You are on page 1of 114

Development of Constitution Diagram for Dissimilar Metal Welds in Nickel Alloys and

Carbon and Low-Alloy Steels

THESIS

Presented in Partial Fulfillment of the Requirements for the Degree Master of


Science in the Graduate School of The Ohio State University

By

Elijah Katunich Gould

Graduate Program in Welding Engineering

The Ohio State University

2010

Master's Examination Committee:

Professor John Lippold, Advisor

Professor Sudarsanam Babu


Copyright by

Elijah Katunich Gould

2010
ABSTRACT

The prevalence of dissimilar metal joints between carbon and low-alloy

steels and solid-solution strengthened nickel-base filler materials has resulted in an

increased desire to understand the metallurgy that controls the properties of welds

between these materials. The increased use of these combinations has been driven

by their successes in applications where austenitic stainless steel weld metals have

failed. Their superior resistance to stress corrosion cracking, elevated temperature

properties, and close match in coefficient of thermal expansion with steels make

them an ideal choice in a variety of situations. However, these materials do have

weldability issues of their own to mitigate. Among the more formidable concerns

regarding dissimilar joints between these materials is the formation of martensite in

the composition transition zone at the fusion boundary, which often accompanies

the formation of Type II boundaries during welding. The presence of substantial

amounts of martensite in this region greatly increases the likelihood of failure

during service. The Schaeffler diagram has been demonstrated to have an ability to

predict the formation of martensite to a certain degree in such situations. Having

been developed based on 300 series stainless steels its overall predictive capability

with regards to these materials can be improved upon.

ii
The goal of this study is to use previously established techniques to develop a

predictive diagram which has greater accuracy when applied to dissimilar welds

between carbon and low-alloy steels and nickel-base alloys. Button melt samples

were created by mixing various proportions of carbon and low-alloy steels and

nickel-base filler metals to produce samples whose microstructure consist of

varying amounts of martensite and austenite. The use of a martensite

characterization technique based on magnetic response (ferrite number) using the

Magne-Gage proved an effective way of determining martensite content of button

melt samples. Additional characterization techniques were used to establish a

relationship between magnetic response and martensite volume fraction. Magnetic

response data generated by the Magne-Gage was analyzed using statistical methods

and fit to a model in order to allow prediction of magnetic response based on alloy

content. The model was then used for construction of a predictive diagram for

dissimilar metal welds between these materials. Finally, predictive capabilities of

the developed diagram were assessed and discussed.

iii
DEDICATION

Dedicated to my mother, without whose constant love, support, and encouragement

I would never have accomplished what I have in life.

iv
ACKNOWLEDGMENTS

I would like to thank my advisors, Professor John Lippold and Professor

Boian Alexandrov whose guidance on this project was pivotal to its success, and

from whom I was able to learn continuously along the way.

Thanks are also extended to Professor Theodore Allen of the Integrated

Systems Engineering Department at the Ohio State University, and Dr. Jerry Gould of

the Edison Welding Institute for their advice and assistance with statistical analysis

preformed as part of this work.

Thanks to Professor Suresh Babu for his advice during parts of this project,

and for taking the time to serve as a member of my master’s examination

committee.

Thanks to the members of the Welding and Joining Metallurgy Group, both

past and present, whose contribution to an enjoyable work atmosphere cannot be

overstated. Thank you to Adam Hope for his work digitizing the Schaeffler diagram.

Thank you to Jeff Rodelas for his assistance with EBSD work.

Finally, I would like to thank the many people in my life whose support,

advice, and encouragement during this project proved essential to its completion.

v
VITA

November 22, 1983………………………Born – Valparaiso, IN. United States of America

June, 2008…................................................B.S.W.E. The Ohio State University

Columbus, Ohio

2008 – Present……………………………...Graduate Research Associate

The Ohio State University

FIELD OF STUDY

Major Field: Welding Engineering

vi
TABLE OF CONTENTS

Abstract ....................................................................................................................................... ii

Dedication.................................................................................................................................. iv

Acknowledgments .................................................................................................................... v

Vita ............................................................................................................................................... vi

Table of Contents.................................................................................................................... vii

List of Tables ............................................................................................................................... x

List of Figures ........................................................................................................................... xi

1. Introduction ...................................................................................................................... 1

2. Background ....................................................................................................................... 3

2.1 Carbon and Low-Alloy Steels............................................................................................ 3

2.1.1 Composition..................................................................................................................................... 3

2.1.2 Metallurgy ........................................................................................................................................ 5

2.1.3 Weldability ..................................................................................................................................... 12

2.1.4 Applications ................................................................................................................................... 13

2.2 Solid-solution strengthened nickel-base alloys ..................................................... 13

2.2.1 Composition................................................................................................................................... 13

vii
2.2.2 Metallurgy and Mechanical Properties............................................................................... 14

2.2.3 Weldability ..................................................................................................................................... 15

2.2.4 Applications ................................................................................................................................... 18

2.3 Weld Metal Constitution Diagrams ............................................................................. 19

2.3.1 Early Constitution Diagrams................................................................................................... 20

2.3.2 The Schaeffler Diagram ............................................................................................................. 23

2.3.3 Austenitic-Martensitic Constitution Diagrams................................................................ 27

2.4 Nickel-Steel Dissimilar Metal Welding ...................................................................... 31

2.4.1 Applications ................................................................................................................................... 31

2.4.2 Weldability ..................................................................................................................................... 32

3. Objectives ......................................................................................................................... 39

4. Experimental Procedures........................................................................................... 41

4.1 Materials............................................................................................................................... 41

4.2 Sample Production ........................................................................................................... 46

4.3 Characterization ................................................................................................................ 48

4.3.1 Sample Prep ................................................................................................................................... 48

4.3.2 Optical Microscopy ..................................................................................................................... 51

4.3.3 Optical Phase Fraction Analysis ............................................................................................ 51

4.4 Ferrite Number Measurement...................................................................................... 52

4.5 X-ray Diffraction ................................................................................................................ 54

4.6 Microhardness Testing.................................................................................................... 56

4.7 Chemical Analysis ............................................................................................................. 56

4.8 Data Analysis ...................................................................................................................... 57

viii
4.9 Diagram Development .................................................................................................... 58

5. Results and Discussions .............................................................................................. 59

5.1 Button Melt Microstructures......................................................................................... 59

5.2 Microhardness Testing.................................................................................................... 63

5.3 Button Melt Analysis ........................................................................................................ 65

5.4 Magne-Gage Ferrite Number (FN) Measurement .................................................. 66

5.5 X-ray Diffraction ................................................................................................................ 72

5.6 Point Count .......................................................................................................................... 75

5.7 Development of FN to martensite phase fraction relation................................. 78

5.8 Data Analysis & Diagram Development .................................................................... 79

5.9 Use of the Diagram ............................................................................................................ 83

6. Conclusions ...................................................................................................................... 85

References ................................................................................................................................ 87

A. Additional Micrographs .............................................................................................. 91

B. Button Melt Sample Data ............................................................................................ 97

ix
LIST OF TABLES

Table 4.1: Chemical composition of materials used in this study ......................... 43

Table 4.2: List of etchants used in this study................................................................ 51

Table 5.1: X-ray diffraction and Magne-Gage analysis results for selected

samples ............................................................................................................................. 73

Table 5.2: Point Counting, X-ray diffraction and Magne-Gage analysis results

for selected samples ..................................................................................................... 76

Table B.1: Button melt sample composition and ferrite number data ............... 98

x
LIST OF FIGURES

Figure 2.1 Fe-C phase diagram (region of interest for steels) .................................. 8

Figure 2.2: Modified Strauss-Maurer Diagram [25] ................................................... 21

Figure 2.3: The Schaeffler Diagram (1947) [3] ............................................................ 25

Figure 2.4: The Schaeffler Diagram (1949)................................................................... 27

Figure 2.5: Diagram of Self et al. [33]............................................................................... 29

Figure 2.6: Schaeffler Diagram as modified by Béres [34] ...................................... 31

Figure 2.7: Schematic representation of Type II boundaries and their

formation [35] ................................................................................................................ 33

Figure 2.8: Example usage of Schaeffler diagram as it relates to dissimilar

metal welding [12] ........................................................................................................ 34

Figure 4.1: Schaeffler diagram with materials of interest plotted ....................... 44

Figure 4.2: Schaeffler diagram enlarged to show study focus area ...................... 45

Figure 4.3: Schematic representation of button melting setup ............................. 48

Figure 4.4: Calibration curves used to calculate FN using Magne-Gage .............. 54

Figure 5.1: TCP phase within interdendritic regions of 8630-625 85% dilution

button melt sample (Etch 2) ...................................................................................... 60

Figure 5.2: : Predominantly austenitic microstructure of F22-800H 55%

dilution button melt sample (Etch 4) ..................................................................... 61


xi
Figure 5.3: Predominantly martensitic structure with pockets of retained

austenite in 100% 8630 button melt sample (Etch 2) ..................................... 62

Figure 5.4: Duplex martensite + austenite microstructure of P92-800H 74%

button melt sample (Etch 5) ...................................................................................... 63

Figure 5.5: Vickers hardness indents in selected phases of F22-800H 63%

dilution button melt sample ...................................................................................... 64

Figure 5.6: : Vickers hardness indents in selected phases of F22-800H 60%

dilution button melt sample ...................................................................................... 65

Figure 5.7: Ferro-fluid staining of F22-800H 55% button melt sample ............. 66

Figure 5.8: JMatPro graphical output of solidification nature of compositional

makeup of P92-800H 88% dilution ........................................................................ 68

Figure 5.9: JMatPro graphical output of nickel and chromium segregation

within δ ferrite during solidification...................................................................... 69

Figure 5.10: JMatPro graphical output of solidification nature of enriched δ

ferrite in P92-800H 88% dilution ........................................................................... 69

Figure 5.11: FN versus dilution of button melt samples determined by Magne-

Gage .................................................................................................................................... 71

Figure 5.12: FN versus dilution of button melt samples determined by Magne-

Gage with superimposed XRD analysis values .................................................... 73

Figure 5.13: X-ray diffraction spectra of X80-C276 73% dilution button melt

sample ............................................................................................................................... 74

xii
Figure 5.14: X-ray diffraction spectra of F22-800H 60% dilution button melt

sample ............................................................................................................................... 75

Figure 5.15: Plot of martensite phase fraction vs. FN for this study .................... 79

Figure 5.16: Measured FN versus FN as predicted by regression equation ...... 80

Figure 5.17: Preliminary constitution diagram for microstructure prediction

in dissimilar metal welds ........................................................................................... 82

Figure 5.18: Preliminary constitution diagram for microstructure prediction

in dissimilar metal welds with FN data plotted by developed equivalency

formulae ........................................................................................................................... 83

Figure A.1: Predominantly martensitic microstructure of 100% dilution 8630

steel (Etch 5) ................................................................................................................... 91

Figure A.2: 8630-625 85% dilution microstructure (Etch 5) ................................. 92

Figure A.3: F22-800H 76% dilution microstructure (Etch 5)................................. 92

Figure A.4: X80 – C276 70% dilution microstructure (Etch 2) .............................. 93

Figure A.5: P92-800H 63% dilution microstructure (Etch 2) ................................ 93

Figure A.6: F22-800H 60% dilution microstructure (as polished) ...................... 94

Figure A.7: A508-FM52M 82.5% dilution microstructure (as polished)............ 94

Figure A.8: F92-800H 63% dilution microstructure (ferro-fluid) ........................ 95

Figure A.9: X80-C276 70% dilution microstructure (ferro-fluid) ........................ 95

Figure A.10: X80-C276 78% dilution microstructure (ferro-fluid)...................... 96

Figure A.11: F22-800H 55% dilution microstructure showing possible TCP

phases (Etch 4) ............................................................................................................... 96

xiii
1. INTRODUCTION

CHAPTER 1

INTRODUCTION

The use of nickel-base filler materials as part of dissimilar metal

combinations involving carbon and low-alloy steels is commonplace in a number of

applications including the petrochemical, power generation industries and the

fabrication of pressure vessels to name just a few. The nickel-base filler metals offer

several advantages over the use of austenitic stainless steels typically associated

with dissimilar metal welding. They minimize coefficient of thermal expansion

mismatches common in dissimilar metal joints welded with austenitic stainless

steels, and offer superior corrosion resistance in chloride containing environments

[1, 2].

The use of constitution diagrams for prediction of weld metal microstructure

in stainless steel welds has been the subject of study for over 50 years. The

Schaeffler diagram, created in 1947 [3], and published in its final form in 1949 [4]

was one of the first such diagrams. Due to its large composition range it has been

used for prediction of weld metal microstructures in a number of dissimilar weld

metal combinations [5-7] including those between nickel-base filler metals and

carbon and low-alloy steels. This diagram was created based on the 300 series
1
stainless steels, and a select number of ferritic and martensitic stainless steels. For

this reason, its accuracy when applied to the aforementioned combinations is

unknown.

Previous research has illustrated the effectiveness of the Schaeffler diagram

for prediction of the presence of martensite in the weld metal of dissimilar metal

welds involving nickel-base alloys and carbon and low-alloy steels. The relative

amount of martensite present in the fusion zone has been closely attributed to

dilution level required for martensite formation based on the Schaeffler diagram.

Furthermore, the presence of increased amounts of martensite along the fusion

boundary has been shown to increase the susceptibility of these welds to cracking

by various mechanisms, including hydrogen induced cracking typically associated

with martensite microstructures [6].

The effects of microstructure on cracking susceptibility in dissimilar metal

welds where nickel-base alloys and carbon and low-alloy steels are involved

highlights the need for a diagram that is specifically intended for prediction of weld

metal microstructure in these combinations. The purpose of this study is to develop

such a diagram using combinations of nickel-base filler materials and carbon and

low-alloy steels. Such a diagram will provide information about the microstructure

of weldments between materials that the Schaeffler diagram is not intended for and

provide an ideal starting point for determination of dissimilar metal combinations

for use in a variety of applications.

2
2. BACKGROUND

CHAPTER 2

BACKGROUND

2.1 Carbon and Low-Alloy Steels

Carbon and low-alloy steels account for 95% of the world’s construction and

fabrication metals [8]. Low cost and favorable mechanical properties make these

materials ideal for a variety of applications. Joining of these materials is predictably

common due the large quantity being utilized. Based on composition, strength, heat-

treatment, and high-temperature properties carbon and low-alloy steels are

generally categorized into the following five groups [8]:

1. Carbon Steels

2. High-strength low-alloy steels

3. Quenched and tempered steels

4. Heat-treatable low-alloy steels

5. Chromium-molybdenum steels

2.1.1 Composition

Carbon steels form the basis from which all the other steel alloy groups are

created. Due to its ability to impart significant strength when added to iron in small

3
amounts, carbon is the most important alloying element in steel. As a result the Fe-C

system is the basis for all carbon and low-alloy steel grades. Carbon steels contain

carbon contents of less than 1.0 wt.%. Carbon steels are designated in order of

increasing carbon content as either low-carbon, mild-steel, medium-carbon steel, or

high carbon steel [8]. In addition to carbon, manganese, silicon, and copper may also

be present in carbon steels as a result of the steel making practice [9]. Manganese

contents in carbon steels are kept below 1.65 wt.%, and copper and silicon contents

are kept below 0.60 wt.% [8].

High-strength low-alloy (HSLA) steels contain small amounts of varying

combinations of chromium, nickel, molybdenum, copper, nitrogen, vanadium,

niobium, titanium, and zirconium based on mechanical property requirements and

product thickness. In general the contents of these elements individually are kept

below 1.0 wt.%. Carbon is present in these alloys in the range of 0.05-0.25 wt.%.

Manganese is present at levels below 2.0 wt.% for increased formability and

weldability [9].

Quenched and Tempered steels have chemical compositions similar to the

HSLA steels but can also include increased amounts of varying elements to promote

specific properties. Addition of 8-9% nickel to A553 steel results in increased low-

temperature toughness in this grade for instance.

Heat-treatable low-alloy (HTLA) steels contain the same alloying additions as

the HSLA and quenched and tempered steels. However, these steels contained

increased amounts of various elements relative to these other steels. The increase in

4
alloying additions results in the ability to be heat-treated to high-strength following

welding, as the name would suggest. Carbon contents in these alloys range from

0.25 wt.% to 0.45 wt.%. Because of cracking concerns associated with heat-

treatment of these alloys sulfur and phosphorus levels are kept below 0.040 wt% or

lower dependent on steel melting technique and intended final strength level [9].

Chromium-molybdenum steels are so named because of the increased

alloying additions of these elements. Chromium is added to improve oxidation and

corrosion resistance in these steels. Molybdenum is added to increase high

temperature properties. Additions of small amounts of titanium, vanadium,

niobium, and nitrogen are also used to increase high-temperature properties. These

elements form high temperature precipitates that increase resistance to grain

boundary sliding thus increasing creep strength. Alloying additions are also tailored

such that the desired matrix of ferrite, bainite, or martensite can be achieved with

proper heat treatment. Chromium contents in these steels range from 0.5–9 wt.%,

and molybdenum contents range from 0.5-1.0 wt.%. Carbon contents in these steels

are generally kept below 0.15 wt.%.

2.1.2 Metallurgy

The metallurgy of carbon and low-alloy steels is based on the iron-carbon

phase diagram and the non-equilibrium transformations that occur under certain

conditions. The compositions previously discussed are chosen on the basis of the

effects different alloy additions have on these reactions. Additionally, because of the

5
importance of welding as a means for joining these materials for many applications

the effect of welding on the metallurgy of these alloys is equally important.

The Fe-C phase diagram shown in Figure 2.1 [9] provides a starting point for

a basic understanding of the metallurgy of carbon steels. This diagram illustrates

that pure iron will exist as a variety of different phases, each with different

properties dependent on temperature. The transformation of austenite (γ) to

various decomposition products is the most important metallurgical aspect of steels

to be considered. The equilibrium Fe-C phase diagram shows that two

transformation products are possible from the decomposition of austenite. Under

equilibrium conditions, or very slow cooling, austenite will transform into ferrite (α)

or pearlite, a mixture of ferrite and cementite (Fe3C) for the range of carbon

contents in carbon and low-alloy steels. However, under non-equilibrium conditions

other phases may form such as bainite and martensite. Each of the aforementioned

phases has different mechanical properties as a result of the differences in their

microstructure. Furthermore, many of the alloying elements present in steels will

participate in precipitation reactions which will also influence mechanical

properties of a given alloy.

Llewellyn et. al [10]has reported that there are five practical options for

increasing the strength of low-carbon steels:

1. Solid solution strengthening

2. Refining the grain size

3. Precipitation strengthening

6
4. Transformation strengthening

5. Dislocation strengthening

Carbon and low-alloy steels derive their mechanical properties based on any

combination of these strengthening methods. A detailed discussion of each of these

strengthening methods is beyond the scope of this review, for detailed information

the reader is referred to the references cited throughout this section. The following

is intended as a basic introduction to steel metallurgy as it relates to the steels of

interest mentioned above.

7
Figure 2.1 Fe-C phase diagram (region of interest for steels)

The metallurgy of carbon and low-alloy steels varies tremendously based on

the intended application. Carbon steels typically exhibit microstructures consisting

of ferrite and pearlite in varying fractions dependent on carbon content and


8
processing. These steels are typically strengthened through a combination of solid

solution grain size refinement, strengthening, and precipitation strengthening.

Llewellyn et. al [11] has summarized details of these strengthening mechanisms in

steels. Generally speaking ferrite is the softest and therefore weakest of the

aforementioned phases present in carbon steels. The addition of alloying elements

to iron increases the strength of ferrite by solid solution strengthening, with carbon

being potentially potent in this regard [9]. Reducing the ferrite grain size also has a

beneficial effect on grain size, with the particular advantage that toughness is not

sacrificed. Ferrite grain size can be controlled in a number of ways. Alloying

elements that form a fine dispersion of particles can be added such that at heat

treatment above the A3 temperature austenite grains growth is limited by pinning of

grain boundaries by these particles. Fine austenite grain size will result in fine

ferrite grain on cooling to room temperature. Decreasing the temperature at which

the austenite to ferrite transition occurs can also be utilized to control grain size.

This can be accomplished by alloying additions of carbon and manganese, or

increased cooling rates. Precipitation strengthening is typically achieved through

additions of alloying elements that undergo precipitation reactions either on cooling

or during subsequent heat treatment, notably niobium and vanadium in carbon

steels. Final microstructure in carbon steels will consist of a mixture of ferrite and

pearlite based on carbon content, with possible presence of precipitates. Controlled

rolling is the preferred processing route as it eliminates the need for heat treatment.

9
HSLA steels are strengthened on the same principals as carbon steels.

Increased alloying in these steels means that precipitation strengthening plays a

larger role than in carbon steels where grain refinement is more significant. HSLA

steels may be used in the as-rolled or normalized condition dependent on alloying

and mechanical property requirements.

The quenched and tempered steels take advantage of transformation

hardening to achieve high strength levels. Transformation hardening is achieved by

the transformation of austenite to the non-equilibrium transformation products

bainite and martensite that are stronger than ferrite and cementite. The increased

amounts of alloying additions in these elements increase the materials

hardenability, the ease of martensite formation. Alloy additions are also made to

these steels to impart specific properties such as low-temperature fracture

toughness or oxidation resistance. Following quenching, which is performed in a

variety of mediums, these steels undergo tempering treatments which restores

fracture toughness at the expense of some strength. Final microstructure in these

alloys will consist of tempered martensite, with or without the presence of second

phases such as precipitates and retained austenite. These steels often require

complex heat treatments, particularly if the presence of precipitates or retained

austenite is necessary for mechanical properties.

HTLA steels have increased levels of most alloying elements compared to

quench and tempered grades. For this reason their metallurgy is closely related to

the quenched and tempered steels. The most important difference is that the

10
increase in alloying additions increases the hardenability making the formation of

martensite easier, which has significant impact on the welding metallurgy of these

steels as will be discussed. Additionally, the higher carbon levels exhibited in these

steels make the martensite formed harder and more brittle. Like quenched and

tempered steels they require complex processing routes to ensure adequate

mechanical properties.

Chromium-molybdenum steels have increased levels of the alloying

additions chromium and molybdenum as their name suggests. The metallurgy of

these steels is similar to the other more highly alloyed classes already discussed.

The addition of chromium to these steels account for their increased oxidation

resistance. Molybdenum increases the resistance to creep at the elevated

temperatures at which these steels are often used by forming precipitates that

increase resistance to grain boundary sliding. Other elements, such as V, Nb, and Cr,

will also participate in precipitation reactions that have similar effects. Complex

heat treatments are generally required for reasons already discussed as well as the

susceptibility of these steels to temper embrittlement [8].

The welding metallurgy of all the above steels is quite complicated as well.

They are welded with a variety of filler materials to suit particular needs, which can

result in a variety of issues related to the welding metallurgy of a given system. The

main concern when welding these steels as will be discussed in more detail below is

the formation of hard brittle martensite within the fusion zone and heat affected

zone (HAZ). This is particularly problematic in the HTLA grades that are highly

11
alloyed resulting in high hardenability. When welded, these grades are likely to form

hard martensite in the HAZ. Loss of mechanical properties can also occur for a

number of other reasons during welding. Grain growth in the HAZ of alloys where

ferrite grain size plays an important strengthening role can reduce base metal

properties. Loss of precipitation strengthening can also occur when precipitates go

back into solution during welding and don’t have a chance to re-precipitate on

cooling. The welding metallurgy can be quite complicated and will vary based on the

specific alloys involved and the processes used to join them.

2.1.3 Weldability

Carbon and low-alloy steels encounter a wide variety of concerns when

joined [8]. Different alloy groups are susceptible to solidification cracking, lamellar

tearing, and reheat cracking. The number of different weldability concerns varies in

severity and understanding. However, the primary weldability issue encountered

when joining carbon and low-alloy steels is cold cracking; also known as hydrogen

embrittlement [8], or hydrogen induced cracking (HIC); and is common to most

steel alloys. This type of cracking is associated with hard, brittle martensite often

encountered in the heat-affected zone (HAZ) of carbon and low-alloy steels. When

this hard brittle microstructure is combined with stress, and hydrogen from many

common welding processes, cracking occurs. Although the details of the phenomena

surrounding this type of cracking are disputed and not well understood, general

methods for avoiding it do exist. These focus mainly on eliminating one of the

contributing factors mentioned above. Post weld heat treatment is often applied to
12
welds where brittle microstructures would be expected to form in the HAZ upon

cooling. Furthermore, steel grades such as the HSLA and quenched and tempered

steels, where hardenability has been reduced by lowering the amount of alloying

additions, such that martensite formed on cooling is of a softer and less brittle

nature if formed at all, have been developed with joining in mind. The prevalence of

this form of cracking has led to many ways of mitigating it in the industry.

2.1.4 Applications

Carbon and low-alloy steels find use in the greatest variety of applications of

any engineering alloy. The carbon varieties are popular in construction of a number

of structures such as buildings and bridges [11]. The HSLA grades are used in

applications ranging from ship hulls to automobiles. The quench and tempered

varieties are often used in construction of pressure vessels. The chromium-

molybdenum varieties are tailored for high temperature applications such as

equipment utilized by the power industry [8]. Carbon and low-alloy steels remain

popular due to the wide range of mechanical properties they offer that can be

achieved at low cost.

2.2 Solid-solution strengthened nickel-base alloys

2.2.1 Composition

The solid-solution strengthened nickel-based alloys are so named because

they derive the majority of their strength from alloying additions that act as solid

solution strengtheners. Several classes of these alloys exist based upon the primary
13
alloying additions. Nickel-copper alloys were among the first nickel-based alloys to

be produced and form the basis of the 400 series alloys in use today [12]. Chromium

is often present in these alloys to improve corrosion and oxidation resistance.

Additional systems are based on Ni-Mo, and Ni-Fe with both of these alloying

additions having large maximum solubility in nickel and acting as solid solution

strengtheners. Additions of Co, W, Ta may also be added for their effects as solid-

solution strengtheners. Niobium, along with many of the elements already listed

may also be added for its effect as a carbide former.

2.2.2 Metallurgy and Mechanical Properties

The solid-solution strengthened nickel-base filler metals will always solidify

as austenite. The Ni-Cu series will remain fully austenitic to room temperature due

to the complete solubility that exists between these two elements. Carbon

concentrations over 0.02 wt. % will lead to the formation of Cr rich M23C6 carbide at

temperatures below 950ºC during the cooling cycle. MC type carbides can form

during solidification when carbide-forming elements Ti and Nb are present in

conjunction with carbon. In Ni-Mo and Ni-Cr-Mo alloys the topologically close

packed (TCP) phases σ, μ, and Ρ may form if sufficient Mo and Cr is present. These

phases appear at interdendritic regions and are therefore are associated with a

eutectic-like reaction occurring from liquid enriched with solute by segregation

[12]. The practical aspects of the presence of these phases will be discussed in

subsequent sections.

14
The mechanical properties of solid-solution strengthened nickel-base filler

metals in the as-welded condition are similar to those of their matching base

materials. They often may have superior properties to matching base metals due to

the inclusion of Al, Ti, Mn, Nb, and Mo in the filler metals imparting additional

strength. Many of these filler metals also exhibit reasonable elevated temperature

strength up to 650ºC [12].

Postweld heat treatment (PWHT) may be necessary in the solid-solution

strengthened nickel-base filler metals in order to relieve residual stresses,

homogenize the fusion zone, improve corrosion resistance, or dissolve undesirable

second phases. Postweld heat treatment has been shown to result in precipitation

the undesirable phase δ (Ni3(Nb,Mo,Cr,Fe,Ti)) in alloy 625 when carried out in the

temperature range of 750ºC – 950ºC [13]. Precipitation of the orthorhombic δ

phase, coupled with dissolution of body centered tetragonal γ’’ phase (Ni3Nb), which

precipitates out at the low end of the temperature range and imparts strength in

alloy 625 after PWHT, results in a reduction of all mechanical properties of the

weldments.

2.2.3 Weldability

The solid-solution strengthened nickel-base filler metals can be susceptible

to solidification cracking typical in alloys that solidify as primary austenite. This

type of cracking is associated with the presence of liquid films along solidification

grain boundaries that lead to cracking at stresses below the yield stress. Increased

15
solidification temperature range is closely associated with an increase in

solidification cracking susceptibility. Additionally, segregation known to occur

during solidification can result in the formation of low melting point phases that

increase solidification cracking susceptibility. Sulfur and phosphorus have been

shown to increase susceptibility to solidification cracking in alloy 600 due to their

tendency to expand the solidification temperature range when added to nickel. The

strong tendency of these elements to segregate into the liquid during solidification,

coupled with their tendency to reduce solid-liquid interfacial energy, promoting

wetting of liquid films along grain boundaries, add to their propensity to increase

solidification cracking susceptibility in solid-solution strengthened nickel-base filler

metals. Additions of Mn, Si, Al, and Ti have been shown to combat some of these

effects and reduce the solidification cracking susceptibility [14]. In alloys with

greater concentrations of elements, such as Nb, Cr, and Mo, which participate in

carbide and intermetallic reactions, solidification cracking is more influenced by the

formation of these phases at the end of solidification. These phases tend to form at

lower temperatures than the alloy solidus thus increasing the solidification

temperature range and solidification cracking susceptibility [12]. Studies on the

Hastelloy series of solid-solution strengthened nickel-based filler metals, C-4, C-22,

and C-276 have shown an increased risk of solidification cracking susceptibility due

to the presence of the TCP phases discussed above [15]. Cieslak has reported on the

negative effect of Nb, C, and Si additions on solidification cracking susceptibility in

alloy 625 caused by the formation of laves phase and MC type carbides [16]. It is

16
noteworthy that although alloy 625 often shows poor resistance to solidification

cracking during Varestraint testing, it is actually quite weldable in practice, being

aided by the fact that it often will generate sufficient liquid to backfill and heal

cracks [12]. In practice most of the solid-solution strengthened nickel-based filler

metals are generally quite weldable as a result of either tight control of alloying

additions, small solidification temperature ranges, and the ability to backfill and

heal cracks.

Ductility dip cracking (DDC) is type of cracking associated with a number of

alloys including solid-solution strengthened nickel-base filler materials. It is a solid-

state phenomenon and occurs at temperatures between the solidus (Ts) and 0.5Ts.

Materials susceptible to this form of cracking will experience a dramatic drop in

ductility over this temperature range, and the resultant stresses from welding will

cause cracking. The mechanism for DDC is not well understood, and many theories

exist to try and explain the phenomenon. DDC is found to be present along migrated

grain boundaries (MGBs) in the fusion zone of solid-solution strengthened nickel-

base alloys. Recent studies [17-19] have indicated that DDC is related to grain

boundary sliding that occurs at high temperatures. It has been shown [12, 17-20]

that susceptibility to DDC can be reduced by the formation of intergranular

precipitates that effectively lock MGBs preventing sliding from occurring.

Corrosion of solid-solution strengthened nickel-base alloy weldments can

also be a concern when such alloys are used in situations where corrosion

17
resistance is important. This can be attributed to the tendency for elemental

segregation in the weld metal leading to interdendritic secondary phases and

composition gradients which can lead to localized attack [12]. PWHT can have both

positive and negative effects on weldments corrosion depending on the temperature

and time it is carried out at. When stress corrosion cracking is a concern the use of

PWHT to relieve residual stress can be beneficial. Cortial et al. [13] has reported an

increased susceptibility to corrosion when alloy 625 is heated in the range

discussed earlier that results in formation of γ’’ and δ phases. PWHT of these alloys

must therefore be carried out carefully to insure the best mix of properties.

2.2.4 Applications

Solid-solution nickel-based base metals are used in a variety of applications

and their matching filler metals would therefore often be expected to be present in

many of the same applications where joining of these materials would be necessary.

These include use in the power generation, chemical processing and petrochemical

industries [12]. Use in these industries is often closely related to their superior

corrosion resistance at high temperature [21]. They are also used as cladding

material for variety of carbon, low-alloy, and stainless steels [8, 12, 21].

Furthermore, they find extensive use in dissimilar metal combinations as discussed

in detail in section 2.4.

18
2.3 Weld Metal Constitution Diagrams

Diagrams that allow the prediction of a material’s microstructure based upon

its chemical composition, known as constitution diagrams, have been in existence

for almost a century. The majority of the research into constitution diagram

development has been focused on stainless steels that exhibit a final microstructure

consisting of some mixture of δ-ferrite and austenite, and to a lesser degree

martensite. The carbon and low-alloy steels and nickel-base alloys will be shown to

have microstructures consisting of some mixture of martensite and austenite when

welded. For this reason, the focus of this section will be to present background on

those constitution diagrams that have been developed which are capable of

predicting these microconstituents. Much of the focus of early diagrams was on

austenitic and duplex austenitic plus ferritic systems. Many of these diagrams still

included regions that predicted martensitic microstructures. Furthermore, these

diagrams were the building blocks for modern constitution diagrams used to predict

martensitic, austenitic, and martensitic plus austenitic microstructures. A brief

history of their development is therefore essential to an understanding of the

constitution diagrams of concern.

Due to the significance of constitution diagrams to the welding community,

as well as their prolonged history, a number of comprehensive reviews on the

subject have been undertaken. Olson [22] has compiled a review of constitution

diagram history up to 1985. While reviewing many of the same studies as Olson,

Balmforth [23], conducted a review of literature on the subject up to 1998. Lippold

19
et al. [5], have also presented a thorough review of the subject. Those authors have

presented much of the following previously.

2.3.1 Early Constitution Diagrams

Strauss and Maurer [24] produced the first of these diagrams in 1920. Their

diagram allowed prediction of the following phases: austenite, martensite,

troostosorbite (tempered martensite and bainite), and pearlite in slowly cooled

wrought steels. Nickel and chromium percentages were the y-axis and x-axis

respectively. A modified version of their diagram was published in 1939 by Scherer,

Riedrich and Hoch [25]and is presented in Figure 2.2. Their diagram included the

addition of austenite-ferrite stability lines. This diagram was applicable for alloys

with 0 to 26 weight percent chromium and 0 to 25 weight percent nickel. Within

normal contents of carbon, silicon, and manganese this diagram was useful in

predicting microstructure based on composition.

20
Figure 2.2: Modified Strauss-Maurer Diagram [25]

In 1938, Newell and Fleischman [26] published an equation that predicted

the austenite-austenite plus ferrite boundary on the basis of alloying additions.

Their equation is as follows:

(Cr  2Mo  16) 2 Mn


Ni    30(0.10  C )  8
12 2

The chemical symbols in the equation above indicate weight percentage of

the given element, as they will in all equations that follow. The equation published

by Newell and Fleischman would become the standard for constitution diagram

development up to present time. Future investigators would separate the elements

that promoted austenite and ferrite onto opposing sides of their equations. These

groupings were the predecessors to the chromium and nickel equivalent formulas

used in modern day constitution diagrams. Much of the research into constitution

21
diagrams has been in altering of the nickel and chromium equivalents to ensure

accurate prediction of microstructure.

In 1943, Field, Bloom, and Linnert [27] modified the Newell-Fleischman

equation to improve its accuracy when applied to welding. The constant 8 was

changed to 11 to reflect the effect of higher cooling rates experienced during

welding as compared to wrought products, and its influence on martensite and

ferrite formation. Presented with austenite forming elements on the right and

ferrite forming elements on the left, as it would eventually be written, their equation

for the austenite – austenite plus ferrite boundary was:

(Cr  2Mo  16) 2


Ni  0.5Mn  30C   14
12

The work of Campbell and Thomas [28] on 25 chromium 20 nickel weld

metals found that microstructure and mechanical properties could be correlated to

small additions of molybdenum and niobium. They developed the following

chromium equivalent to demonstrate this:

Creq  Cr  1.5Mo  2 Nb

As previously mentioned, the use of equivalency formulae such as these would be

the trend for constitution diagram development to present day.

Thomas [29] reported the following equation for the stability of austenite

relative to δ-ferrite:

Ni  0.5Mn  30C  1.1(Cr  Mo  1.5Si  0.5Nb)  8.2

22
This equation was significant for two reasons. It included terms for silicon and

niobium. More importantly, it reflects a linear boundary versus the quadratic

boundary featured on the previously discussed diagrams. This trend would become

predominant in future constitution diagram development.

2.3.2 The Schaeffler Diagram

Schaeffler [3] realized the potential application of the previously discussed

research to be applied practically to welding applications. Schaeffler sought out to

combine the concepts of equivalency formulae with the Maurer diagram. The

Maurer diagram was used in his initial study; however, he modified it to include

nickel and chromium equivalents as the y-axis and x-axis respectively. His

chromium equivalent was chosen to be:

Creq  Cr  2.5Si  1.8Mo  2 Nb

and his nickel equivalent was:

Nieq  Ni  0.5Mn  30C

Multiplying factors for chromium, nickel, manganese, and carbon were chosen based

on those published in much of the research discussed above. He chose multiplying

factors for silicon, molybdenum, and niobium based on his own experience.

Balmforth [23] notes that the absence of a nitrogen term in Schaeffler’s nickel

equivalency formula, despite its strong effect as an austenite stabilizer, was likely

due to limitations in determining nitrogen contents of steels at the time.

23
Schaeffler’s original diagram was created by plotting data points from 34

different samples created using various combinations of austenitic stainless steel

filler material on 4340 steel and mild steel plate, as well as undiluted weld metal on

the modified Maurer diagram. Using the chromium and nickel equivalents he

determined, data points were determined based off of tested chemical compositions

of his samples. Metallographic analysis was performed on each of the sample to

determine microstructural phases present and relative quantities. The austenite –

austenite plus ferrite and austenite – austenite +martensite boundaries were

redrawn to coincide with the data points. This diagram is presented Figure 2.3.

Included in Figure 2.3 are the data points from Schaeffler’s original work, as well as

lines corresponding to the modified Maurer diagram illustrating how it was adapted

by Schaeffler.

24
Figure 2.3: The Schaeffler Diagram (1947) [3]

Additionally, Schaeffler reported an equation for the austenite – austenite +

ferrite boundary in his original diagram:

(Creq  16) 2
Nieq   12
12

where Nieq and Creq are his nickel and chromium equivalents as reported above. He

notes that his equation only differs from those proposed by Newell and Fleischman,

and Field, Bloom and Linnert by the final constant term.

Schaeffler published modified versions of his diagram in 1948 [30], and again

in 1949 [4]. The 1948 version of his diagram featured straight lines for the

boundaries and included iso-ferrite lines allowing for better prediction of weld-

metal microstructure to be accomplished within the two-phase austenite plus ferrite


25
region. The final version published in 1949 has little differences from his 1948

version. However, the adjustment of the chromium equivalent in his final version is

noteworthy. Schaeffler [4] notes that examination of additional welded samples

resulted in revision of the multiply factors for silicon, niobium, and molybdenum, as

well as a slight relocation of the phase boundaries. His final chromium equivalent

was:

Creq  Cr  1.5Si  Mo  0.5Nb

Schaeffler’s final diagram is presented in Figure 2.4. This diagram was reported to

be accurate to within 4% ferrite when used for prediction of microstructure in 300

series weld materials. The left side of the diagram was determined to be

quantitatively accurate for prediction of martensite and martensite plus ferrite

microstructures in straight chromium and chromium-molybdenum stainless steels.

The final version of the Schaeffler diagram still finds use today as will be discussed

in subsequent sections.

26
Figure 2.4: The Schaeffler Diagram (1949)

2.3.3 Austenitic-Martensitic Constitution Diagrams

Following the Schaeffler diagram, constitution diagram development began

to focus on specific regions of interest within earlier diagrams [23]. The reader is

referred to the previously mentioned reviews [5, 22, 23] for details in development

of constitution diagrams for prediction of austenitic plus ferritic weld metal

microstructure as well as those concerned with ferritic plus martensitic weld metal

microstructure. The remainder of discussion on constitution diagrams will be

focused on those that predict weld metal microstructures consisting austenite plus

martensite.

27
Balmforth [23] reports that Eichelman and Hull [31] were the first to study

the effect of alloying additions on martensite start temperature (Ms). In alloys

containing 10 to 18 wt% chromium they determined Ms could be approximated by:

M s  75(14.6  Cr )  110(8.9  Ni )  60(1.33  Mn)


 50(0.47  Si)  3000(0.068  [C  N ])

When the Ms is assumed set to be 20ºC this equation is shown to reduce to:

0  38.55 1.25Cr 1.83Ni  Mn  0.83Si  50(C  N )

This equation demonstrates a manganese to nickel ratio close to the 0.5 reported in

Schaeffler’s nickel equivalent. Andrews [32] study of Fe-Mn-Ni weld metal with

chromium contents below 5 wt% reported the following Ms equation:

M s  539  423C  30.4Mn  17.7 Ni  12.1Cr  7.5Mo

When a Ms of 20ºC is assumed this equation becomes:

0  17.07 13.9C  Mn  0.58Ni  0.4Cr  0.25Mo

The manganese to nickel ratio illustrated here is almost the opposite of that

predicted by Schaeffler.

Self et al. [33], noting the Schaeffler diagram’s ability to qualitatively

determine austenitic plus martensitic microstructures based on composition, as

well as the work of Andrews, and of Eichelman and Hull, conducted a study to

determine the correlation between chromium, nickel and manganese contents on

the position of the austenite – austenite plus martensite on the Schaeffler diagram.

Several welds were made that resulted in nominal 9 wt% Cr and 1.2 wt% Cr

concentrations with varying levels of Mn and Ni. It was shown that the equation of

28
Eichelman and Hull correlated well to the 1.2% Cr data generated, while the

equation of Andrews produced good correlation with the 9% Cr data. On this basis

an equation was generated which predicts austenite formation in weld metal with 0-

16 wt% Cr:

14.00  Mn  (0.0833Cr  0.5) Ni  0.0742Cr 2  1.2Cr

Two significant features of this equation were noted. First, the interaction of

chromium with nickel in producing austenitic microstructures was noted as the

reason increased levels of chromium increased the propensity of nickel as an

austenite former as is seen in stainless steels. Second, the nature of chromium to self

interact as demonstrated by the Cr2 term. The author also proposed a modification

to the Schaeffler diagram as presented in Figure 2.5.

Figure 2.5: Diagram of Self et al. [33]

29
Béres [34] has also proposed a modified version of the Schaeffler diagram.

He notes that the Schaeffler diagram was designed using stainless steels with carbon

contents of less than 0.1 wt%. However, the effect of carbon as an austenite

stabilizer is related to its ability to depress the Ms, which is drastically reduced at

levels above 0.1 wt% as are often typical in carbon and low-alloy steels, and could

thusly be expected in dissimilar welds on these materials. Coupling previously

published formulae for Ms as a function of alloying additions Ni, Mn, Cr, Mo, and Si

and Ms variance in unalloyed steels as a function of carbon content he arrives at the

following equivalency formulae:

Creq  Cr  Mo 1.5Si

0.2
Nieq  Ni  0.5Mn  10C 
C

These equations combined with the Ms equation:

MS  MsC  21(Ni  0.5Mn) 16.8(Cr  Mo 1.5Si)

where MsC is the martensite start temperature of unalloyed steel as a function of



carbon content. Between 0.03 wt% carbon and 0.35 wt% carbon MsC is given by:

4.2
M sC  454  210C 
C

By assuming at a Ms temperature of 0°C no martensite will form he adds various


 diagram as a function of carbon concentration. The
boundaries to Schaeffler’s

modified version of the Schaeffler diagram he proposed is presented in Figure 2.6.

Below chromium equivalents of 18, the left side of the diagram is used with the

austenite – austenite plus martensite boundary being determined by the carbon

30
concentration. The use of his modified nickel equivalent, Nieqβ is required in this

region of the diagram. The right side of the diagram is to be used as the original

Schaeffler diagram would, using the original nickel equivalent.

Figure 2.6: Schaeffler Diagram as modified by Béres [34]

2.4 Nickel-Steel Dissimilar Metal Welding

2.4.1 Applications

The solid-solution strengthened nickel-base filler metals are used to join

carbon and low-alloy steels in a variety of applications. When carbon and low-alloy

steels are welded to stainless steels using a stainless filler metal the large difference

in coefficient of thermal expansion (CTE) between alloys (7.5-8.0 µin/in/°F versus

9.5-10 µin/in/°F) of these two groups can lead to cracking in elevated temperature

service (greater than 425ºC). Nickel-alloys possess a CTE of 8 µin/in/°F, that lies

31
between carbon and low-alloy steels and stainless steels. For this reason they are

frequently used to successfully join these alloy groups for use in the power-

generation industry when elevated temperature service is expected [12]. Nickel

alloys are often used to clad carbon and low-alloy steels in application where

corrosion resistance is important.

2.4.2 Weldability

In addition to the aforementioned weldability issues that can accompany

welds made on carbon and low-alloy steels and those made with solid-solution

strengthened nickel-base filler metals, weldability concerns exist that are specific to

combinations of these two alloy groups.

Although the carbon and low-alloy steels have a variety of different final

microstructures based upon composition and processing, they are all similar in that

they will solidify as body-centered cubic (BCC) delta ferrite. The different crystal

structures of the solidification phases of the two alloy groups give rise to an

interesting metallurgical feature in welds between them known as Type II

boundaries, which have been linked with cracking in some instances. The

mechanism by which Type II boundaries form is the result of suppression of

epitaxial growth of the weld metal off of the base metal as is typically observed in

welding. Because of the difference in crystal structure previously mentioned the

austenitic (FCC) filler material is unable to grow epitaxially from the ferritic (BCC)

heat-affected zone of the base metal. The result is heterogeneous nucleation within

the weld metal. As the HAZ of the base metal cools through the δ-ferrite region into
32
the austenite region it becomes FCC and a high-angle boundary is left between the

HAZ and fusion zone. This resultant boundary has high energy and is mobile, and

migrates into the FCC weld metal as a result. As the material cools to room

temperature the boundary is locked into place and the final microstructure is left

with this so called Type II boundary which runs essentially parallel to the fusion

boundary[12]. A schematic illustrating Type II boundaries is shown in Figure 2.7.

Figure 2.7: Schematic representation of Type II boundaries and their formation [35]

The microstructural zone in which Type II boundaries are found within the

fusion zone is commonly referred to as the transition region. Along with Type II

boundaries the transition region is frequently associated with a thin layer of

martensite that forms on cooling as well as after PWHT [36]. The formation of

martensite that often accompanies Type II boundary formation is driven by the

diffusion of carbon from the higher-carbon base material into the weld metal. The

33
Schaeffler diagram discussed previously can be used to predict the presence of

martensite in nickel-base weld metals diluted by carbon and low-alloy steel base

metals as shown in Figure 2.8 [6, 12]. Although the dilution levels necessary to cause

martensite formation (>75%) are rare, a compositional transition across narrow

region (~10μm) will always exist that includes the high dilution levels necessary for

the formation of martensitic structures.

Figure 2.8: Example usage of Schaeffler diagram as it relates to dissimilar metal welding [12]

Numerous studies have been conducted that examine factors that contribute

to the formation of this layer of martensite, and the deleterious effects it has on joint

properties [6, 7, 36]. It has been previously noted that martensite is frequently

associated with hydrogen induced cracking. Rowe et al. [6] investigated the

34
susceptibility to hydrogen induced cracking of welds made with two austenitic

stainless steel filler metals as well as one nickel-base filler metal. They found that

welds made with all three filler metals showed increased cracking susceptibility

when hydrogen was introduced through the shielding gas. Metallographic analysis

showed that the cracking was associated with a layer of martensite that formed

within the transition region that exhibited high hardness relative to the weld metal.

The nickel-base filler metal exhibited the highest resistance to hydrogen

induced cracking of the three filler metals tested. This was attributed to the

increased dilution necessary to form martensite of the nickel-base filler metal

compared to the two austenitic stainless steel filler metals examined, which resulted

in a lower overall amount of martensite present at the fusion boundary.

Additionally, cracking could not be induced in multi-pass welds made using the

same materials. This was attributed to the increased ability of hydrogen to diffuse

away from the martensite given the higher temperatures associated with multi-pass

welding. The importance of filler metal selection in reducing the susceptibility of

dissimilar welds to hydrogen induced cracking by keeping the amount of martensite

present in the transition region to a minimum is apparent from this study; as well as

the confirmation of the benefits of traditional techniques for mitigating hydrogen

induced cracking.

Dupont and Kusko [7] also investigated the formation of martensite along the

fusion boundary of a carbon steel clad with nickel-base and austenitic stainless steel

filler metals. Their study matched the findings of Rowe et al., noting that nickel-base

35
filler materials resulted in the formation of less martensite within the transition

region. They used concentration gradients measured from the base material into an

undiluted region of the filler metal to determine the martensite start temperature

(Ms) as a function of location. Because of the significant effect nickel has in

suppressing the Ms temperature the distance into the fusion zone the Ms

temperature was calculated to be above room temperature was ~10μm for the

nickel-base filler metal compared to ~40μm for the austenitic stainless steel filler

metal. As a result, martensite is able to form a greater distance into the cladding

material when the austenitic stainless steel filler metal was used. The agreement

with the work of Rowe et al. can be attributed to the fact that as the necessary

dilution to form martensite was increased by the use of nickel-base filler metals

over stainless steel filler metals, the Ms temperature was being maintained at or

below room temperature up to this dilution as is shown in the work of Dupont and

Kusko.

Gittos and Gooch [36] also confirmed the presence of martensite near the

fusion boundary in nickel-base and austenitic stainless steel claddings on low-alloy

steel, with an increased quantity of martensite present in the austenitic stainless

steel cladding. They also discussed the effect of PWHT on the transition region. The

composition gradient that exists following welding results in suppression of the Ac1

temperature in the transition region and the PWHT temperatures in the range of

650-720 ºC overshoot the Ac1 resulting in the formation of fresh austenite. PWHT

causes carbon diffusion from the base material into the transition region and weld

36
material to occur due to the difference in carbon concentration between the base

and filler materials. The results of this are formation of Cr-rich carbides within the

transition region and weld metal. The carbides themselves contribute to hardening

in this region and also tie up Cr resulting in a lower level in the matrix that increases

the Ms temperature and promotes formation of fresh martensite on cooling. The

result is that the PWHT has a limited effect on increasing ductility in the transition

region. However, the larger problem associated with PWHT of dissimilar metal

welds such as this is revealed to be how carbon depletion influences HAZ

microstructure as will be discussed below.

Carbon migration from carbon and low-alloy steel base materials into the

weld metal in dissimilar metal welds is associated with the formation of a soft

ferritic microstructure in the HAZ [12, 36, 37]. Gittos and Gooch [36] noted this

microstructure following PWHT of samples created by cladding low-alloy steel with

nickel-base and austenitic stainless steel filler metals, and that it had a lowered

hardness compared to the as-welded HAZ and was the location of many failures

during mechanical testing. However, in practice, this type of failure is often

associated with creep at high temperature exacerbated by large CTE differences

between materials in the joint. Therefore, these types of failures are more

associated with dissimilar metal welds where an austenitic stainless steel filler

material is used because of the large difference of CTE. This mismatch in CTE results

in high local strain along the interface between the carbon/low-alloy steel and the

austenitic stainless steel filler metal. Over time at high temperature, grain boundary

37
sliding occurs in the soft ferritic region resulting in creep failures. Nickel-base filler

metals are often the solution for this type of problem. Nickel-base alloys possess a

CTE that closely matches carbon/low-alloy steels, and are also stronger than the

carbon denuded ferritic HAZ. As a result they will transfer the strain caused by

mismatches in CTE between different materials onto the austenitic stainless steel

HAZ which can accommodate them better [12].

Solidification cracking issues common to nickel-base filler materials can be

increased when they are used to weld carbon and low-alloy steels. In these

combinations the susceptibility to solidification cracking is increased by dilution of

the filler metal by the base material. As previously noted, the solidification

temperature range has a direct correlation to the solidification cracking

susceptibility of a given alloy. Increased levels of Fe, Cr, Cu, Si in nickel-base weld

metal results in increased solidification temperature ranges. Additionally, carbon

and low-alloy steels high levels of impurities such as S and P will result in locally

increased solidification cracking susceptibility [12]. Insuring minimal dilution is

therefore key to minimizing cracking susceptibility in dissimilar metal welds

between carbon and low-alloy steels and nickel-base alloys.

38
3. OBJECTIVES

CHAPTER 3

OBJECTIVES

The objective of this study was to evaluate the ability of the Schaeffler

diagram to predict weld metal microstructure for when carbon and low-alloy steels

are joined with solid-solution nickel-base filler materials. The data compiled

throughout this process will be used to accomplish the main goal of development of

a new weld metal constitution diagram aimed specifically at welds between these

two materials from which samples in this study were created. The specific

objectives of this study were:

1. Using previously defined techniques; develop samples whose

composition would cover a wide range of the area within the

Schaeffler diagram that predicts microstructures consisting of

austenite, martensite or a mix of the two.

2. Determine quantitatively the microstructural makeup of previously

created samples using a variety of analytical techniques, such that the

accuracy of the Schaeffler diagram for prediction of microstructure in

these materials can be determined.


39
3. Using statistical methods, develop new nickel and chromium

equivalency formulae that are more accurate when applied to these

materials.

4. Combined with new equivalency formulae, develop a new constitution

diagram with boundaries between austenitic, austenitic plus

martensitic, and martensitic regions redrawn for greater accuracy in

predicting microstructure of solidified material with compositions

falling between carbon and low-alloy steels and nickel-base alloys.

40
4. EXPERIMENTAL PROCEDURES

CHAPTER 4

EXPERIMENTAL PROCEDURES

In order to meet the objectives of the study a range of data was required that

covered the majority of the martensite plus austenite region of the Schaeffler

diagram, as well as that area immediately surrounding it. Much of the procedure

that follows has been adapted from the work of Balmforth [23], whose work

provided a good starting point for the following study. This section contains the

procedures that were followed during this study.

4.1 Materials

Carbon and low-alloy steels and nickel-base alloys for this study were

selected on the basis of their position when plotted on the Schaeffler diagram, their

use in the industry and their availability. The materials used in this study are

presented along with their composition in Table 4.1. The selection of these materials

covered a wide range of the Schaeffler diagram as shown in Figure 4.1. Material

combinations for sample production were selected that allowed the easiest drawing

of tie-lines to be accomplished. This generally resulted in material combinations

whose connecting ties lines were roughly parallel to each other. Initial material

41
combinations are illustrated in Figure 4.2, which demonstrates how tie-lines were

drawn between intended material combinations. Three dilutions were selected

within each material combination for sample production. One dilution was selected

that would fall in the middle of the martensite plus austenite region of the Schaeffler

diagram. The two other dilutions were selected such that they fell just within the

austenite region on the other side of the austenite – austenite plus martensite

boundary, and just within the martensite region on the other side of the martensite

– martensite plus austenite boundary. The initial dilutions selected are illustrated

graphically in Figure 4.2. Additional dilutions were later melted to provide

additional data points.

42
Table 4.1: Chemical composition of materials used in this study

Material

Mn

Mo

N*
Nb

Cu

Co
Cr
Ni

W
Al
Si

Ti

V
P
C

S
8630 0.299 0.856 0.009 0.008 0.257 0.770 0.782 0.363 0.002 0.004 - 0.035 0.009 0.198 - 0.008

A553 0.044 0.582 0.004 0.001 0.189 8.970 0.140 0.010 0.001 0.004 0.004 0.018 0.001 0.034 0.050 0.012
Steels

A508 0.182 1.449 0.005 0.002 0.181 0.920 0.125 0.514 0.002 0.002 - 0.016 0.004 0.040 - 0.025

X80 0.023 1.370 0.014 0.004 0.113 0.013 0.040 0.010 0.059 0.042 - 0.031 0.005 0.033 - 0.003

F22 0.141 0.430 0.011 0.022 0.292 0.128 2.420 1.030 0.004 0.002 - 0.025 0.017 0.190 - 0.010

P92 0.120 0.430 0.010 0.001 0.280 0.120 8.270 0.380 0.010 0.010 0.045 0.006 0.174 0.190 1.615 -

625 0.017 0.037 0.003 0.001 0.084 64.40 22.29 8.680 3.660 0.212 - 0.102 - 0.014 - -

FM82
Ni-Base

0.040 2.810 0.004 0.001 0.030 70.40 21.00 0.005 2.210 0.050 - - - 0.090 - 0.050
Alloys

FM52M 0.020 0.760 0.003 0.001 0.110 61.37 29.06 0.040 0.820 0.210 - 0.050 - 0.020 - -

C276 0.003 0.330 0.007 0.006 0.300 57.75 16.06 16.01 0.179 0.010 0.014 0.030 0.020 0.060 3.350 0.200

800H 0.068 0.710 0.006 0.001 0.380 30.90 20.90 0.250 0.040 0.330 - 0.280 - - - -
*Note: Omission of value does not reflect lack of presence in given alloy

43
Figure 4.1: Schaeffler diagram with materials of interest plotted

44
Figure 4.2: Schaeffler diagram enlarged to show study focus area

45
4.2 Sample Production

Balmforth [23] demonstrated the effectiveness of the button technique in

creating weld metal microstructures. For this reason the button melting technique

was utilized as the means of sample production for this study. A schematic of the

button melting setup is presented in Figure 4.3. The intended final sample size was

four grams. Masses of base materials used to create the samples were determined

by calculation using the intended dilution and intended final sample mass. The

common convention for dilution is weld metal dilution by the base metal was used.

Masses of base materials for use in creation of button melt samples were

determined by calculation based on intended final mass of 4g and intended dilution.

Because base materials were available in varying forms (pipe, sheet, welding wire,

etc.) they often needed to be sectioned to the necessary mass. This was usually

accomplished by sectioning with an abrasive saw and wet grinding on a belt sander.

Base materials were cleaned in ethyl alcohol and dried following this procedure.

They were then massed on a high precision digital scale to within 0.01 gram of the

target mass.

After being massed out, base materials were placed within the water-cooled

copper crucible of the button melting apparatus. The chamber was then closed and

purged with argon. Melting of samples was accomplished with the GTAW torch

using a DCEN polarity and a 2% thoriated tungsten electrode with a nominal current

setting of 150 A. The current was controlled with a foot pedal and was gradually

increased to its maximum as applying excessive current before base materials had

46
been allowed to melt could result in ejection of material. The GTAW torch is swirled

at maximum current, as shown in Figure 4.3 to try to create motion within liquid

button that would eliminate concentration gradients. Current was similarly ramped

down gradually following melting. Samples were flipped upside down and re-melted

twice to insure sample homogeneity. Following solidification, samples were allowed

to cool for five minutes in argon atmosphere to prevent oxidation.

47
Figure 4.3: Schematic representation of button melting setup

4.3 Characterization

4.3.1 Sample Prep

Following button melting, samples were sectioned in half using a LECO VC-

50 precision diamond saw. Samples intended for metallographic analysis were

mounted in Bakelite. Samples were polished to final polish using two methods. The

automated method utilized a LECO Spectrum System 1000 grinder/polisher

48
base unit equipped with semi-automatic head and the use of LECO magnetic

CAMEO disks, and LECO FAS disks. The polishing steps for this procedure were a

grinding until planar using the blue CAMEO disk, followed by use of the silver

CAMEO disk with 6μm diamond suspension, followed by use of 3μm diamond

paste on a adhesive backed polishing cloth attached to FAS disk, before finishing

with colloidal silica followed by water to flush the pad and remove residue from

samples. Wheel speed of 250 RPM was used for all steps but colloidal silica when the

speed was reduced to 150 RPM. Cylinder head pressure varied based on application.

In general pressure was about 8-9 psi/sample when 1.5in samples were used, and

about 5-6 psi/sample when 1.25in samples were used. These levels were generally

reduced 1-3 psi/sample during colloidal silica step. Time at each step was also

varied; the author suggests 3 min on each step following rough grinding except for

final polishing with colloidal silica, where 1.5 min polish and 1.5 min water flush

should be used initially. Samples were rinsed in soapy water followed by ethyl

alcohol between each step listed. Manual polishing was performed on Leco

Spectrum System 1000 grinder/polisher base unit. Following mounting, samples

were wet ground to 800 grit, followed by polishing with 3μm, then 1μm diamond

paste. Samples were rinsed in cold water, followed by ethyl alcohol, and then

ultrasonically cleaned between each step. Samples intended for ferrite number (FN)

measurement x-ray diffraction (XRD) were not mounted if metallographic analysis

was unnecessary. They were simply sectioned following procedure outlined above,

49
ground to 800 grit and rinsed in water, followed by ethyl alcohol and dried in hot

air.

A number of different etchants were used to reveal the two-phase

microstructure of melted buttons. These etchants and procedure for their use are

listed in Table 4.2. Additionally, the use of ferro-fluid, a colloidal suspension of Fe3O4

particles was used to differentiate between austenite and martensite. The Fe3O4

particles are attracted to the ferromagnetic martensite resulting in staining of this

phase. A single drop of ferro-fluid is placed on the surface of a mounted sample; the

sample is then rinsed in a bath of petroleum ether and agitated to remove the ferro-

fluid. The sample is then placed in another bath of petroleum ether and agitated

again to remove any remaining residue.

50
Table 4.2: List of etchants used in this study

Etchant ID Etchant Time


1 5% Nital 15-45 sec.
2 5% Nital 2-5 sec.
followed by: +
10 g Na2S205 + 100 mL H2O 15-45 sec.
3 4% Picric Acid 15 sec.
4 5 %Chromic Acid
(5 g CrO3 + 100 mL Ethanol)
30-120 sec.
Electrolytic at:
3 V ; 0.5 A
5 10 mL (4g Picral + 100 mL
H20 Solution) + 3-4 drops
HCl (Pt. 1)
20-40 sec.
followed by:
10 g Na2S205 + 100 mL H2O
(Pt. 2)

4.3.2 Optical Microscopy

An Olympus GX51 light microscope was used to examine microstructures of

button melt samples and take photomicrographs. Photomicrographs were shot at

various magnifications which best allowed inspection of solidification

microstructure as well as individual phase detail. Samples were primarily examined

without the aid of a polarizer or differential image contrast filter. Samples were

examined following etching, application of ferro-fluid, and in the as polished

condition to suit various needs.

4.3.3 Optical Phase Fraction Analysis

Quantitative measurement of the phase fraction of martensite and austenite

contents of button melted samples was determined using two optical analysis
51
methods. The first involved use of image analysis software ImageJ [38]to selectively

threshold micrographs of button melted samples. Micrographs were taken in an

etched or ferro-fluid stained condition that resulted in maximum contrast between

austenite and martensite present. The image analysis software was then used to

convert micrographs to black and white images, fine tune brightness and contrast to

distinguish phases, selectively threshold a given phase based on black or white level,

and determine amount of threshold area based on pixel count relative to total pixels.

Point counting was also used as a means of determining austenite and martensite

phase fraction. Point counting was performed in accordance with ASTM E562. Point

counting was conducted on micrographs taken at 200X in the as-polished condition.

ImageJ was used superimpose a grid onto micrographs, as well as tune contrast as

necessary to give greater phase differentiation. The grid was chosen such that on

each micrograph three regions of 100 points existed for analysis. For each sample

analyzed 10 micrographs were chosen resulting in a total of 30 fields analyzed per

sample. Phase fraction was determined for each sample based on the standard.

Micrographs used for phase fraction analyses were taken in the region of button

melted sample where dendritic solidification was observed as this was the most

representative of a corresponding weld metal microstructure.

4.4 Ferrite Number Measurement

Ferrite number (FN) measurement was chosen as a means of characterizing

the martensite content of the button melt samples. Despite the fact that FN is

generally a means of characterizing the amount of ferrite in austenitic and duplex


52
stainless steels, measurement by magnetic response has been used by previous

investigators [39] to quantify the amount of martensite present in weld metals. The

Magne-Gage correlates a measured force generated by a spring necessary to pull a

magnet from a sample to a dial reading that can be converted into a FN. The FN is

correlated to the dial reading by a calibration as described in ANSI/AWS A4.2-91.

This calibration results in a plot of FN versus dial readings. The calibration involves

taking dial readings from several calibration test specimens that have an established

FN. Plotting of these readings versus the FN established for each calibration

specimens resulted in a several data points which a line can be fit through. This

calibration was carried out prior to use of the Magne-Gage for this study and is

presented in Figure 4.4.

Because the Magne-Gage was designed with the intention of identifying small

amounts of ferrite in austenitic weld metal the spring can only provide enough force

by itself to free the magnet from samples with a small overall percentage of ferro-

magnetic constituent within the microstructure. In order to compensate for this

counterweights were utilized which provided additional detachment force. This

method has been demonstrated by Kotecki [40].Calibration curves are shown in

Figure 4.4 for all counterweights used as part of this study. As can be seen, use of

counterweights allows FN as high as 160 to be determined. It should be noted that

because calibration blocks only existed up to 76 FN, actual samples were used with

approximated slopes based on actual calibration data to generate curves for

calibration weights of 20.5g, 27g, and 35g. FN number measurements were made for

53
all button melt samples. Measurements were made until three successive readings

were obtained that deviated by no more than three dial marks, and the average of

these three readings was taken as the dial reading from which a FN was calculated.

Figure 4.4: Calibration curves used to calculate FN using Magne-Gage

4.5 X-ray Diffraction

X-ray diffraction (XRD) was performed on selected samples to attempt to

determine a relationship between measured FN and martensite phase fraction as

determined by XRD. X-ray diffraction (XRD) studies were performed using a Scintag

XDS2000 diffractometer. The system has a Cu anode and 2kW sealed tube X-ray

source. The configuration was the theta-theta type goniometer. The X-ray detector

54
was an i-Ge solid-state type. Accelerating voltage and operating current were 45 kV

and 20 mA respectively.

Measurements were made over 2θ values from 37º to 87º at a rate of one

degree per minute. These values were chosen because they allowed for

measurement of four distinct peaks, the austenite (200) and (220) peaks, and the

martensite (002)-(200) and (112-211) doublets. Martensite phase fraction was

calculated based on the direct comparison method [41]. This method relates

martensite phase fraction (cα) to austenite phase fraction (cγ) by the equation:

I R  c

I R  c

Where Iγ and Iα are the integrated intensities of a chosen peaks of austenite and
 values are calculated by first taking the total number
martensite respectively. These

of counts in a given peak from a 2θ value safely on one side of that peak to a 2θ

value safely on the other side. Following that, the counts per second value

associated with the starting and finishing 2θ values are each multiplied by half the

number of second associated with scanning from one value of 2θ to the other. These

numbers are then added together and represent an estimate of the total number of

counts in the background underneath the measured spectra. This number is

subtracted from the sum of the counts across a given peak as calculated initially to

give a measure of integrated intensity for a given peak. Rγ and Rα depend on the

Bragg angle, hkl, and the phase in question and can be calculated from data

published in the reference above. Once integrated intensities and R-values have

55
been calculated for crystallographic planes of interest, a value for the ratio of cγ to cα

can be calculated from the above expression. This value can then be used to

calculate martensite and austenite phase fractions by the equation:

c  c 1

Comparisons were made between each of the peaks measured resulting in a total of
 four values was taken as the martensite phase
four values. The average of these

fraction.

4.6 Microhardness Testing

Vickers microhardness indents were made in select samples using a LECO

LM100AT microidentation hardness tester. Indents were placed within selected

phase fields to substantiate metallographic evidence that phases in question were

martensite and austenite. All hardness indents were made with 25g load in order to

minimize phase interaction affecting the result.

4.7 Chemical Analysis

Material compositions from Table 4.1 above were collected from various

sources. Materials for this study were largely obtained from past research

conducted members of the Welding and Joining Metallurgy Group at the Ohio State

University. Where available chemical compositions of base materials were obtained

from the vendor. Other base materials were sent out for chemical analysis.

Worthington Steel and MSI performed chemical analysis of base materials. Chemical

analysis at MSI was performed using a LECO Glow Discharge Spectrometer.

56
Balmforth [23] demonstrated that the composition of buttons could be

calculated accurately based on dilution. This method assumes that the amount of

element X in a sample of dilution D (expressed as a fraction) can be calculated by the

value of X in the steel alloy and nickel-base alloy used to create it by:

X button  D(X steel )  (1 D)(X nickel )

Composition of all melted buttons was calculated using this method.



4.8 Data Analysis

Ferrite number data generated by the Magne-Gage was analyzed using

standard statistical methods. In some instances, FN data was normalized according

to techniques described by Gould et al. [42, 43] in order to achieve a better

regression model fit. Normality of response data was achieved by applying the

following equation to FN values:

 FN  70 
FN Norm  sinh 
 20 

The equation above was developed by assessment of the normality of FN data based

on mean and standard deviation. Observation of the data scatter led to the choice of

hyperbolic sine function as the transformation equation. The constants 70 and 20

were determined by trial and error until the data with highest normality was

achieved. Both original and normalized response data was analyzed in Minitab 15

statistical software package using regression techniques in order to fit a model that

correlated FN to chemical composition. Stepwise regression was first applied to

normalized FN data to determine which elements had significant effect of FN


57
response. Furthermore, two-factor interactions and self-interactions were included

in stepwise regression analysis. Regression models were then generated using

factors determined from stepwise regression analysis. Regression models were

evaluated based on R2 and p-value statistics. Those regression models deemed

suitable were used in development of equivalency formulae used for diagram

development.

4.9 Diagram Development

The constitution diagram for prediction of microstructure in dissimilar metal

welds was developed based on regression models determined previously. The terms

in the regression equation were separated into equivalency formulae based on their

tendency to increase or reduce FN. Those terms that had a positive effect on FN

were grouped into a martensite equivalent due to their tendency to promote

martensite. Terms that reduced FN were grouped into an austenite equivalent, as

they tended to promote austenite. Martensite equivalent was placed on the x-axis

and austenite equivalent on the y-axis. Boundary lines were determined based on

FN expected to give approximately 0, 50, and >95 volume percent martensite, and

plotted on the diagram. Diagrams were tested for relative conformity with FN data

by of calculation of martensite and austenite equivalents for button melt samples,

and plotting on diagram.

58
5. RESULTS AND DISCUSSIONS

CHAPTER 5

RESULTS AND DISCUSSIONS

5.1 Button Melt Microstructures

Initial button melt samples as plotted in Figure 4.2 were prepared using

metallographic procedures outlined in 4.3.1 and examined using light microscopy.

The dilutions selected were intended to result in microstructures that represented

the fully martensitic, fully austenitic, and duplex austenite plus martensite regions

of the diagram. The various etchants, as well as ferro-fluid staining proved an

effective method for distinguishing austenite and martensite present in button

melted samples. Application of Etchant No. 2 from Table 4.2 proved the most

successful at distinguishing martensite from austenite. The results of this and all

other etchants utilized as part of this study varied greatly from sample to sample

based on the different compositions of each button melt sample. Additionally, with

the exception of Etch No. 4, etching response was very non-uniform across the

sample surface. This led to difficulties in resolving microstructural features

consistently. However, general structure of button melt samples is evident from

micrographs taken in polished, etched, and ferro-fluid stained conditions.

Micrographs confirmed the microstructure of button melt samples to consist of


59
martensite, austenite, or some mix of the two. In samples made from molybdenum

containing nickel-base filler metals, such as C-276 and 625, the TCP phases σ and P

may be present in small amounts. This is illustrated in Figure 5.1 where small

amounts of an interdendritic phase consistent with the appearance of a TCP phase

[12] are present within the austenite.

Figure 5.1: TCP phase within interdendritic regions of 8630-625 85% dilution button melt sample (Etch 2)

60
Figure 5.2: : Predominantly austenitic microstructure of F22-800H 55% dilution button melt sample (Etch 4)

Figure 5.2 illustrates a microstructure consisting primarily of austenite, with small

amounts of what appears to be P phase. Figure 5.3 shows the microstructure of a

100% dilution 8630 base metal that would be expected to be fully martensitic. Small

islands of retained austenite are present, as is typical of all highly diluted samples.

Figure 5.4 depicts the typical microstructure of a duplex austenite plus martensite

sample. The martensite is present at dendrite cores appears dark, while the

austenite present within the interdendritic areas appears light. During solidification,

the liquid will become enriched in solute, thereby lowering the Ms temperature

relative to the first to solidify dendrite cores. As the sample cools, those regions that

61
are solute lean, the dendrite cores, whose Ms and Mf temperatures remain above

room temperature transform to martensite. The solute rich regions, liquid which

will transform to austenite upon cooling, whose Ms and Mf temperatures are below

room temperature, will remain austenite. The same type of solidification

segregation phenomena results in the formation of TCP phases at the interdendritic

areas of molybdenum bearing samples. Additional micrographs are presented in

Appendix A.

Figure 5.3: Predominantly martensitic structure with pockets of retained austenite in 100% 8630 button melt

sample (Etch 2)

62
Figure 5.4: Duplex martensite + austenite microstructure of P92-800H 74% button melt sample (Etch 5)

5.2 Microhardness Testing

Microhardness measurements were taken as per methods outlined in Section

4.6. Measurements are presented in Figure 5.5 and Figure 5.6.The results of the

measurements are consistent with the expectation that these samples illustrate two

distinct phases within their microstructure. The relatively high value of 393 HV is

also within the hardness range that would be expected for martensite. Although the

value of 261 may be a little less than expected, it should be noted that the size of the

indents relative to the features investigated means values may be distorted by

interaction of a neighboring dendrite of the opposing phase. Additionally, the trend


63
of increased hardness with increased dilution is anticipated based on both the

increased amount of martensite within the sample, and the increased carbon within

the sample resulting in additional hardening of the martensite.

Figure 5.5: Vickers hardness indents in selected phases of F22-800H 63% dilution button melt sample

64
Figure 5.6: : Vickers hardness indents in selected phases of F22-800H 60% dilution button melt sample

5.3 Button Melt Analysis

The initial intent of metallographic preparation was to yield micrographs that

had sufficient difference in phase contrast that phase fraction quantification could

be achieved by means of the first method discussed in 4.3.3. Achieving the level of

contrast necessary to accomplish this proved difficult in practice. Several different

etchants were tried but they all gave similar results. The lath structure of martensite

led to a significant variation in contrast within this phase. As a result, accurate

quantitative assessment of the austenite and martensite contents by this method

proved unreliable. Ferro-fluid application was attempted as a solution to this

problem. Staining with ferro-fluid resulted in attraction of small globules to regions


65
of austenite as shown in Figure 5.7. This was believed to be a result of the slight

ferro-magnetic nature of nickel. Based on this, and phase fraction results which

were inconsistent with results obtained using the Magne-Gage (as discussed below),

the use of image analysis as the primary method of phase fraction quantification

was abandoned in favor of the Magne-Gage.

Figure 5.7: Ferro-fluid staining of F22-800H 55% button melt sample

5.4 Magne-Gage Ferrite Number (FN) Measurement

The Magne-Gage was selected along with the ferro-fluid as possible

alternative means for quantification of phase fraction in button melted samples.

This project was concerned with creating samples that by prediction of the
66
Schaeffler diagram would result in microstructures that were free of ferrite. This

was a key aspect of the choice of the Magne-Gage for martensite detection. As

previously noted, nickel is slightly ferromagnetic as is evidenced by the ferro-fluid

response on primarily austenitic microstructures as discussed above. The amount of

magnetic response from austenite within button melt samples would be expected to

be negligible when compared to that of martensite. However, if ferrite was also

present in the microstructure of button melt samples then quantification by use of

the Magne-Gage would prove difficult [23]. The additional ferromagnetic influence

of ferrite would make it impossible to determine which constituent was necessary

for the magnetic response of the sample and to what magnitude each was

responsible. No ferrite was detected during microscopy of button melt samples used

as part of this study so this was not believed to be an issue that would preclude the

use of the Magne-Gage.

Analysis using the thermodynamic simulation software JMatPro can also be

used to demonstrate the ferrite free final microstructure of button melt samples.

The composition of sample P92 – 800H 88% dilution has a calculated chromium

content of 9.8 wt% and a calculated nickel content of 3.9 wt%. This represents the

largest chromium to nickel ratio of any button created during this project. The large

chromium to nickel ratio suggests that this sample is the most likely to form delta

ferrite on solidification, as well as retain delta ferrite in the final microstructure.

JMatPro was used to calculate phase fraction versus temperature data for this

composition. The graphical output, as illustrated in Figure 5.8 demonstrates that

67
during equilibrium conditions this sample will solidify as delta ferrite, but will

transform to 100% austenite during cooling. However, segregation that occurs

within the delta ferrite during solidification may enrich it to the point that regions

may be stable to room temperature. JMatPro was used to determine segregation

tendencies within delta ferrite during solidification to assess this possibility. As

shown in Figure 5.9, delta ferrite will become enriched in both chromium and nickel

as solidification proceeds. The last to solidify areas will contain approximately 9.6

wt% chromium and 3.8 wt% nickel. The graphical output from JMatPro of phase

fraction versus temperature data for the enriched composition is presented in

Figure 5.10. It shows that even at this enriched composition, transformation to

100% austenite will still occur. This demonstrates that it is unlikely delta ferrite will

be present within final microstructure of button melt samples.

Figure 5.8: JMatPro graphical output of solidification nature of compositional makeup of P92-800H 88% dilution
68
Figure 5.9: JMatPro graphical output of nickel and chromium segregation within δ ferrite during solidification

Figure 5.10: JMatPro graphical output of solidification nature of enriched δ ferrite in P92-800H 88% dilution

69
The results of the Magne-Gage are presented graphically in Figure 5.11.

Qualitatively the results agree with the prediction provided by the Schaeffler

diagram quite well. As dilution is increased, every material combination showed an

increase in FN.

Another characteristic of the data plotted in Figure 5.11 is the presence of

three regions within each curve. Moving from high to low dilution each curve seems

to exhibit a region of moderate slope, followed by a region of increased slope,

followed by a level of moderate slope where data exists. This can be explained by

the fact that magnetism is affected by chemical composition as well as

microstructural makeup. Kotecki [40] has shown that increasing the amount of iron

in fully ferritic samples will result in an increase in FN, with FN approaching 200

being observed for commercially pure iron. The effect of composition on magnetic

response would be expected to be minimal compared to that of increasing the

amount of a ferromagnetic constituent, such as martensite, within the

microstructure. The data in Figure 5.11 shows this phenomenon.

At low dilution the microstructure of the sample is primarily austenite and

results in little magnetic response. The Magne-Gage will have difficulty resolving

any change in magnetism due to alloy content when the microstructure is primarily

austenite, so the slope of the curve would be expected to be essentially zero until

martensite is present within the sample. Once martensite is encountered within the

sample the FN will raise steadily with increasing dilution as the amount of

martensite present in the sample is increased. As noted previously many of these

70
materials will retain some austenite to room temperature even at 100% base

material. This results in the FN at high level being affected somewhat by small

changes in the amount of retained austenite from sample to sample; but more

consistently by the increasing level of iron present in the sample as the dilution

increases. Once the level of retained austenite reaches these levels (believed to be

around 2-5 vol%) the FN increases to a lesser degree with increasing dilution. More

data points are necessary to confirm this trend as accurate.

Figure 5.11: FN versus dilution of button melt samples determined by Magne-Gage

71
5.5 X-ray Diffraction

X-ray diffraction was performed in order to establish quantitative values of

martensite phase fraction for selected buttons, such that a correlation between

martensite phase fraction and FN could be determined. The results of XRD analysis

are presented in Table 5.1 below, and are superimposed on the previously

presented FN data in Figure 5.12 XRD results from highly diluted samples are in line

with expectations based on FN results. Large variation in results of analysis from

peak to peak were seen in two samples as evidenced by the large standard deviation

observed in analysis of F22-800H 66% dilution and 63% dilution. This could be the

result of due to texture effects present in button melt samples due to preferred

solidification that occurs from button melting apparatus. The result of XRD analysis

that is most concerning to this study is the determination of martensite fractions

between 54.3-76.1 volume percent in samples with FN of around 2. This result has

significant confliction with FN results as determined by the Magne-Gage. XRD

spectra from these samples, X80-C276 73% dilution and F22-800H 60%, are

presented in Figure 5.13 & Figure 5.14 respectively. Examination of the spectra from

these samples reveals that distinct peaks are present for both martensite and

austenite. This confirms the presence of martensite in these samples, despite the

possible inaccuracy of the quantitative result presented in Table 5.1. This result is

confirmed by point counting results and will be discussed in more detail in

preceding sections.

72
Table 5.1: X-ray diffraction and Magne-Gage analysis results for selected samples

Material Dilution FN XRD XRD St. Dev.


Combination %Martensite
A553-Alloy 82 85% 119.6 97.9 1.60%
X80-C276 78% 133.5 97.6 1.47%
X80-C276 74% 65.9 48.8 3.76%
X80-C276 73% 2.3 54.3 3.23%
F22-800H 66% 90.3 79.5 8.38%
F22-800H 63% 54.2 60.7 16.99%
F22-800H 60% 2.4 76.1 3.60%

Figure 5.12: FN versus dilution of button melt samples determined by Magne-Gage with superimposed XRD

analysis values

73
Figure 5.13: X-ray diffraction spectra of X80-C276 73% dilution button melt sample

74
Figure 5.14: X-ray diffraction spectra of F22-800H 60% dilution button melt sample

5.6 Point Count

Point counting was performed on selected samples following the results of

XRD analysis in the hope that the results would correlate better to FN data. Point

counting was performed as described in 4.3.3. The results of point counting on

selected samples are presented in Table 5.2.

75
Table 5.2: Point Counting, X-ray diffraction and Magne-Gage analysis results for selected samples

Material Dilution FN Point Point XRD XRD


Combination Counting Count %Martensite St.
%Martensite St. Dev.
Dev.
A508-FM52M 82.5% 116.5 91.3 2.28% - -
A553-Alloy 82 85% 119.5 - - 97.9 1.60%
A553-Alloy 82 82% 36.1 68.6 4.87% - -
X80-C276 78% 133.5 - - 97.6 1.47%
X80-C276 74% 65.9 80.2 4.57% 48.8 3.76%
X80-C276 73% 2.3 73.2 4.22% 54.3 3.23%
F22-800H 66% 90.3 88.4 3.19% 79.5 8.38%
F22-800H 63% 54.2 85.5 4.09% 60.7 17.0%
F22-800H 60% 2.4 58.8 6.12% 76.1 3.60%

The results of point counting on selected samples confirm data trends

predicted from XRD analysis as shown in Table 5.2. Martensite phase fractions are

slightly greater than those predicted by XRD in most cases. This is a result of heavy

segregation of alloying elements to the last to solidify region at the top of the button

melt samples. Due to this segregation, coupled with reduced cooling rates

experienced in this area, a fully austenitic will be present to some varying extent.

The contribution of this area to the overall phase fraction would be accounted for in

XRD analysis, but not point count analysis, which was confined to the region of

button melt samples where dendritic solidification occurred.

The phase fraction measurements determined by point counting also

confirmed the presence of martensite in samples with minimal ferro-magnetic

response based on Magne-Gage results. The weak magnetic response of alloys

exhibiting seemingly high martensite content may be caused by a number of factors.

The most predominant of these being the weaker ferro-magnetic response of


76
martensite compared to ferrite. Previous studies[39, 44, 45] which used various

techniques, including XRD have demonstrated that magnetic test instruments

intended for detection of ferrite respond less when used on samples consisting of

austentic plus martensitic microstructures. The ferro-magnetic response of

martensite has been reported to be as little as half of that of ferrite.

Additionally, Kotecki [40], has demonstrated the influence of chemical

composition on FN response. The magnetic response of ferrite will be increase with

increasing iron content in fully ferritic microstructures. This is likely a contributing

cause of the reduced ferro-magnetism of martensite. The diffusionless martensitic

transformation prevents the partitioning of iron that occurs during the

transformation to ferrite. Thermodynamic influence on the Curie temperature may

also play a role in the reduced magnetic response of button melt samples. The

addition of alloying elements can be shown to reduce the Curie temperature of pure

iron, which could account for the weakened magnetic response of samples tested as

part of this project.

The sample size used for FN number testing may also play a role in the

reduced magnetic response exhibited by certain button melt samples. The Magne-

Gage was designed for samples whose total mass is approximately 4 grams, about

twice the size of the sample size used for testing. This may result in some lowering

of the FN. The presence of retained austenite between martensite laths may also

contribute to FN response. This would cause the martensite phase fraction reported

by point counting to be higher than the true martensite phase fraction within the

77
sample. The discrepancies reported between the characterization techniques used

are likely a result of contributions from all the above-mentioned factors.

5.7 Development of FN to martensite phase fraction relation

An equation that relates martensite phase fraction to FN was developed using

XRD and point counting data to define an upper bound and the assumption that a

ferrite number of zero corresponds to a sample free of martensite. The upper bound,

where greater than 95% volume fraction martensite is present was predicted to

occur at greater than 135 FN. For samples examined as part of this study, martensite

phase fraction can be estimated from FN based on the equation:

Martensite %  0.7037FN

The equation is represented graphically in Figure 5.15. Ferrite number data from

button melt samples that were 100% steel base material are plotted in the top right

hand corner of this diagram. These buttons would be expected to represent as close

to a fully martensitic microstructure. Data from XRD and point counting analysis are

also plotted on this diagram. As can be seen a line drawn through these points

would intersect the y-axis at a value much greater than zero volume percent

martensite. This is due to the reasons discussed in section 5.6. As a result the

accuracy of this equation is unknown.

78
Figure 5.15: Plot of martensite phase fraction vs. FN for this study

5.8 Data Analysis & Diagram Development

A number of regression trials were run according to the methods outlined in

4.8. The regression model chosen for diagram development yielded the following

formula for the prediction of FN:

FN Norm  30.8  296C  Nb  1.94 Ni  30.9Mo  Nb  0.102 Ni 2


 0.406 Ni  Mo  0.275Ni  Cr  1.90Mn  Cr

This model was chosen based on the highest overall R2 value of those tested

combined with the low average p-values of its terms. Note that this equation reports

79
FN in the normalized form, so use of the equation reported in 4.8 used to normalize

FN data for analysis, is required to back-calculate actual FN from normalized FN

values. This has been done to calculate predicted values of FN based on composition

for button melt samples such that a plot of measured FN versus predicted FN can be

generated as presented in Figure 5.16.

160.0

140.0

120.0

100.0
Measured FN

80.0

60.0

40.0

20.0

0.0
-5.0 15.0 35.0 55.0 75.0 95.0 115.0 135.0 155.0

Predicted FN

Figure 5.16: Measured FN versus FN as predicted by regression equation

Although this equation features mostly two-factor interactions that drive its

response the tendencies of particular elements with regards to FN response remain

the same as models developed using only single terms. Niobium was the only

element with a positive coefficient in regression models developed using only

80
composition data. In the above formula it is seen to have the same effect, albeit in

combination with molybdenum. Additionally, the effect of alloying additions as

demonstrated in the above equation parallel many equations published for

calculation of Ms [5, 7]. The greatest effect appears to be due to carbon

concentration, with manganese, chromium and nickel also playing key roles in

depressing FN response just as they do the Ms temperature. The equation above is in

many ways, just another form of an equation for Ms. Austenite forms when alloying

additions suppress the Ms temperature to the point where austenite is stable to

room temperature. The ratio of austenite to martensite is increased with decreasing

Ms temperature up to the point where the Ms temperature passes room temperature

and fully austenitic structures are formed. The above equation also demonstrates

this behavior. Alloy additions for the most part tend to decrease the FN, just as they

do the Ms temperature, and eventually the FN drops below zero, indicating a fully

austenitic sample.

The above equation was separated into those terms that increased and

decreased FN in order to generate the following martensite and austenite

equivalency formulae:

Aeq  296C  Nb  0.406 Ni  Mo  0.275Ni  Cr  1.90Mn  Cr

M eq  1.94 Ni  0.102 Ni 2  30.9Mo  Nb

The preliminary constitution diagram for prediction of microstructure in

dissimilar metal welds is presented in Figure 5.17. The martensite equivalency

formula above is plotted on the x-axis and the austenite equivalency from above on
81
the y-axis. Borders of the duplex martensite plus austenite region have been created

by calculation based on ferrite numbers of 0, 70, and 135; lines equal to

approximately 0, 50, and <95 volume percent martensite. These numbers were

chosen on the basis of XRD, and point counting analyses’ quantification of

martensite phase fraction, as well as the position these values are found on curves in

Figure 5.12. The diagram is presented again in Figure 5.18 with the addition of FN

data superimposed on it. The vast majority of the data falls into the expected region

suggesting the success of the model at fitting the data.

Figure 5.17: Preliminary constitution diagram for microstructure prediction in dissimilar metal welds

82
Figure 5.18: Preliminary constitution diagram for microstructure prediction in dissimilar metal welds with FN data

plotted by developed equivalency formulae

5.9 Use of the Diagram

The preliminary constitution diagram for use in dissimilar metal welds

represents a starting point for evaluation of microstructure in dissimilar metal

joints. It has been created from limited data that therefore may not accurately

predict microstructure in all steel and nickel base alloy systems. Furthermore,

limitations prevented testing of the diagram on actual weldments. Its ability to

successfully predict microstructure resulting from actual welding situations is

assumed based on previous research conducted with similar techniques.

83
Specific attention should be brought to the gray band that borders the

austenite – austenite plus martensite border. As previously discussed, issues were

experienced determining martensite volume percent at low FN with the Magne-

Gage. This has led to the addition of a gray band to the diagram that represents the

possible variation of the austenite – austenite plus martensite border.

In its present form this diagram should be used as an analytical tool only.

This diagram is not qualified for use in weldments intended for actual service, and

should therefore not be used for this purpose. The use of the diagram as a means of

predicting the presence of martensite which may be expected at the interface of

welds between carbon and low-alloy steels and solid-solution nickel base filler

materials is an example of where this diagram may provide benefits over its

predecessors. It is expected to provide improved prediction of weld metal

microstructures of these alloy groups, from which it was created.

84
6. CONCLUSIONS

CHAPTER 6

CONCLUSIONS

1. A preliminary constitution diagram for prediction of microstructure in

dissimilar metal welds between carbon and low-alloy steels and nickel-base

filler metals has been developed.

2. The diagram is expected to be capable of prediction of martensite and

austenite formation tendencies in regions of dissimilar metal welds between

carbon and low-alloy steels and nickel-base filler materials that have

compositions within the ranges of the data used to create the diagram.

3. The diagram is limited by the fact that it was created from ferrite number

data. Ferrite number is not a quantification of martensite volume percentage

and therefore variation can exist in ferrite number response of two samples

with equal martensite content. Therefore the diagram developed in this

study should not be used as a quantitative indicator or martensite volume

percentage in a given sample.

4. The accuracy of the diagram near the austenite – austenite plus martensite

border is not precise. The ferrite number data generated from Magne-Gage

analysis in this region was found to be inconsistent with regards to


85
martensite prediction. The weaker ferro-magnetic response of martensite

compared to ferrite is thought to be the main factor influencing this

phenomena. As a result of inconsistent data generated in this region a band

has been drawn on the new diagram to compensate for this effect.

86
REFERENCES

1. Kiser, S., Nickel-Alloy Consumable Selection for Severe Service Conditions.


NI, SP, Nickel base alloys, Superalloys; NI, Nickel base alloys, 1990.
69(11): p. 30-35.
2. Kiser, S.D., Pulp and paper industry depends on nickel alloys. Welding
Journal (Miami, Fla), 1997. 76(Compendex): p. 53-57.
3. Schaeffler, A.L., Selection of austenitic electrodes for welding dissimilar
metals. Welding Journal, 1947. 26(10): p. 601-620.
4. Schaeffler, A.L., Constitution diagram for stainless steel weld metal.
Metal Progress, 1949. 56(5): p. 680680.
5. Lippold, J.C., Welding metallurgy and weldability of stainless steels / John
C. Lippold, Damian J. Kotecki, ed. D.J. Kotecki. 2005, Hoboken, NJ :: John
Wiley. xvi, 357 p. :.
6. Rowe, M.D., T.W. Nelson, and J.C. Lippold, Hydrogen-induced cracking
along the fusion boundary of dissimilar metal welds. Welding Journal
(Miami, Fla), 1999. 78(Compendex): p. 31S-37S.
7. Dupont, J.N. and C.S. Kusko, Technical Note: Martensite Formation in
Austenitic/Ferritic Dissimilar Alloy Welds. Welding Journal.Vol.86, 2007.
86(2): p. 51s-54s.
8. William R. Oates, A.M.S., ed. Welding Handbook: Volume 4 - Materials
and Applications - Part 2. 8th ed. Vol. 4. 1998, American Welding
Society: Miami.
9. ASM materials information [electronic resource] : ASM handbooks online,
ed. A.S.M. International. 2002, Materials Park, OH :: ASM International.
Electronic data.
10. Llewellyn, D.T., Steels : metallurgy and applications / D.T. Llewellyn and
R.C. Hudd. 3rd ed ed, ed. R.C. Hudd. 1998, Oxford [England] ; Woburn,
MA :: Butterworth-Heinemann. x, 389 p. :.
11. Llewellyn, D.T. and R.C. Hudd, Steels - Metallurgy and Applications (3rd
Edition), Elsevier.
12. DuPont, J.N., J.C. Lippold, and S.D. Kiser, Welding Metallurgy and
Weldability of Nickel-Base Alloys. 2009, Hoboken, New Jersey: John
Wiley and Sons, Inc.
13. Cortial, F., J. Corrieu, and C. Vernot-Loier, Influence of heat treatments on
microstructure, mechanical properties, and corrosion resistance of weld

87
alloy 625. Metallurgical and Materials Transactions A, 1995. 26(5): p.
1273-1286.
14. Savage, W.F., E.F. Nippes, and G.M. Goodwin, Effect of Minor Elements on
Hot Cracking Tendencies of Inconel 600. Welding Journal.Vol.56, 1977.
56(8): p. 245s-253s.
15. Cieslak, M., T. Headley, and A. Romig, The welding metallurgy of
HASTELLOY alloys C-4, C-22, and C-276. Metallurgical and Materials
Transactions A, 1986. 17(11): p. 2035-2047.
16. Cieslak, M.J., The Welding and Solidification Metallurgy of Alloy 625. NI,
SP, Nickel base alloys, Superalloys, 1991. 70(2): p. 49s-56s.
17. Noecker Ii, F.F. and J.N. Dupont, Metallurgical Investigation into Ductility
Dip Cracking in Ni-Based Alloys: Part II. Welding Journal, 2009. 88(3): p.
62s-77s.
18. Noecker Ii, F.F. and J.N. DuPont, Metallurgical Investigation into Ductility
Dip Cracking in Ni-Based Alloys: Part I. Welding Journal, 2009. 88(1): p.
7S-20S.
19. Ramirez, A.J. and J.C. Lippold, High temperature behavior of Ni-base weld
metal. Materials Science and Engineering: A, 2004. 380(1-2): p. 245-
258.
20. Lippold, J.C. and N.E. Nissley, Ductility-Dip Cracking in High Chromium,
Ni-Base Filler Metals. 2008. p. 409-425.
21. Oates, W.R., ed. Welding Handbook: Volume 3 - Materials and
Applications - Part 1. 8th ed. Vol. 4. 1998, American Welding Society:
Miami.
22. Olson, D.L., PREDICTION OF AUSTENITIC WELD METAL MICROSTRUCTURE
AND PROPERTIES. Welding Journal (Miami, Fla), 1985. 64(Compendex):
p. 281. s-295. s.
23. Balmforth, M.C., -, Development of a ferritic-martensitic stainless steel
constitution diagram. 1998. p. xv, 140 leaves.
24. Strauss, B. and E. Maurer, Die hochlegierten chromnickelstahle als
michtrostende stahle. Kruppsche Monatshefte, 1920. 1(8): p. 129-146.
25. Scherer, R., G. Riedrich, and G. Hoch, Einfluss eines gahaltes an ferrit in
austenitischen chrom-nickel-stahlen auf den kornzerfall. Archiv fur das
Eisenhuttenwesen, 1939. 13: p. 52-57.
26. Newell, H.D. and M. Fleischmann, Hot rolled metal article and method of
making same., U.S. Patent, Editor. 1938.
27. Field, A.L., F.K. Bloom, and G.E. Linnert, Development of armor welding
electrodes: relation to the composition of austenitic (20Cr-10Ni)
electrodes to the physical and ballistic properties of armor weldments.
OSRD. 1943.
28. Campbell, H.C. and J.R.D. Thomas, Effect of alloying elements on tensile
properties of 25-20 weld metal. Welding Journal, 1946. 25(11): p. 760-
768.

88
29. Thomas, J.R.D., A constitution diagram application to stainless weld
metal. Schweizer Archiv fur Angewandte Wissenschaft und Technik,
1949. 1: p. 3-24.
30. Schaeffler, A.L., Welding dissimilar metals with stainless electrodes. Iron
Age, 1948. 162(1): p. 72-79.
31. Eichelman, J.G.H. and F.C. Hull. Effect of composition on temperature of
spontaneous transformation of austenite to martensite in 18-8-type
stainless steel. in American Society for Metals -- Meeting, Oct 20-24 1952.
1952. Cleveland, OH, United States: American Society for Metals.
32. Andrews, K.W., Empirical formulae for calculation of some
transformation temperatures. Iron and Steel Institute -- Journal, 1965.
203(Part 7): p. 721-727.
33. Self, J.A., D.K. Matlock, and D.L. Olson, EVALUATION OF AUSTENITIC Fe-
Mn-Ni WELD METAL FOR DISSIMILAR METAL WELDING. Welding Journal
(Miami, Fla), 1984. 63(Compendex): p. 282. s-288. s.
34. Beres, L., Proposed modification to Schaeffler diagram for chrome
equivalents and carbon for more accurate prediction of martensite
content. Welding Journal (Miami, Fla), 1998. 77(Compendex): p. 273-s-
276-s.
35. Nelson, T.W., J.C. Lippold, and M.J. Mills, Nature and evolution of the
fusion boundary in ferritic-austenitic dissimilar metal welds - Part 2: on-
cooling transformations. Welding Journal (Miami, Fla), 2000.
79(Compendex): p. 267s-277s.
36. GITTOS, M.F. and T.G. GOOCH, The interface below stainless steel and
nickel-alloy claddings. Vol. 71. 1992, Miami, FL, ETATS-UNIS: American
Welding Society.
37. Lundin, C.D., DISSIMILAR METAL WELDS - TRANSITION JOINTS
LITERATURE REVIEW. Welding Journal (Miami, Fla), 1982.
61(Compendex): p. 58. s-63. s.
38. ImageJ: <http://rsbweb.nih.gov/ij/>.
39. Kotecki, D.J., Martensite boundary on the WRC-1992 diagram. Welding
Journal (Miami, Fla), 1999. 78(Compendex): p. 180s-192s.
40. Kotecki, D.J., EXTENSION OF THE WRC FERRITE NUMBER SYSTEM. 1982. V
61(Compendex): p. 352. s-361. s.
41. Cullity, B.D., Elements of x-ray diffraction / B. D. Cullity. 2d ed ed. 1978,
Reading, Mass. :: Addison-Wesley Pub. Co. xii, 555 p. :.
42. Jerry Gould, W.P., James Cruz, An Examination of Electric Servo-Guns for
the Resistance Spot Welding of Complex Stack-Ups, in Sheet Metal
Welding Conference XIV. 2010.
43. Ramasamy, S., J. Gould, and D. Workman, Design-of-experiments study to
examine the effect of polarity on stud welding. Welding Journal (Miami,
Fla), 2002. 81(Compendex): p. 19/S-26/S.
44. Hecker, S.S., et al., EFFECTS OF STRAIN STATE AND STRAIN RATE ON
DEFORMATION-INDUCED TRANSFORMATION IN 304 STAINLESS STEEL - 1.
89
MAGNETIC MEASUREMENTS AND MECHANICAL BEHAVIOR. Metallurgical
transactions. A, Physical metallurgy and materials science, 1982. 13
A(Compendex): p. 619-626.
45. Talonen, J., P. Aspegren, and H. Hanninen, Comparison of different
methods for measuring strain induced -martensite content in austenitic
steels. Materials Science and Technology, 2004. 20(Compendex): p.
1506-1512.

90
A. ADDITIONAL MICROGRAPHS

APPENDIX A

ADDITIONAL MICROGRAPHS

Figure A.1: Predominantly martensitic microstructure of 100% dilution 8630 steel (Etch 5)

91
Figure A.2: 8630-625 85% dilution microstructure (Etch 5)

Figure A.3: F22-800H 76% dilution microstructure (Etch 5)


92
Figure A.4: X80 – C276 70% dilution microstructure (Etch 2)

Figure A.5: P92-800H 63% dilution microstructure (Etch 2)


93
Figure A.6: F22-800H 60% dilution microstructure (as polished)

Figure A.7: A508-FM52M 82.5% dilution microstructure (as polished)


94
Figure A.8: F92-800H 63% dilution microstructure (ferro-fluid)

Figure A.9: X80-C276 70% dilution microstructure (ferro-fluid)


95
Figure A.10: X80-C276 78% dilution microstructure (ferro-fluid)

Figure A.11: F22-800H 55% dilution microstructure showing possible TCP phases (Etch 4)

96
B. BUTTON MELT SAMPLE DATA

APPENDIX B

BUTTON MELT SAMPLE DATA

Button melt sample composition and ferrite number data are reported in

Table B.1 below. The dilution of a given button is listed in the first column between

the two materials from which it is formed.

97
Table B.1: Button melt sample composition and ferrite number data

Ferrite Number
Material

Mn

Mo

Nb

Cu

Co
Cr
Ni

Al

W
Ti
Si

V
P
C

S
8630 0.299 0.856 0.009 0.008 0.257 0.770 0.782 0.363 0.002 0.004 0.000 0.035 0.009 0.198 0.000 0.008
100.00% 0.299 0.856 0.009 0.008 0.257 0.770 0.782 0.363 0.002 0.004 0.000 0.035 0.009 0.198 0.000 0.008 151.4
89.99% 0.271 0.774 0.008 0.007 0.240 7.137 2.934 1.195 0.368 0.025 0.000 0.042 0.008 0.180 0.000 0.007 139.9
85.41% 0.258 0.736 0.008 0.007 0.232 10.056 3.921 1.577 0.536 0.034 0.000 0.045 0.008 0.171 0.000 0.007 117.6
80.35% 0.244 0.695 0.008 0.007 0.223 13.271 5.007 1.997 0.721 0.045 0.000 0.048 0.007 0.162 0.000 0.006 33.0
77.08% 0.234 0.668 0.008 0.006 0.217 15.351 5.711 2.269 0.840 0.052 0.000 0.050 0.007 0.156 0.000 0.006 0.5
51.14% 0.161 0.456 0.006 0.005 0.172 31.863 11.292 4.427 1.789 0.106 0.000 0.068 0.005 0.108 0.000 0.004 0.5
625 0.017 0.037 0.003 0.001 0.084 64.40 22.29 8.680 3.660 0.212 0.000 0.102 0.000 0.014 0.000 0.000
A553 0.044 0.582 0.004 0.001 0.189 8.970 0.140 0.010 0.001 0.004 0.004 0.018 0.001 0.034 0.050 0.012
100.00% 0.044 0.582 0.004 0.001 0.189 8.970 0.140 0.010 0.001 0.004 0.004 0.018 0.001 0.034 0.050 0.012 154.8
94.86% 0.044 0.697 0.004 0.001 0.181 12.127 1.212 0.010 0.115 0.006 0.004 0.017 0.001 0.037 0.047 0.014 152.7
89.05% 0.044 0.826 0.004 0.001 0.172 15.697 2.424 0.009 0.243 0.009 0.004 0.016 0.001 0.040 0.045 0.016 144.7
84.95% 0.043 0.917 0.004 0.001 0.165 18.218 3.280 0.009 0.334 0.011 0.003 0.015 0.001 0.042 0.042 0.018 119.6
83.18% 0.043 0.957 0.004 0.001 0.162 19.303 3.649 0.009 0.373 0.012 0.003 0.015 0.001 0.043 0.042 0.018 84.2
82.06% 0.043 0.982 0.004 0.001 0.160 19.989 3.882 0.009 0.397 0.012 0.003 0.015 0.001 0.044 0.041 0.019 36.1
80.10% 0.043 1.025 0.004 0.001 0.157 21.192 4.290 0.009 0.440 0.013 0.003 0.014 0.001 0.045 0.040 0.020 0.5
Alloy 82 0.040 2.810 0.004 0.001 0.030 70.40 21.00 0.005 2.210 0.050 0.000 0.000 0.000 0.090 0.000 0.050

98
Table B.1: Button melt sample composition and ferrite number data (Continued)

Ferrite Number
Material

Mn

Mo

Nb

Cu

Co
Cr
Ni

Al

W
Ti
Si

V
P
C

S
A508 0.182 1.449 0.005 0.002 0.181 0.920 0.125 0.514 0.002 0.002 0.000 0.016 0.004 0.040 0.000 0.025
100.00% 0.182 1.449 0.005 0.002 0.181 0.920 0.125 0.514 0.002 0.002 0.000 0.016 0.004 0.040 0.000 0.025 153.1
95.06% 0.174 1.415 0.005 0.002 0.177 3.908 1.555 0.491 0.042 0.012 0.000 0.018 0.004 0.039 0.000 0.024 149.2
87.02% 0.161 1.360 0.005 0.002 0.172 8.766 3.881 0.452 0.108 0.029 0.000 0.020 0.003 0.037 0.000 0.022 136.1
82.43% 0.154 1.328 0.005 0.002 0.169 11.539 5.208 0.431 0.146 0.039 0.000 0.022 0.003 0.036 0.000 0.021 116.5
79.88% 0.149 1.310 0.005 0.002 0.167 13.085 5.948 0.419 0.167 0.044 0.000 0.023 0.003 0.036 0.000 0.020 61.8
77.99% 0.146 1.297 0.005 0.002 0.165 14.225 6.494 0.410 0.182 0.048 0.000 0.023 0.003 0.036 0.000 0.019 2.0
FM52M 0.020 0.760 0.003 0.001 0.110 61.37 29.06 0.040 0.820 0.210 0.000 0.050 0.000 0.020 0.000 0.000
X80 0.023 1.370 0.014 0.004 0.113 0.013 0.040 0.010 0.059 0.042 0.000 0.031 0.005 0.033 0.000 0.003
100.00% 0.023 1.370 0.014 0.004 0.113 0.013 0.040 0.010 0.059 0.042 0.000 0.031 0.005 0.033 0.000 0.003 156.0
84.85% 0.020 1.212 0.013 0.004 0.141 8.761 2.466 2.433 0.077 0.037 0.002 0.031 0.007 0.037 0.508 0.033 147.0
77.98% 0.019 1.141 0.012 0.004 0.154 12.726 3.566 3.532 0.085 0.035 0.003 0.031 0.008 0.039 0.738 0.046 133.5
75.82% 0.018 1.118 0.012 0.004 0.158 13.976 3.913 3.878 0.088 0.034 0.003 0.031 0.009 0.040 0.810 0.051 110.7
73.90% 0.018 1.099 0.012 0.005 0.162 15.080 4.219 4.184 0.090 0.034 0.004 0.031 0.009 0.040 0.874 0.054 65.9
73.01% 0.018 1.089 0.012 0.005 0.163 15.599 4.363 4.328 0.091 0.033 0.004 0.031 0.009 0.040 0.904 0.056 2.3
70.13% 0.017 1.059 0.012 0.005 0.169 17.261 4.824 4.788 0.095 0.032 0.004 0.031 0.009 0.041 1.001 0.062 0.3
C276 0.003 0.330 0.007 0.006 0.300 57.75 16.05 16.00 0.179 0.010 0.014 0.030 0.020 0.060 3.350 0.200

99
Table B.1: Button melt sample composition and ferrite number data (Continued)

Ferrite Number
Material

Mn

Mo

Nb

Cu

Co
Cr
Ni

Al

W
Ti
Si

V
P
C

S
F22 0.141 0.430 0.011 0.022 0.292 0.128 2.420 1.030 0.004 0.002 0.000 0.025 0.017 0.190 0.000 0.010
100.00% 0.141 0.430 0.011 0.022 0.292 0.128 2.420 1.030 0.004 0.002 0.000 0.025 0.017 0.190 0.000 0.010 143.1
82.14% 0.128 0.480 0.010 0.018 0.308 5.625 5.721 0.891 0.010 0.061 0.000 0.071 0.014 0.156 0.000 0.008 146.8
78.82% 0.126 0.489 0.010 0.018 0.311 6.644 6.333 0.865 0.012 0.071 0.000 0.079 0.013 0.150 0.000 0.008 145.4
76.15% 0.124 0.497 0.010 0.017 0.313 7.469 6.828 0.844 0.013 0.080 0.000 0.086 0.013 0.145 0.000 0.008 137.6
66.05% 0.116 0.525 0.009 0.015 0.322 10.574 8.693 0.765 0.016 0.113 0.000 0.112 0.011 0.126 0.000 0.007 90.3
63.20% 0.114 0.533 0.009 0.014 0.324 11.453 9.221 0.743 0.017 0.123 0.000 0.119 0.011 0.120 0.000 0.006 54.2
60.17% 0.112 0.542 0.009 0.014 0.327 12.385 9.781 0.719 0.018 0.133 0.000 0.127 0.010 0.114 0.000 0.006 2.4
54.87% 0.108 0.556 0.009 0.013 0.332 14.015 10.760 0.678 0.020 0.150 0.000 0.140 0.009 0.104 0.000 0.005 0.3
800H 0.068 0.710 0.006 0.001 0.380 30.90 20.90 0.250 0.040 0.330 0.000 0.280 0.000 0.000 0.000 0.000
P92 0.120 0.430 0.010 0.001 0.280 0.120 8.270 0.380 0.010 0.010 0.045 0.006 0.174 0.190 1.615 0.000
100.00% 0.120 0.430 0.010 0.001 0.280 0.120 8.270 0.380 0.010 0.010 0.045 0.006 0.174 0.190 1.615 0.000 139.2
87.82% 0.114 0.464 0.010 0.001 0.292 3.868 9.808 0.364 0.014 0.049 0.039 0.039 0.153 0.167 1.418 0.000 141.0
84.93% 0.112 0.472 0.009 0.001 0.295 4.757 10.173 0.360 0.015 0.058 0.038 0.047 0.148 0.161 1.372 0.000 133.2
74.20% 0.107 0.502 0.009 0.001 0.306 8.062 11.529 0.346 0.018 0.093 0.033 0.077 0.129 0.141 1.198 0.000 90.2
68.71% 0.104 0.518 0.009 0.001 0.311 9.750 12.222 0.339 0.019 0.110 0.031 0.092 0.120 0.131 1.110 0.000 33.4
62.91% 0.101 0.534 0.009 0.001 0.317 11.538 12.955 0.332 0.021 0.129 0.028 0.108 0.109 0.120 1.016 0.000 0.4
800H 0.068 0.710 0.006 0.001 0.380 30.90 20.90 0.250 0.040 0.330 0.000 0.280 0.000 0.000 0.000 0.000
Min 0.017 0.456 0.004 0.001 0.141 0.770 0.782 0.009 0.002 0.004 0.000 0.014 0.001 0.036 0.000 0.000
Max 0.299 1.415 0.013 0.018 0.332 31.863 12.955 4.788 1.789 0.150 0.039 0.140 0.153 0.198 1.418 0.062

100

You might also like