the particles in the

composite solder are embedded in such a way that the concentration fields of dissolving

particles do not overlap. The non-dimensional parameter, r* as defined in equation (8.5)

is used for comparison of numerical results.

8.6.1 Effect of Heating Rate

Effect of the heating rate of the composite solder on the dissolution behavior (Fig

8.8 a) and IMC growth (Fig 8.8 b) around the copper particle was studied. To analyze the

effect of heating rate, the cooling rate and holding time was kept same for all three

heating rates. The plots also compare the results with the condition of instant heating. It

was evident from the plots that the dissolution of Cu particles in the molten solder was

rapid for the higher heating rates. The highest heating rate was very close to the instant

heating case. Numerical results also revealed that for the high heating rate, the dissolution

of copper particles was very rapid first and slowed down with later stage of reflow,

however, for slower heating rates the dissolution of particles was steady.

135
1.0

0.9

0.8
r* = r1(t)/ro Medium Slow Heating
Fast
0.7

0.6
Instan
t Heati
ng/Coo
ling
0.5

0.4
(a) 0 20 40 60 80 100 120 140 160
t (sec)

1.2 ng
ooli
C
g/
atin
1.0 e
tH
t an
Ins
IMC Thickness (µ m)

0.8
Slow Heating
0.6
Fast Medium

0.4

0.2

0.0
(b) 0 20 40 60 80 100 120 140 160
t (sec)

Figure 8.8: Plots showing effect of heating rate on the dissolution of copper particle (a)
and IMC growth (b) in Sn-Ag-0.9Cu (wt %) solder alloy

Rapid dissolution of the particles at high heating rate can be attributed to the

greater solubility of copper in solder at elevated temperature. It was observed that the

136
heating rate also had significant effect on the growth of IMC around the copper particles.

It was interesting to note that although the end thickness of the IMC was almost same for

all the heating rates, however, the slower heating rates resulted in higher initial thickness

of the IMC layer. A higher thickness of IMC can also be attributed to the lower value of

equilibrium concentration at low temperature.

8.6.2 Effect of Cooling Rate

In the industrial reflow soldering process the packaging assembly is cooled at

various cooling rates after the joining process is completed. The cooling rate has a

significant effect on the microstructure of the IMC at the metallization/solder interface. In

the present study, the effect of cooling rate of the solder on the dissolution of micron-size

Cu particle in composite solder was also studied. Plots in Fig 8.9 show the dissolution

behavior of copper particle and the growth of IMC at the Cu/solder interface in eutectic

Sn-Ag-Cu solder. In this study the heating rate and holding time was same for all three

cooling rates. It was evident from the plots that the dissolution rate of copper particles

was practically same for all the three cooling rates but was different from the case of

instant heating/cooling of the solder. The identical dissolution rate of copper particles for

all the cooling rates could be a result of the saturation of the solder with copper prior to

the cooling stage. It was evident from the plots that the IMC gained significant thickness

during the cooling stage. A slower cooling rate yielded even thicker IMC.

137
1.0

0.9
Fa
st
/M
0.8 ed
iu
m
r* = r1(t)/ro /S
lo
w
0.7 Co
ol
in
g
0.6

Instan
0.5 t Heati
ng/Coo
ling

0.4
(a) 0 20 40 60 80 100 120 140 160
t (sec)

1.4

1.2
Slow Cooling g
lin
oo
1.0 g/ C
µm)

ti n
Medium ea
IMC Thickness (µ

tH
0.8 tan
Ins

0.6
Fast
0.4

0.2

0.0
(b) 0 20 40 60 80 100 120 140 160
t (sec)

Figure 8.9: Plots showing effect of cooling rates on the dissolution of copper particle (a)
and IMC growth (b) in Sn-Ag-0.9Cu (wt %) solder alloy

The thickening of the IMC layer during cooling is also observed experimentally

by researchers and is said to be a result of precipitation effect. This is particularly true for

the stage when the solder was saturated with copper before cooling started.

138
8.6.3 Effect of Holding Time

The effect of holding time of the Cu-particle in the composite solder at the peak

process temperature was also analyzed through the numerical model.

1.0

0.9
10
/2
0/
0.8 30
se
c
r* = r1(t)/ro

ho
ld
0.7 in
g
t im
e
0.6

Instant
0.5 Heatin
g/Cooli
ng

0.4
0 20 40 60 80 100 120 140 160
(a)
t (sec)

1.4

1.2
30s
20s
1.0
µm)

10s
IMC Thickness (µ

0.8

0.6 g
o lin
Co
g/
0.4 tin
Hea
t
an
st
0.2 In

0.0
(b) 0 20 40 60 80 100 120 140 160
t (sec)

Figure 8.10: Plots showing effect of holding time interval on the dissolution of copper
particle (a) and IMC growth (b) in Sn-Ag-0.9Cu solder alloy

139
Plots in Fig 8.10 revealed that the holding time interval did not have a significant

effect on the dissolution of copper particles, however, the IMC grew slightly thicker at

longer holding durations. The phenomenon follows a similar explanation as for the

cooling rate on the dissolution behavior and IMC growth. At longer reflow time, after the

solder was saturated with copper any addition diffusion of Cu or Sn through the IMC

actually contributed towards the thickening of the IMC, hence the IMC grew thicker for

longer holding time intervals.

8.6.4 Effect of Initial Copper Content in Sn-Ag-Cu (SAC) Alloy

In the growing trend of lead-free solders, the Sn-Ag-Cu (SAC) alloy has now

become an industry standard. Other elements as additives to the SAC alloy are also

researched to improve the performance of the solder. Besides the eutectic composition of

the Sn-Ag-Cu, these alloys are also researched for the effects of varying amount of initial

copper content in the solder.

Since a given solder volume can only dissolve a limited amount of copper, any

predoping of the solder with copper will affect this capacity. In this view, the knowledge

of the effect of the initial copper content in the SAC alloy on the dissolution of copper

particles in composite solder becomes an important issue. In this view the effect of

particle dissolution in Sn-Ag-Cu solder with different initial copper content was studied.

A moderate heating and cooling rate were considered and the kinetics effects at the

IMC/solder interface around the copper particle was also considered. It was evident from

the plots in Fig 8.11 that dissolution of copper particles was slower in the solders pre-

doped with higher concentration of copper.

140
1.0
Ci = 1.1w
t%
0.8
Ci = 0.9
wt%
0.6
r* = r1(t)/ro

0.4

Ci
0.2 = 0.7
wt
%
0.0

(a) 0 20 40 60 80 100
t (sec)

1.2

1.0
µm)

0.8
IMC Thickness (µ

0.6
Ci = 1.1wt%
0.4

0.2 Ci = 0.9wt%
Ci = 0.7wt%
0.0
(b) 0 20 40 60 80 100
t (sec)

Figure 8.11: Plots showing effect of initial copper concentration in Sn-Ag-Cu alloy on Cu
particle dissolution (a) and IMC growth (b) during non-isothermal reflow cycle

The decline in the dissolution process with the increase in the copper content of

the solder was due to the reason that an increase in the copper content in the SAC alloy

141
decreased its capacity to dissolve the copper from the dissolving particles and hence

retarded the dissolution process. The growth of IMC showed an interesting behavior. The

initial copper concentration of 0.7 % and 0.9 % showed a uniform growth on IMC during

heating and holding stage and the IMC grew thicker during cooling phase of the process.

For the initial copper concentration of 1.1 wt %, the IMC thickness plot showed an

interesting trend. During the initial stage of heating, the IMC grew rapidly for short time

period followed by a decrease in the thickness during the uniform temperature holding

stage. The thickness of IMC started to grow again during the cooling stage of the process.

The possible explanation for the phenomenon during the heating and holding stage could

be that in this case the solder was pre-doped with the copper content close to the

equilibrium concentration at the peak temperature of the process. As the temperature was

increased gradually during the heating stage, the equilibrium concentration also increased

but remained lower than the initial copper content in the solder for most of the heating

stage. A higher copper concentration in the solder than the equilibrium concentration at

the temperature would not allow any dissolution of copper, consequently, any diffusion

of Cu or Sn through the IMC was consumed in the formation of IMC. As the temperature

reached the normal process temperature and the solubility of copper also increased, the

IMC started to dissolve and the net IMC thickness decreased during the constant

temperature holding stage. The IMC again started to grow during the cooling stage of the

process due to the same reason as explained in the previous sections.

142
8.7 Summary of Dissolution Behavior of Micro-Particles in Composite Solders

The numerical model was applied to simulate the dissolution of micro-size copper

particle in Sn-Ag-Cu composite solder during reflow process. The overall rate of

dissolution of copper particles was high at the beginning of the process but slowed down

with time. Smaller particles were observed to dissolve completely in the solder. Moderate

interface kinetics was considered at the IMC/solder interface. It was deduced from the

numerical results that the interface kinetics played a significant role in the dissolution rate

of the particles. Increase in initial copper concentration in Sn-Ag-Cu slowed down the

dissolution of the particles and increased the IMC thickness around the particles as

compared with the equilibrium concentration conditions at the interface.

To simulate the actual reflow environment in the industrial soldering process,

non-isothermal reflow conditions were considered and the effect of heating rate, holding

time and cooling rate on the dissolution of particles and IMC growth was analyzed. The

dissolution rate of particles was observed to be greater at higher heating rates but the

terminal IMC thickness remained unaffected with the heating rate. The cooling rate effect

on the dissolution of particles was negligible, however, the final IMC thickness was

observed to be higher for slower cooling rates during reflow. Similar behavior as cooling

rate was observed for the variation in holding time also.

143
CHAPTER IX

CONCLUSIONS AND FUTURE WORK

A one-dimensional model based on mass diffusion equation was developed to

predict the intermetallic phase growth and substrate dissolution during the substrate-

solder interaction. The boundary conditions at the two moving interfaces were derived

from the mass balance equations. An interface kinetic coefficient, governing the rate of

the interface reaction, was applied at the IMC/solder interface. The complete one-

dimensional model consisted of two moving interfaces viz. substrate/IMC interface and

IMC/solder interface. The model was solved numerically and the movement of the

interfaces was tracked using the coordinate transformation technique. Extensive

parametric study of the effect of interface reaction kinetic coefficient (µ), long-range

diffusion of substrate elements in the solder (DSn) and the intermetallic diffusivity (Dimc)

was conducted. Thermodynamic constraints such as substrate element (solute in the

present case) concentration at various interfaces and its solubility in solder at the given

temperature were obtained from the relevant phase diagram. The binary system of Cu as

substrate (metallization layer) and Sn-3.5wt%Ag solder were used as a representative

case and the study was conducted for the reflow conditions. However, the model can be

applied to any system by supplying the relevant physical and thermodynamic properties.

144
9.1 Model Development and Parametric Study

Numerical results revealed that the consideration of the interface kinetics at the

IMC/solder interface was essential to the accurate modeling of the process. At the

beginning of the soldering process the growth of IMC and dissolution of the metal layer

was both governed by the interface kinetics and long-range diffusion of substrate

elements in the solder bulk. The two parameters balance their respective contribution to

achieve the critical mass conservation at the moving interfaces. The solubility limit of the

copper in solder, also known as the equilibrium concentration, is the maximum possible

concentration which can exist at the IMC/solder interface. Application of the interface

reaction kinetics (µ) decreased the interface concentration, thereby slowing down the

dissolution of copper in the solder.

Dissolution of copper substrate was analyzed for cases of slow, medium and fast

interface kinetics (µ) and for fast, moderate and slow diffusivity of copper in solder (DSn).

Normalized parameters were derived for the interface velocity, interface concentration

and time. The results were also obtained and compared with the equilibrium conditions at

the IMC/solder interface.

Effect of the interface kinetics (µ), on the interface concentration and dissolution

rate of copper in solder was analyzed. The non-equilibrium interface concentration of

copper at the moving interface was observed to increase rapidly with time after the

copper dissolution started but slowed down at later stage. The rate of increase of interface

concentration was observed to be a strong function of the interface reaction kinetics and

increased with increasing value of µ. It was inferred that for an infinitely high value of µ,

the interface would attain the equilibrium concentration very quickly. It was interesting to

145
note that after the onset of dissolution process, it took a finite time for the interface

concentration to reach a constant value, highlighting the fact that non-equilibrium

conditions existed at the moving interface especially during early stages of wet soldering

process. Effect of interface reaction kinetics on the dissolution rate of the dissolving

copper substrate was studied and the results were also compared with the equilibrium

interface conditions. It was observed that for any value of µ, the copper dissolution rate

was very high at the beginning and slowed down as the process advanced. The

dissolution rate of copper was relatively slower for lower values of µ. As the value of µ

was increased, the dissolution rate approached the equilibrium condition very quickly. A

higher dissolution rate at the beginning of the process was attributed to the strong

interface reaction between copper substrate and liquid solder. Due to the intense copper-

solder reaction, copper concentration started to buildup rapidly at the copper/solder

interface. Since the copper diffusivity in solder was limited, the dissolution rate,

consequently, had to slow down after short period of time to maintain the mass balance at

the moving interface. These observations revealed that at the beginning, the copper

dissolution rate was controlled by interface kinetics and in the later stages by diffusion

mechanism.

Numerical experiments revealed that the solder diffusivity (DSn) also had the

similar effects on dissolution rate of copper as the interface kinetics (µ). The dissolution

rate of copper was rapid at the beginning of the process but decreased with time. Higher

dissolution rate was observed for the higher values of solder diffusivity. However, the

solder diffusivity had an opposite effect on the interface concentration. The rate of

increase of interface concentration was observed to be a strong function of solder


146
diffusivity (DSn) and increased with decreasing value of DSn. It was interesting to note

that similar to the interface kinetics, increasing the solder diffusivity beyond a certain

value didn’t significantly change the dissolution rate or copper concentration at the

interface. The reason for the minor changes at higher values of interface kinetics or solder

diffusivity was due to the fact that during the wet stage of soldering the dissolution

process is mutually controlled both by interface reaction kinetics and long-range

diffusion of copper in the bulk of solder. The two parameters must balance their

respective contributions to maintain the necessary mass balance at the interface. Any

excessive change in the value of one parameter will not cause significant effect in the

overall behavior, because the process will also be limited by the other controlling

parameter.

In addition to the interface kinetics and solder diffusivity, other important

parameters which affects the terminal IMC thickness and the dissolution of substrate is

the intrinsic diffusivity of the intermetallic compound (Dimc). The diffusivity of the IMC

is a measure of the rate at which the substrate and solder elements can diffuse through

intermetallic phase. Effect of IMC diffusivity on the dissolution behavior of the copper

substrate and growth of the IMC was also studied for the wet phase of Cu-Solder

interaction. A moderate value of interface kinetics (µ = 4 x 10-7 m/s) at the IMC/solder

interface and copper diffusivity in liquid tin solder (DSn = 1 x 10-9 m2/s) was used in the

calculations. Numerical results showed that the dissolution rate of copper for various

IMC diffusivity values was high at the beginning of the process but slowed down after

short period of time. The copper dissolution rate was slightly higher for low values of

IMC diffusivity.

147
Effect of different values of IMC diffusivity on the movement of the two

interfaces viz. the Cu/IMC interface and IMC/solder interface was analyzed. Tracking of

the copper/IMC interface through numerical model revealed that the recorded velocity of

copper/IMC interface always showed a negative sign for any value of IMC diffusivity.

However, the velocity of IMC/solder interface was negative upto a certain limit but then

changed to positive. A negative interface velocity of copper/IMC interface indicated that

copper always dissolved at all stages of the process. Part of the substrate was consumed

in the formation of the IMC and part was dissolved in the solder bulk. A change in the

sign of IMC/solder interface velocity was an indication of the fact that initially the IMC

was dissolving and the front of IMC was receding but at later stage, the positive interface

velocity indicated that the IMC started to grow at the IMC/solder interface into the solder

matrix. The growth of the IMC front in the solder happened when the solder in the

vicinity of the IMC became saturated with the dissolving copper elements and the IMC

can no longer dissolve in the solder. Based on the results of the numerical actual reflow

experiments it was inferred that as the substrate comes in contact with the molten solder,

the two moving fronts are established very quickly. The copper/IMC front moves at

slightly higher rate than IMC/solder interface. The difference in the velocity of the two

interfaces determines the net growth rate of the IMC at the copper/solder interface.

Effect of change in IMC diffusivity on the net dissolution of the copper and

terminal thickness of the IMC was also analyzed in the present study. It was evident from

the results that an increase in the IMC diffusivity increased the IMC thickness and the

consequential increase in the dissolved copper layer. The low IMC thickness at the lower

diffusivity value was a result of the slow diffusion of species through IMC. The net

148
growth of the IMC thickness is governed by the difference in the rates of the IMC

formation and dissolution. At low diffusivity, the IMC dissolved at almost the same rate

as it was formed resulting in a very thin IMC thickness. This was true particularly during

the early stage of the process.

9.2 Applications of Numerical Model

The model was applied to estimate the IMC thickness and copper dissolution

during reflow process. Results from the model were compared with the experimental

data. It was observed that the model was able to accurately predict the IMC growth for

the cases where the intermetallic phase grows in planar morphology. For the cases

involving planar morphology of the IMC, a constant value of IMC diffusivity worked

well. However, use of a constant value of IMC diffusivity grossly under-predicted the

thickness of the IMC with scalloped morphology during the wet phase of the Cu-Solder

interaction. Careful observation of the morphology and the support provided with the

parametric study helped conclude that the diffusion of elements across the IMC is not

constant and can not be estimated directly. The overall IMC diffusivity is a combination

of the diffusivities of various phased in the IMC which involves a complicated

morphology and phase change with time. A variable IMC diffusivity value was then

suggested using the experimentally determined and parametrically analyzed values of

diffusivity. The exponentially varying IMC diffusivity, when used in the model was able

to predict the IMC thickness and the corresponding copper dissolution very well.

The numerical model was also applied to simulate the dissolution of micro-size

copper particle in Sn-Ag-Cu composite solder during reflow process. The overall rate of

149
dissolution of copper particles was high at the beginning of the process but slowed down

with time. Smaller particles were observed to dissolve completely in the solder. Moderate

interface kinetics was considered at the IMC/solder interface. It was deduced from the

numerical results that the interface kinetics played a significant role in the dissolution rate

of the particles. Increase in initial copper concentration in Sn-Ag-Cu slowed down the

dissolution of the particles and increased the IMC thickness around the particles as

compared with the equilibrium concentration conditions at the interface.

To simulate the actual reflow environment in the industrial soldering process,

non-isothermal reflow conditions were considered and the effect of heating rate, holding

time and cooling rate on the dissolution of particles and IMC growth was analyzed. The

dissolution rate of particles was observed to be greater at higher heating rates but the

terminal IMC thickness remained unaffected with the heating rate. The cooling rate effect

on the dissolution of particles was negligible, however, the final IMC thickness was

observed to be higher for slower cooling rates during reflow. Similar behavior as cooling

rate was observed for the variation in holding time also.

9.3 Future Work

The author wishes to extend the present study to further explore the variable IMC

diffusivity and to establish the relationships pertaining to the experimental conditions and

solder composition. Experimental study especially related to the size distribution and

dissolution behavior of micro/nano-size metal particles in lead-free composite solders

during reflow is also an area of interest to extend the present study. Dissolution behavior

of the leads of electronic components, which are essentially cylindrical shaped, and the

150
prediction of IMC thickness and the dissolved lead element by the model also seems an

interesting part for future work. The author also plans to explore the extension of the

present model to two-dimensional geometry to further analyze the complicated physics of

the scallop formation.

151
BIBLIOGRAPHY

1. S. Chada, W. Laub, R. A. Fournelle, and D. Shangguan (1999) “An Improved


Numerical Method for Predicting Intermetallic Layer Thickness Developed during
the Formation of Solder Joints on Cu Substrates,” J. Electron. Mater., Vol. 28:
1194-1202.

2. M. Shaefer, W. Laub, J. M. Sabee, and R. A. Fournelle (1996) “A Numerical for


Predicting Intermetallic Layer Thickness Developed During the Formation of
Solder Joints,” J. Electron. Mater., Vol. 25: 992-2003.

3. M. Shaefer, R. A. Fournelle, and J. J. Liang (1998) J. Electron. Mater., Vol. 27:


1167.

4. O. Fouassier, J. Chazeals, and J. Silvan (2002) “Conception of a consumables


copper reaction zone for a NiTi/SnAgCu composite material,” Composites: Part A,
Vol. 33: 1391-1395.

5. D. Ma, W. D. Wang, and S. K. Lahiri (2002) “Scallop formation and dissolution of


Cu-Sn intermetallic compound during solder reflow,” J. Appl. Phys., Vol. 91: 3312
– 3317.

6. H. K. Kim and K. N. Tu (1995) “Rate of consumption of Cu in soldering


accompanied by ripening,” Appl. Phys. Lett. Vol. 67: 2002 – 2004.

7. Y. G. Lee, and J. G. Duh (1998) “Characterizing the formation and growth of


intermetallic compound in the solder joint,” J. Mater. Sc. Vol. 33: 5569 – 5572.

8. T. M. Korhonen, P. Su, S. J. Hong et. al. (2000) “Reaction of Lead-Free Solders


with CuNi Metallizations,” J. Electron. Mater., Vol. 29: 1194 – 1199.

9. K. Suganuma (2001) “Advances in lead-free electronics soldering,” Current


Opinion in Solid State and Materials Science, Vol. 5: 55 – 64.

10. S. W. Chen and Y. W. Yen (1999) “Interfacial Reaction in Ag-Sn/Cu Couples,” J.


Electron. Mater., Vol. 28: 1203 – 1208.

11. Y. C. Chan, C. K. Alex, and J. K. L. Lai (1998) “Growth kinetic studies of Cu-Sn
intermetallic compound and its effect on shear strength of LCCC SMT solder

152
joints,” Mat. Sc. Eng. B, Vol, B55: 5 – 13.

12. K. H. Prakash and T. Sritharan (2001) “Interface Reaction Between Copper and
Molten Tin-Lead Solders,” Acta Mater., Vol. 49: 2481 – 2489.
13. Y. G. Lee and J. G. Duh (1999) “Interfacial morphology and concentration profile
in the unleaded solder/Cu joint assembly,” J. Mater. Sc., Vol. 10: 33 – 43.

14. Rudolf Strauss (1994) “Surface Mount Technology,” Butterworth-Heinemann Ltd.

15. Giles Humpston, David M. Jacobson (1993) “Principles of Soldering and Brazing,”
ASM International.

16. Donald P. Seraphim, Ronald C. Lasky and Che-Yu Li (1989) “Principles of


Electronic Packaging,” McGraw-Hill Book Company.

17. Armin Rahn (193) “The Basics of Soldering,” John Wiley & Sons Inc.

18. A. J. Sunwoo, J. W. Morris, Jr., and G. K. Lucey, Jr. (1992), Metall. Trans. A:23:
1323.

19. A. J. Sunwoo, H. Hayashigaani, J. W. Morris, Jr., and G. K. Lucey, Jr. (1991), JOM
43 (6): 21.

20. K. N. Tu, T. Y. Lee, J. W. Jang, L. Li, D. R. Frear, K. Zeng and J. K. Kivilahti


(2001) “Wetting reaction versus solid state aging of eutectic Sn-Pb on Cu” J. of
Appl. Phys. 89: 4843-4849.

21. A. S. Zuruzi, C.-h. Chiu, S. K. Lahiri, and K. N. Tu (1991) “Roughness evolution of


Cu6Sn5 intermetallic during soldering” J. of Appl. Phys. 86: 4916-4921.

22. G. Ghosh (2000) “Coarsening kinetics of Ni3Sn4 scallops during interfacial reaction
between liquid eutectic solders and Cu/Ni/Pd metallization” J. of Appl. Phys. 88:
6887-6897.

23. B. J. Lee, N. M. Hwang and H. M. Lee (1997) “Prediction of interface reaction


products between Cu and various solder alloys by thermodynamic calculation” Acta
Mater. 45: 1867-1874.

24. L. A. Clevenger, B. Arcot, W. Ziegler, E. G. Colgan, Q. Z. Hong, F. M. d’Heurle,


C. Cabral, Jr., T. A. Gallo, and J. M. E. Harper (1998) “Inter-diffusion and phase
formation in Cu(Sn) alloy films” J. of Appl. Phys. 83: 90-99.

25. J. London and D. W. Ashall, (1986), Brazing and Soldering 11: 49.

153
26. P. G. Kim, J. W. Jang, T. Y. Lee and K. N. Tu (1999) Interfacial reaction and
wetting behavior in eutectic Sn-Pb solder on Ni/Ti thin films and Ni foils, J. of App.
Phys. 86: 6746-6751.

27. Y. Wang and K. N. Tu (1995), Ultrafast intermetallic compound formation between


eutectic Sn-Pb and Pd where the intermetallic is not a diffusion barrier, App. Phys.
Lett. 67(8): 1069-1071.

28. M. Faizan, R. A. McCoy, D. C. Lin, and G.-X. Wang (2003), An investigation of


copper dissolution and microstructural development in lead-free solders,
Proceedings of ASME IMECE.

29. M. Faizan, R. A. McCoy, D. C. Lin, T. S. Srivatsan and g.-X. Wang (2003), An


Investigation of Copper Dissolution and Formation of Intermetallic Compound
During Soldering, Proceedings of ASME Summer Heat Transfer Conference.

30. K. N. Tu and K. Zeng, (2001) “Tin-Lead (Sn-Pb) solder reaction in flip chip
technology, Mat. Sc. And Engr. R 34, 1-58.

31. A. A. Liu, H. K. Kim, K. N. Tu and P. A. Totta (1996), Spalling of Cu6Sn5


spheroids in the soldering reaction of eutectic Sn-Pb on Cr/Cu/Au multilayer thin
film, J. App. Phy. 80(5) 2774-2780.

32. J. G. Lee, F. Guo, K. N. Subramanian and L. P. Lucas, (2002) “Intermetallic


morphology around Ni particles in Sn-3.5%Ag solder”, Soldering and Surface
Mount Technology, 14/2, 11-17.

33. C. B. Lee, J. W. Yoon, S. W. Suh, S. B. Jung et.al., (2003) “Intermetallic compound


layer formation between Sn-3.5%Ag BGA solder ball and Cu, Au, Ni substrates”, J.
of Mat. Sc., 14, 487-493.

34. H. K. Kim and K. N. Tu (1996) “Kinetic analysis of the soldering reaction between
eutectic Sn-Pb alloy and Cu accompanied by ripening” Phys. Rev. B, 53:23, 16027
– 16034.

35. S. Bader, W. Gust and H. Hieber (1994), “Rapid formation of Intermetallic


compounds by inter-diffusion in the Cu-Sn and Ni-Sn systems”, Acta. Mater. Vol.
43, No. 1, 329-337.

36. A. M. Gusak and K. N. Tu (2002), “Kinetic theory of flux driven ripening”, Phys.
Rev. B, 66:11, 1- 14.

37. V. I. Dybkov, Growth kinetics of chemical compound layers, (1998), Cambridge


International Science, Canbridge.

154
38. F. J. Vermolin and S. van der Zwaag (1996), A numerical model for the dissolution
of spherical particles in binary alloys under mixed mode control, Mat. Sc. & Eng.
A, Vol. 220, pp. 140-146.

39. F. J. Vermolin C. Vuik and S. van der Zwaag (1998), The dissolution of a
stioichiometric second phase in ternary alloys: a numerical analysis, Mat. Sc. &
Eng. A, Vol. 246, pp. 93-103.
40. S. P. Chen, M. S. Vossenberg, F. J. Vermolin, J. van der Langkruis and S. van der
Zwaag (1999), dissolution of β particles in an Al-Mg-Si alloy during DSC runs,
Mat. Sc. & Eng. A, Vol. 272, pp. 250-256.

41. C. Vuik, A. Segal, F. J. Vermolen (1999), A conservative discretization for a Stefan


problem with an interface reaction at the free boundary, Comput and Visual in Sci,
3: 109-114.

42. F. J. Vermolen and C. Vuik (2000), A mathematical model for the dissolution of
particles in multi-component alloys, J. of Comp. and App. Math., 126: 233-254.

43. F. J. Vermolin, C. Vuik and S. van der Zwaag (2002), A mathematical model for
the dissolution of stoichiometric particles in multi-component alloys, Mat. Sc. &
Eng. A, Vol. 328, pp. 14-25.

44. F. J. Vermolen, C. Vuik and S. van der Zwaag (2005), Cross-diffusion controlled
particle dissolution in metallic alloys, Comput and Visual in Sci, 8: 27-33.

45. Z. Mei, A. J. Sunwoo and J. W. Morris Jr. (1992), Analysis of low-temperature


intermetallic growth in copper-tin diffusion couple, Metal. Trans. A, 23, 857-863.

46. Ahmed Sharif, Y. C. Chan, Rashed Adnan Islam (2003), Effect of volume in
interfacial reaction between eutectic Sn-Pb solder and Cu metallization in
microelectronic packaging, Mat. Sci. & Eng. B, 106: 120-125.

47. M. O. Alam and Y. C. Chan (2003), Effect of 0.5 wt% Cu addition in Sn-3.5%Ag
solder on the dissolution rate of Cu metallization, J. of App. Phy., 94:12, 7904-
7909.

48. S. V. Patankar (1980), Numerical Heat Transfer and Fluid Flow, Hemisphere, New
York.

49. D. Q. Yu, C. M. L. Wu, C. M. T. Law, L. Wang and J. K. L. Lai (2004),


Intermetallic compound growth between Sn-3.5%Ag lead-free solder and Cu
substrate by dipping method, J. of Alloys and Comp., 392: 192-199.

155
50. A. Hayashi, C. R. Kao, and Y. A. Chang (1997), Reactions of solid copper with
pure liquid tin and liquid tin saturated with copper, Scripta Mater. 37:4, 393-398.

51. I. Kaur… fundamentals of Grain boundary diffusion….

52. R. A. Lord and A. Umantsev (2005), Early stages of soldering reactions, J. of App.
Phy. 98: 1-11.

53. J. Görlicin, G. Schmitz and K. N. Tu (2005), On the mechanism of the binary Cu/Sn
solder reaction, App. Phy. Lett. 86, 053106, 1-3.

54. Mohammad Faizan, (2003), An investigation of copper dissolution and Formation


of Intermetallic compound during soldering, M.S. Thesis.

55. Fu Guo, (2007), Composite lead-free electronic solders, J Mater Sci: Mater Electron
18:129–145

56. J. L. Marshall, J. Sees, J. Calderon, (1992) Micro-characterization of Composite


Solders, Proceedings of Technical Program, Nepcon West Conference, Anahein,
CA, Cahners Exhibition Group, Des Plains, IL, 3, , pp. 1278–1283

57. J. L. Marshall, J. Calderon, J. Sees, G. Lucey, J.S. Hwang, IEEE Trans. Comp. Hyb.
Manufact. Technol. 14(4), 698 (1991)

58. J. L. Marshall, J. Calderon, (1997) Sold. Surf. Mount Technol. 9(3), 6

59. J.L. Marshall, J. Calderon, (1997) Sold. Surf. Mount Technol. 9(3), 11–14

60. R.F Pinizzotto, Y. Wu, E.G. Jacobs, L.A. Foster, (1992) Microstructural
Development in Composite Solders Caused by Long Term, High Temperature
Annealing, Proceedings of Technical Program, Nepcon West Conference, Anaheim,
CA, Cahners Exhibition Group, Des Plains, IL, 3, p. 1284

61. Y. Wu, J.A. Sees, C. Pouraghabragher, L.A. Foster, J.L. Marshall, E.G. Jacobs, R.F.
Pinizzotto, J. Electron. Mater. 22(7):769 (1993)

62. S.M.L. Sastry, T.C. Peng, R.J. Lederich, K.L. Jerina, C.G. Kuo, (1992)
Microstructures and Mechanical Properties of In-situ Composite Solders,
Proceedings of Technical Program, Nepcon West Conference, Anaheim, CA,
Cahners Exhibition Group, Des Plains, IL, 3, p. 1266

63. H.S. Betrabet, S.M. McGee, J.K. McKinlay, (1991) Scripta Metall. Mater. 25,
2323–2328

156
64. D.C. Lin, T.S. Srivatsan, G.-X. Wang and R. Kovacevic, (2006) Microstructural
development in a rapidly cooled eutectic Sn–3.5% Ag solder reinforced with copper
powder, Powder Technology 166, p 38–46

65. D. C. Lin, G.X. Wang, T.S. Srivatsan, Meslet Al-Hajri, M. Petraroli (2002), The
influence of copper nano-powders on microstructure and hardness of lead-tin
solder, Materials Letters 53, p 333– 338

66. K. Mohan Kumar, V. Kripesh, Lu Shen, Andrew A.O. Tay (2006), Study on the
microstructure and mechanical properties of a novel SWCNT-reinforced solder
alloy for ultra-fine pitch applications, Thin Solid Films 504, p 371 – 378

67. Paul G. Harris, Kaldev S. Chaggar (1998), The role of intermetallic compounds in
lead-free soldering, Soldering and surface mount technology, 38-52

68. M. E. Loomans and M. E. Fine (2000), Metall. Mater. Trans. A 31A, 1155

69. K.-W. Moon, W. J. Boeetinger, U. R. Kattner, F. S. Biancaniello and C. A.


Handwerker (2000), J. Electron. Mater. 29, 1122

157
APPENDICES

158
APPENDIX A

COORDINATE TRANSFORMATION FOR MOVING BOUNDARY

The substrate/solder interface was fixed in the computational domain by

transforming the original physical coordinates into computation coordinates (Fig A1).

Assuming the substrate and the solder as η and ξ domain respectively, the following

transformation can be used.

r r −ε
η= and ξ= (A1)
ε b −ε

where r = ε(t) represents the substrate/solder interface and b is the total thickness of the

substrate and solder. A typical control volume element is explained in Fig A1.

Substrate Solder
Vi

b
r
r=0 ε(t)

η ξ
η=0 η=1 ξ =1
ξ=0

Figure A1: Coordinate system used for transformation

159
∆η w ∆η e
w e

J-1 J J+1

∆η J-1 ∆η J ∆η J+1

Figure A2: Control volume element for the interface tracking

To treat the large concentration gradient at the interface, a non-uniform grid can

be used with higher grid density near the interface. The position of the faces of the

control volume elements can be calculated by the following relation [48]

ps
 J −1 
ηs (J ) =   J = 1, 2, 3,…M-1 (A2 a)
M −2

pl
 J −1 
ξ s (J ) =   J = 1, 2, 3,…N (A2 b)
 N −1

where M – 1 and N are the total number of control volume faces in the solid and liquid,

respectively, and the subscripts s means the faces of the elements i.e. the locations

denoted by e or w in Fig A2. The grid nodes are placed at the center between two

adjacent faces for a total of M – 2 internal nodes in the substrate region and N – 1 in the

solder region. The interface and each of the two boundary elements are considered as

zero-volume elements. There are, therefore, a total of M + N grid nodes. The exponent ps

and pl are arbitrary parameters used to control the variation of element width in the

substrate and solder domain respectively. A typical value of these parameters between 1.5

and 2.0 gave satisfactory results.


160
APPENDIX B

DISCRETIZATION OF DIFFERENTIAL EQUATIONS

The central differencing of the convective term in the concentration equation after

coordinate transformation may create a stability problem at high interface velocities. A

universal function derived for these schemes by Patankar [48] can be used to derive the

difference equations for node concentration. The concentration diffusion equation after

the coordinate transformation can be written for the substrate region (0 ≤ η ≤ 1) as

∂C S 1 ∂  a ∂C  dε  ∂ 
ε2 = a  (η DS ) S  + ε  (ηC S ) − C S  (B1a)
∂t η ∂η  ∂η  dt  ∂η 

Equation (B1a) reduces to

∂C S ∂ ∂C dε ∂C S
η aε 2 = (η a DS ) S + η a +1ε (B1b)
∂t ∂η ∂η dt ∂η

if b is the combined width of the substrate and solder domains, i.e. b = ro + Ro then in the

solder region (0 ≤ ξ ≤ 1) the concentration diffusion equation will be as

∂C L ∂  ∂C L  dε ∂C L
κ (b − ε ) 2 =  κDL  + κ (b − ε )(1 − ξ ) (B2)
∂t ∂ξ  ∂ξ  dt ∂ξ

In equation (B2)
161
κ = [(b − ε )ξ + ε ]a

The finite difference equations for the nodal concentrations are obtained directly by

integrating the transformed differential equations (B1) and (B2) from time t to t + ∆t over

each control volume element. Equations (B1) and (B2) in terms of solute fluxes GS and

GL in substrate and solder domains respectively can be written as

∂C S ∂GS ∂FS
η aε 2 + − CS = 0 (B3)
∂t ∂η ∂η

∂C L ∂G L ∂FL
κ (b − ε ) 2 + − CL = 0 (B4)
∂t ∂ξ ∂ξ

where

∂C S
GS = FS C S − DSη a (B5 a)
∂η

∂C L
G L = FL C L − κDL (B5 b)
∂ξ


FS = −η a +1ε (B5 c)
dt


FL = −κ (b − ε )(1 − ξ ) (B5 d)
dt

Following Patankar [48], the difference equation for concentration at node J (Fig 4.3) can

then be derived as

PJ C Jn +1 = E J C Jn++11 + WJ C Jn−+11 + S J (B6)

162
where the coefficients in equation (B6) are as follows

E J = De ( A(| Pe |) + − Pe − 0 ) (B7 a)

WJ = Dw ( A(| Pw |) + Pw − 0 ) (B7 b)

PJ = PJo + E J + WJ (B7 c)

S J = PJo C Jn (B7 d)

and

Fe Fw
Pe = Pw = (B8 a)
De Dw

η Ja ε 2
P =
J
o
∆η J for substrate region (B8b)
∆t

κ (b − ε ) 2
PJo = ∆η J for solder region (B8c)
∆t

The definitions for De, Dw, Fe and Fw are given as follows for substrate region

η ea DSe η wa DSw
De = Dw = (B9a)
δη e δη w

dε dε
Fe = −η ea+1ε Fw = −η wa +1ε (B9b)
dt dt

and De, Dw, Fe and Fw for the solder region

κDLe κDLw
De = Dw = (B9c)
δξ e δξ w

163
dε dε
Fe = −κ (b − ε )(1 − ξ e ) Fw = −κ (b − ε )(1 − ξ w ) (B9d)
dt dt

The function A(|P|) in equations (B7) was derived by Patankar [48] and depends on the

selection of a differencing scheme. For the power-law scheme used in the present study,

the function has the following definition

A(| P |) = max{0, (1 − 0.1 | P |) 5 } (B10)

where max{A, B} denotes the larger of A and B.

164

You might also like