+ All Categories
Home > Documents > Microstructural Evolution during Heat Treatment and High ...

Microstructural Evolution during Heat Treatment and High ...

Date post: 07-Feb-2022
Category:
Upload: others
View: 2 times
Download: 0 times
Share this document with a friend
139
Microstructural Evolution during Heat Treatment and High Strain Rate Deformation of an Fe-10Ni-0.1C Steel By Ian Harding Master of Science, Brown University, Providence, RI, 2015 Bachelor of Science, Temple University, Philadelphia, PA, 2013 A dissertation submitted to the School of Engineering in partial fulfillment of the requirements for the degree of Doctor of Philosophy Brown University Providence, Rhode Island May, 2019
Transcript
Page 1: Microstructural Evolution during Heat Treatment and High ...

Microstructural Evolution during Heat

Treatment and High Strain Rate

Deformation of an Fe-10Ni-0.1C Steel

By Ian Harding

Master of Science, Brown University, Providence, RI, 2015

Bachelor of Science, Temple University, Philadelphia, PA, 2013

A dissertation submitted to the School of Engineering in partial fulfillment

of the requirements for the degree of Doctor of Philosophy

Brown University

Providence, Rhode Island

May, 2019

Page 2: Microstructural Evolution during Heat Treatment and High ...

© Copyright 2019 by Ian Harding

Page 3: Microstructural Evolution during Heat Treatment and High ...

iii

This dissertation by Ian Harding is accepted in its present form by

the School of Engineering as satisfying the dissertation

requirement for the degree of Doctor of Philosophy.

Date________________

_________________________

Prof. K. Sharvan Kumar, Advisor

Recommended to the Graduate Council

Date________________

_________________________

Prof. Clyde Briant, Reader

Date________________

_________________________

Prof. C. Cem Taşan, Reader

Approved by the Graduate Council

Date________________

_________________________

Andrew Campbell, Dean of the

Graduate school

Page 4: Microstructural Evolution during Heat Treatment and High ...

iv

Curriculum Vitae

Education

- Ph.D. in Materials Science and Engineering 2019

Brown University, Providence, RI, USA

Advisor: Prof. Sharvan Kumar

- Sc.M. in Materials Science and Engineering 2015

Brown University, Providence, RI, USA

- B.S. in Mathematics and Physics 2013

Temple University, Philadelphia, PA, USA

Conference Presentations

Harding I, Mouton I, Gault B, Raabe D, and Kumar S, “The Stability of Precipitated

Austenite in Fe-10Ni-0.5Mn-0.1C Steel”. TMS 148th

Annual Meeting and Exhibition,

March 2019.

Harding I, Kumar S, “Dynamic Deformation Behavior of an Fe-Ni-C High Strength,

High Toughness Steel”. TMS 148th

Annual Meeting and Exhibition, March 2019.

Harding I, Kumar S, “Thermal Stability of Precipitated Austenite in Fe-10Ni-0.1C

Steel”. TMS 147th

Annual Meeting and Exhibition, March 2018.

Harding I, Kumar S, “Microstructural Evolution in Fe-10Ni-0.1C Steel during Dynamic

Deformation” (Poster). TMS 147th

Annual Meeting and Exhibition, March 2018.

Teaching Experience

Introduction to Materials Science, Lab TA (Falls 2015, ’16, ’17)

Page 5: Microstructural Evolution during Heat Treatment and High ...

iv

Acknowledgements

I would like to begin by thanking my family, who have always fostered my curiosity and

firmly believed in the value of education. They instilled in me a love of learning, and I am

grateful for the sacrifices they made for me to have this opportunity.

To Jenn- thank you for your patience, companionship, and endless support. It takes

significant dedication to get a PhD, but comparable devotion to help someone get through a PhD.

Thank you for your sacrifices for my dream, for helping me stay focused, and for looking out for

me when I forget to look after myself.

I would like to acknowledge the help from my labmates, who have taught me the

experimental methods I have used, who have helped me think through problems, who have

shared their extensive knowledge of the literature, and who have helped guide my research. In

particular, thank you to postdocs Hyokyung Sung, Hyunmin Kim, and Hyung Soo Lee. I am

grateful to the many technicians, without whom all research in the university would quickly

grind to a halt. Specifically, I would like to thank Tony McCormick for his help in all things

electron microscopy and John Shilko, Brian Corkum, and Chris Bull for their assistance in Prince

Lab. Additionally, I’d like to thank the Max-Planck-Institut für Eisenforschung for hosting me

and allowing me access to their facilities. I am grateful to the help of Isabelle Mouton, Bat Gault,

and Prof. Dierk Raabe.

I’d like to acknowledge the many friends I have made at Brown. Mariami Bekauri and

Leah Nation for getting through Materials courses together; the ‘Brown Engineering Mafia’ of

Stelios Siontas, Steve Racca, Peter Sun, Gerardo Pradillo, and others; the mentorship of Steve

Ahn, Jay Sheth, and Max Monn; and my good friends Jon Estrada, Hadley Witt, Alex Landauer,

Page 6: Microstructural Evolution during Heat Treatment and High ...

v

Rana Ozdeslik, Mrityunjay Kothari, and Mohak Patel. Research can be arduous and isolating,

and I cherish the time we spent together.

Lastly, I would like to thank the many academic mentors that I have had. I owe my

scientific and mathematical foundation to Mark Hammond and Eric Kemer. They believed the

foundation of science lies in rigorous understanding of fundamental principles, which in turn I

have relied heavily on throughout my higher education. I am forever grateful to my

undergraduate mentors, Dr. Ruth Ost and Prof. Dieter Forster, for believing in me and working

so hard on my behalf. Thank you both for your fierce advocacy and for your help in finding my

way. I’d like to thank Prof. Ke Chen for patiently teaching me the fundamentals of experimental

research; without the exposure I had in his lab, I undoubtedly would not have pursued this career.

Lastly, thank you to my advisor, Prof. Sharvan Kumar. I did not have a background in

engineering, but you had the patience for me to learn and change the trajectory of my career. I

appreciate your infectious dedication to learning and love of science, but more so your ability to

understand the big picture- life, family, happiness, and fulfillment.

Page 7: Microstructural Evolution during Heat Treatment and High ...

vi

Table of Contents

Curriculum Vitae ............................................................................................................................ 4

Education ..................................................................................................................................... 4

Conference Presentations ............................................................................................................ 4

Teaching Experience ................................................................................................................... 4

Acknowledgements ......................................................................................................................... 4

Chapter 1: Introduction ................................................................................................................... 1

Chapter 2: Technical Background .................................................................................................. 4

2.1: Outline .................................................................................................................................. 4

2.2: Fe-Ni Alloys and Fe-Ni low C Steels .................................................................................. 5

2.3: Quench and Partition, TRIP, and other Advanced High Strength Steels ........................... 14

2.3.1: Autotempering ............................................................................................................. 14

2.3.2: Thermal and Mechanical Stability of Austenite .......................................................... 16

2.4: Fe-10Ni-0.1C steels and the QLT Heat Treatment ............................................................ 21

2.5: Dynamic Deformation of Steels and Other Alloys ............................................................ 26

2.6: Scope of this Effort ............................................................................................................ 36

Chapter 3: Experimental Procedure .............................................................................................. 37

3.1: Materials and Heat Treatment Schedules ........................................................................... 37

3.2: High Strain Rate Deformation ........................................................................................... 39

3.3: Microstructure Characterization ......................................................................................... 41

3.3.1: Optical Microscopy: .................................................................................................... 41

3.3.2: Scanning Electron Microscopy (SEM):....................................................................... 42

3.3.3: Transmission Electron Microscopy (TEM): ................................................................ 43

3.3.4: Atom Probe Tomography (APT) ................................................................................. 50

3.3.5: Nano-Indentation ......................................................................................................... 52

Chapter 4: The Partitioning of Carbon During the Heat Treatment of Quenched Fe-10Ni-0.1C

Steel............................................................................................................................................... 53

4.1: Introduction ........................................................................................................................ 53

4.2: Results: L Temper .............................................................................................................. 54

4.3: Results: T’ Temper and QLT ............................................................................................. 61

4.4: Discussion .......................................................................................................................... 62

Page 8: Microstructural Evolution during Heat Treatment and High ...

vii

Chapter 5: The Partitioning of Ni During the Heat Treatment of Quenched Fe-10Ni-0.1C Steel 64

5.1: Introduction ........................................................................................................................ 64

5.2: The As-quenched Microstructure (AQ) ............................................................................. 64

5.3: The L Tempers Microstructure (QL and Q25L) ................................................................ 65

5.4: Discussion of Isothermal Tempering ................................................................................. 78

5.5: The QLT Treatment – Results and Discussion .................................................................. 82

Chapter 6: Microstructural Evolution in an Fe-10Ni-0.1C Steel During Dynamic Deformation 88

6.1: Introduction ........................................................................................................................ 88

6.2: Kolsky-Bar Calibration using 4140 Steel .......................................................................... 88

6.3: Kolsky-Bar Testing of 10Ni-QLT ...................................................................................... 89

6.4: Nanoindentation across the ASB ....................................................................................... 91

6.5: Microstructural Analysis of the dynamically deformed Specimen .................................... 93

6.5.1: Location far from the shear band: ............................................................................... 96

6.5.2: Location ahead of shear band: ..................................................................................... 97

6.5.3: Location adjacent to the shear band: ......................................................................... 102

6.5.4: Location inside shear band: ....................................................................................... 103

6.6: Discussion ........................................................................................................................ 112

Chapter 7: Conclusions ............................................................................................................... 115

7.1: Microstructural Evolution during Heat Treatment of an Fe-10Ni-0.1C Steel ................. 115

7.2: Microstructural Evolution during High Strain-Rate Deformation of an Fe-10Ni-0.1C Steel

Subjected to a Two-Stage Heat Treatment .............................................................................. 116

Chapter 8: Recommendations for Future Work .......................................................................... 118

References ................................................................................................................................... 121

Page 9: Microstructural Evolution during Heat Treatment and High ...

viii

List of Tables

Table 3.1: Nominal alloy composition .......................................................................................... 37

Table 3.2: Example table of calibrated diffraction radii ............................................................... 44

Table 4.1: C content in austenite and ferrite in Q25T’, Q125T’, and QLT as measured by APT

(at.%) ............................................................................................................................................. 61

Table 5.1: Composition of austenite in the four isothermal heat treatments as measured by APT

(at.%) ............................................................................................................................................. 77

Table 5.2: Size and composition of precipitates in isothermal treatments ................................... 78

Page 10: Microstructural Evolution during Heat Treatment and High ...

ix

List of Figures

Figure 2.1: Fe-rich end of the binary Fe-Ni equilibrium phase diagram [30]. Note that as

temperature decreases, the equilibrium Ni content in austenite increases rapidly. ........................ 5

Figure 2.2: Schematic representation of the QLT treatment (adapted from [12]). Three single-

stage tempered samples (QT2, QT100, and QL) were also produced to compare to the full QLT

treatment. ...................................................................................................................................... 12

Figure 2.3: a) The QLT heat treatment- the steel is first austenitized at 800°C for 60 minutes and

water quenched; tempered at 650°C for 40 minutes and water quenched; and then tempered at

590°C and quenched. b) The Fe-Ni binary equilibrium phase diagram with each QLT treatment

temperature marked [30]. c) The Fe-Ni Ms temperature diagram with the equilibrium Ni content

in austenite at each stage of QLT marked [43]. ............................................................................ 22

Figure 2.4: Overview of the QLT treatment. At the austenitizing temperature, the steel is fully

austenitic (with some undissolved fine carbides). When quenched, it forms a lath martensite

structure. The L-temper produces a high volume of precipitated austenite, though the thermal

stability of the austenite is not fully understood. The T-temper produces additional austenite

precipitation that is more Ni-rich than L-generation, and the final result is the presence of an

appreciable fraction of thermally stable austenite. ....................................................................... 23

Figure 2.5: Ballistic field test results that led to the combination of a 10Ni alloy and the QLT

treatment, as reported by Zhang [25]. ........................................................................................... 24

Figure 2.6: DICTRA model results compared to APT results, taken from [28], confirm austenite

precipitation in Ni-rich fresh martensite pockets that are formed by the transformation of some of

the L-generation austenite to martensite during the quench from the L temperature. .................. 26

Figure 3.1: Heat treatment schedules considered. All treatments begin with austenitization and

quench (AQ). Two single-stage tempering temperatures (650°C-L and 540°C-T’) with various

lengths are analyzed, as well as the two-stage temper QLT ......................................................... 38

Figure 3.2: Kolsky-bar setup. The sample is sandwiched between two bars; a projectile is fired at

the end of one bar, and the strain pulses through the incident bar, reflected back through the

Page 11: Microstructural Evolution during Heat Treatment and High ...

x

incident bar, and transmitted through the transmission bar are recorded with strain gauges. A

stress/strain curve is thus derived from these pulses. ................................................................... 39

Figure 3.3: a) an example of EDS analysis of QLT using a manually condensed beam (rough size

marked by pink circles) in the CM20. Austenite precipitates include roughly one datum point

each. b) STEM EDS on the JEOL 2100F allows for a high enough data point density to create

linear gradient plots within grains................................................................................................. 46

Figure 3.4: STEM mode images are mapped to BF using manually selected correlative point

pairs- i.e. a location is selected on the STEM image, and then the same point is selected in BF.

Several of these are used to create an affine map. Data is collected in small sets to avoid

significant drift during collection; these sets are uploaded in STEM space, mapped to BF mode,

and color coded by composition. Dozens of individual sets create dense composition maps. ..... 49

Figure 3.5: Sample EDS Ni gradient on a Q336T’ austenite precipitate ...................................... 50

Figure 3.6: A pictorial summary of the experimental process utilized in this effort to characterize

the undeformed microstructure as a function of heat treatment. .................................................. 52

Figure 4.1: a) TEM brightfield image of an austenitized and quenched (AQ) specimen showing a

lath martensite microstructure; some fine carbides (~20nm) are highlighted with dashed circles.

b) APT composition measurement of a carbide shows predominantly Mo but includes significant

amounts of V and Ti. c) C segregation along interfaces in martensite. ........................................ 56

Figure 4.2: a) TEM EDS/MDP shows Ni-depleted ferrite of about 8 wt.% Ni and Ni-rich (15-17

wt.%), thermally stable austenite precipitates. b) APT composition measurement of a carbide

particle in the QL specimen. A Ti- and V-rich core is surrounded by a Mo-rich but Ti- and V-

depleted outer shell (delineated by dashed lines). ........................................................................ 58

Figure 4.3: a) TEM EDS/MDP show the austenite precipitates are moderately Ni-rich (15 wt.%)

while the ferrite is close to equilibrium value of 5 wt.%. The precipitates are thermally unstable.

b) Interface segregation of C in fresh martensite in the Q25L specimen. c) APT measurement of

the composition across a carbide particle shows a V-rich, Mo-poor carbide is sandwiched by Mo-

rich, V-depleted carbide. ............................................................................................................... 60

Figure 5.1: TEM brightfield image with SADP of the AQ treatment. Note the lath martensite

structure......................................................................................................................................... 65

Page 12: Microstructural Evolution during Heat Treatment and High ...

xi

Figure 5.2: a) SEM images of QL and Q25L. After only 40 minutes of tempering, austenite has

precipitated out.; after 25 hours, the precipitates have evolved and occupy more than 50% of the

total area. b) TEM-EDS arrays start around bulk composition (10 wt.%) for short tempering, but

a bi-modal distribution is visible after 25 hours. .......................................................................... 66

Figure 5.3: a) TEM BF/MDP with corresponding EDS measurements (b) of the QL treatment.

Note that the center of the precipitate is Ni-rich, around 17 wt.% Ni, and radially decreases in

composition. This is an artifact of overlap between the austenite and ferrite, which causes a

dilution effect where the austenite is thinnest. This is confirmed from interfacial composition

measurements using APT (c), where the composition measured near the phase boundary matches

the composition measured by EDS at the center of the austenite particles. ................................. 68

Figure 5.4: a) TEM BF/MDP with corresponding EDS (b) of Q25L shows Ni content varies

throughout, with some pockets of ~17 wt.% Ni and large swaths of 12-15 wt.% Ni. Some of the

moderate Ni values are believed to be true, as they are from the thickest parts of the austenite,

and thus are unlikely to be an average of a higher Ni content averaged with low-Ni ferrite. b)

APT confirms that some of the low-Ni readings are not an artifact from overlapping phases, as

the boundaries can be as low as 12 wt.% Ni. Together, we can conclude that some regions within

Q25L austenite are quite Ni-rich, while others are very Ni-lean. MDP shows the austenite of

Q25L is thermally unstable. .......................................................................................................... 71

Figure 5.5: a) SEM comparison of T’ treatments. Ni diffusion at the T’ temperature is much

slower than at L, and so even after 5 hours there is not a significant volume of precipitates. After

125 hours, precipitates now line most lath/packet/block/grain boundaries, and after 336 hours,

coarsened globules are seen along high-angle boundaries. b) EDS measurements of the Ni

content in ferrite show a progressive shifts towards its equilibrium value with longer tempering

times .............................................................................................................................................. 73

Figure 5.6: a) TEM BF with corresponding EDS (b) of Q25T’ shows the core of precipitates

have a composition between 24-26 wt.% Ni with radially decreasing values. As in QL, APT (c)

confirms that this decrease is an artifact of phases overlapping through the foil thickness – APT

composition measurements near austenite/ferrite interface closely matches those measured with

EDS at the austenite core. As precipitates grow with increasing tempering time (Q125T’, d), the

overlap effect in EDS diminishes while the core value remains constant (e). APT again

Page 13: Microstructural Evolution during Heat Treatment and High ...

xii

corroborates EDS measurements(f). All precipitates in Q5T’-125T’ were found to be thermally

stable (FCC). ................................................................................................................................. 75

Figure 5.7: Q336T’ has large, coarsened, equiaxed austenite precipitates on the scale of 1-2µm.

They have a leaner composition than smaller T’ treatment precipitates (20-22 wt.% Ni vs 24-26

wt.%), and are sometimes thermally stable (a, c) or have transformed to martensite during the

quench (b, d). Note the thermally stable austenite (c) has a low defect structure, while the

thermally unstable austenite (d) has a lath-like internal structure. ............................................... 76

Figure 5.8: The Ms temperature diagram for a binary Fe-Ni alloy [43] predicts both L and T’

austenite to be thermally unstable at room temperature. The observed austenite stability in short

tempering times is thought to be due to a size-effect.................................................................... 80

Figure 5.9: For both L and T’ tempers, the small austenite from short tempers is thermally stable.

However, even the very Ni-rich T’ austenite becomes unstable after 336 hours, suggesting both

small L and T’ tempers are only stable due to their small size. If let coarsen long enough,

presumably all T’ austenite should be unstable. ........................................................................... 80

Figure 5.10: SEM of QL (shown here again for convenience) and QLT. Note that there is

significantly more austenite than QL, which suggests that the T temper contributes heavily to

additional austenite nucleation and growth................................................................................... 82

Figure 5.11: a) TEM BF of QLT with corresponding EDS (b) shows mixed composition of 15-20

wt.% at the center of precipitates. This is not an artifact of phase overlap, but rather is a result of

two distinct generations of austenite with different growth compositions. c) APT tips confirm

that both L-generation and T-generation austenite are present in QLT. ....................................... 83

Figure 5.12: QLT process produces thermally stable austenite around 15-17 wt.% Ni during the

L treatment. During T, growth resumes, resulting in additional growth in the range of 20-22

wt.% Ni. ........................................................................................................................................ 85

Figure 5.13: QLT Process, understood in terms of austenite size and composition. .................... 85

Figure 6.1: Strain/strain-rate pairs for Kolsky experiments on 4140 tempered martensite. Each

datum point correlates to a specific firing pressure with the noted bar. ....................................... 89

Figure 6.2: A diffuse and a well-developed adiabatic shear band ................................................ 90

Page 14: Microstructural Evolution during Heat Treatment and High ...

xiii

Figure 6.3: Sample chosen for microstructural analysis. A well-developed shear band initiates at

the bottom left, propagates towards the center of the sample (top right) and eventually dissipates.

Another band initiated in the opposite corner (not visible in this image) that also propagated

towards the center of the sample, though it was smaller and less developed than the one seen in

the montage above. The total height strain was about 20% and the strain-rate was about 1900s-1

.

Specific TEM specimen lift-out locations are marked using numbers from 1 to 10. The relative

displacements of the vertical etching bands (banding in the rolled plate) provide a sense of the

large shear experienced within the ASB. ...................................................................................... 91

Figure 6.4: Nanoindentation across the shear band (marked in purple). There is an increase in

hardness (~5GPa to ~7.5GPa) that extends beyond the boundary of the unetched portion of the

shear band. .................................................................................................................................... 92

Figure 6.5: SEM micrographs of QLT at increasing magnification (top to bottom), undeformed

(left) and near the shear band (right). Note the lack of features in the band, which gives it the

unetched appearance. Outside of this region, the austenite precipitates can be seen to be sheared

parallel to the shear plane. ............................................................................................................ 94

Figure 6.6: SADP in undeformed QLT compared to in the band using the same aperture. On the

left, the lath martensite structure produces a highly textured SADP; on the right, the highly

misoriented, equiaxed, mechanically recrystallized grains in the band produce a ring-like SADP.

....................................................................................................................................................... 95

Figure 6.7: TEM lift-out from a location that is far from the shear band (circled in blue dashed

line). .............................................................................................................................................. 96

Figure 6.8: Far from the shear localized region, the lath martensite seen in the undeformed

microstructure is still intact. Here, the composition of two austenite precipitates are measured

and found to be between 15-20 wt.% Ni. Thus these precipitates are a mixture of L- and T-

generation austenite, and MDP shows that they are mechanically stable. .................................... 97

Figure 6.9: a) TEM specimen lift-out from a location ahead of the shear band. Notice the stark

white band on the left, which dissipates to the right. b) SADP from the location ahead of the

band microstructure (aperture marked in orange). Ring pattern suggests significant grain

Page 15: Microstructural Evolution during Heat Treatment and High ...

xiv

refinement and grain rotation and recognized from the relatively low magnification bright field

image. ............................................................................................................................................ 98

Figure 6.10: a) Two adjacent subgrains, which are brought into contrast by tilting. b) EDS on the

first subgrain gives a composition of about 22 wt.% Ni (T-generation), and MDP shows it is

mechanically stable FCC. c) Its neighbor is also from T-generation austenite (20 wt.% Ni), and is

also mechanically stable austenite. Together, they show a T-generation austenite precipitate has

mechanically broken down into subgrains and these are mechanically stable. .......................... 100

Figure 6.11: An L- and T-generation austenite precipitate has mechanically broken down into

two subgrains (b and c) and can be brought into contrast through tilting (a). The subgrain before

tilting (b) is mechanically stable and has Ni content varying between L- and T-generation, and

was likely part of a core/shell structure before deformation. Its neighboring subgrain (c) is also

mechanically stable, but its composition shows that it is entirely growth from the T temper. ... 101

Figure 6.12: a) Lift-out location with ASB demarcated by the two orange dashed lines and axes

defined (shear plane normal is X axis, and shear band is propagating in the Z direction). b) TEM

BF image of area about 20µm from the band with rotated axes marked. Microstructure is

elongated in the shear plane (Y-Z).............................................................................................. 102

Figure 6.13: Both, the ferrite and the austenite precipitates adjacent to the band are aligned

parallel to the shear plane (axes marked) and have broken up into subgrains. The subgrain that is

in contrast here has Ni content ranging from 15-22 wt.%, suggesting it originates from a mixture

of L- and T-generation austenite. MDP shows the subgrain to have the FCC structure. ........... 103

Figure 6.14: Exaggerated schematic of TEM specimen lift-out geometries. The austenite

precipitates are stretched parallel to the shear band- when lift-outs are taken along this plane,

there is an averaging effect in the EDS measurements. Lift-outs are taken normal to the band to

mitigate this effect....................................................................................................................... 104

Figure 6.15: Left: Location of in-band lift-out prior to FIB milling. The unetched portion of the

band is approximately 4µm from the lift-out edge, while the other boundary of the band is about

10µm into the lift-out. Right: The measurements taken on the SEM are overlaid on a BF image

of the lift-out specimen (which extends beyond what is shown, as marked by the black arrows).

..................................................................................................................................................... 104

Page 16: Microstructural Evolution during Heat Treatment and High ...

xv

Figure 6.16: Normalized EDS Ni distribution of undeformed QLT (Green) in comparison to data

obtained from a TEM specimen that was lifted out parallel to the band (Blue) and another that

was oriented normal to the shear band (Red). The parallel geometry has a strong, narrow peak

around the nominal Ni content due to an averaging effect rising from the ratio of grain depth to

sample thickness; the normal-to-band sample has a distribution that closely matches the

undeformed material, suggesting it largely has single-grain depth. ........................................... 105

Figure 6.17: BF and corresponding DF image of the region of the deformed specimen that is

estimated to include the white band and the region immediately outside it. There is no observable

difference in microstructure, but note that the demarcation of the boundary is a best estimate

based on markings in Figure 6.15. .............................................................................................. 105

Figure 6.18: A small grain is in contrast. It is low in Ni (boundaries marked in orange on EDS

gradient plot, 5-7 wt.%Ni). Inside, a band-like substructure suggests twinning. ....................... 107

Figure 6.19: The high-Ni region, formerly an austenite precipitate, is spread over a very narrow

strip (about 20nm wide by 200nm in length). The portion that is in contrast is marked in the plot

in orange and has a Ni content of 20 wt.%, suggesting it is T-generation austenite; an MDP from

this location confirms it to be FCC and additionally reveals twinning- it is mechanically stable

austenite. ..................................................................................................................................... 108

Figure 6.20: The region presented in this image falls along the border of the white band. On the

left, the narrow strand of high-Ni (running left to right) is shown to be over 100nm long with

only a small grain of about 20nm in diameter in contrast. On the right, the composition of the

grain is isolated and highlighted in orange: it is about 20 wt.% Ni, and therefore came from T-

generation austenite. The MDP is BCC [111]- this grain is very fine, high in Ni but was

mechanically unstable. ................................................................................................................ 110

Figure 6.21: Another region close to the border of the white band; here, a small grain that

dynamically recrystallized from L-generation austenite (15-17 wt.%, marked in orange) has

mechanically transformed to martensite. .................................................................................... 111

Page 17: Microstructural Evolution during Heat Treatment and High ...

1

Chapter 1: Introduction

There is considerable interest in the U.S. Navy to improve the performance of high strength

low alloy (HSLA) steels used in structural naval applications, due to their cost, mechanical

properties, and fabricability [1]. Ballistic resistance of steels used in hulls and ship decks is a key

property that needs to be improved to meet the requirements of the next generation military

vessels.

Ni-alloyed, low-C steels are promising candidates for the next generation of naval ships

due to their high static strength and toughness, especially at low temperatures [2–18]. This effort

began within the US Navy with off-the-shelf Ni-steels with promising static properties [2–18],

that were then incrementally adjusted in composition and heat treatment through a series of

studies including microstructural analysis, quasi-static mechanical testing, and ballistic field-

testing [20–29]. At the time we began the current study, one particular alloy composition and a

specific heat treatment schedule were isolated as having superior ballistic resistance while

maintaining good static strength and toughness. Specifically, an Fe-10Ni-1.0Mo-0.6Mn-0.6Cr-

0.08V-0.1C (wt.%) steel heat-treated according to a two-stage temper called the QLT treatment

(hereinafter referred to as 10Ni QLT) was deemed optimized [25]. The QLT treatment produces

a dispersion of fine, Ni-rich, thermally stable austenite precipitates in a ferritic matrix [25,28]; it

is thus thought that the high quasi-static toughness and resistance to shear localization during

ballistic impact are due to the mechanical instability of this austenite. Specifically, by

mechanically transforming to martensite, the austenite increases the work hardening capacity of

the steel, which delays the onset of fracture in quasi-static loading and perhaps in high-strain rate

deformation as well [13,25]. However, further improvement in ballistic response is limited by the

Page 18: Microstructural Evolution during Heat Treatment and High ...

2

lack of a thorough understanding of microstructure/heat-treatment relationships and

microstructural evolution during dynamic deformation. These are the goals of this effort.

The thermal and mechanical stability of austenite are both largely governed by the

austenite precipitate size and composition. Therefore, in order to further improve 10Ni QLT, it is

critical to understand the relationship between heat treatment, austenite size and composition,

and its thermal stability. While some work has been done to partially address these questions, it

is not adequate or expansive enough to suggest improvements to 10Ni QLT [20,25,27,28].

Next, how do the various phases in the microstructure (i.e. austenite, ferrite, and carbides)

evolve during high strain rate deformation? Specifically, how does austenite dynamically evolve

as a function of its size, composition, and proximity to shear localization? It is broadly

understood that mechanical transformation does occur during ballistic field tests, but it is unclear

when during the localization phenomenon the transformation occurs, and if it is the main reason

that 10Ni QLT has superior ballistic properties over other treatments and alloys or if the stress-

induced austenite to martensite transformation is simply a side effect of high strain rate

deformation [25]. Understanding the evolution of QLT microstructure during dynamic

deformation will allow us to identify microstructural aspects that could be further tweaked and

could conceivably result in further enhanced ballistic response.

This effort fills some of these gaps in understanding. This thesis is organized in the

following manner. Chapter 2 (Technical Background) discusses the literature relevant to this

work. The areas covered in this review include studies pertaining to: heat treatment and

mechanical properties of Ni-containing alloys and steels for cryogenic applications; mechanical

properties, design, and characterization of advanced high strength steels including

Transformation-Induced Plasticity (TRIP) steels and Quench and Partition (Q and P) steels;

Page 19: Microstructural Evolution during Heat Treatment and High ...

3

microstructure characterization of 10Ni QLT and its predecessors; microstructural evolution of

steels and non-ferrous alloys during high strain rate deformation; and characterization of the

microstructural evolution of 10Ni QLT and its predecessors after high strain rate deformation.

The experimental methods used in this study are explained in Chapter 3. Chapter 4

discusses the partitioning of C during tempering in this alloy and its role on the stability of

precipitated austenite and precipitation and growth of carbides. In particular, the partitioning of

C during isothermal tempers of 40 minutes and 25 hours at 650°C is compared to the as-

quenched material to determine the partitioning behavior as a function of tempering time;

isothermal treatments at 540°C and the QLT treatment are also analyzed to compare the role of

tempering temperature on C partitioning and carbide formation. Chapter 5 details the results of

the analysis of Ni partitioning as a function of tempering time and temperature. Multiple

characterization techniques are used to determine the relationship between tempering time and

temperature on precipitated austenite chemical composition and size, which are then related to

austenite thermal stability. Analysis of isothermal heat treatments additionally provides possible

microstructure-based explanations for the superior mechanical properties of the two-stage QLT

temper. Chapter 6 includes the results and discussion of the microstructural evolution in 10Ni

QLT during high strain rate deformation. Microstructure from various locations relative to an

adiabatic shear band (ASB) is characterized to understand the microstructural evolution during

shear localization. The last chapters are the Conclusions chapter, which summarizes all results,

and a Chapter discussing future considerations.

Page 20: Microstructural Evolution during Heat Treatment and High ...

4

Chapter 2: Technical Background

2.1: Outline

In this Chapter, the literature is reviewed in the following four areas that are relevant to

the research reported in this thesis.

1) Fe-Ni Alloys and Ni-containing, Low-C Steels: The current state of knowledge

regarding austenite precipitation and mechanical properties of Fe-Ni alloys and Ni-

containing steels with relation to heat treatment.

2) Transformation Induced Plasticity (TRIP), Quench and Partition (Q+P), and other

Advanced High Strength Steels (AHSS): Although TRIP/Q+P steels are often

primarily Mn- or C-alloyed to produce thermally stable austenite, the research on

novel heat treatment schedules, austenite precipitation kinetics, and austenite thermal

and mechanical stability may be extended to 10Ni QLT, which is the subject of this

study.

3) Fe-(2.5-10)Ni-0.1C (wt.%) Alloy Development: Summary of research done on Ni-

containing steels for ballistic resistance, associated heat treatment optimization

studies leading to the 10Ni QLT material, and current understanding of the

microstructure evolution in this material.

4) High Strain Rate Deformation and Adiabatic Shear Banding (ASB): Current

understanding of high strain rate deformation in steels, including the 10Ni steel, as

well as in some non-ferrous alloys. The review attempts to connect prior research in

dynamic deformation mechanisms and microstructural evolution in other materials to

post-deformed microstructure observations in ballistic field-tested 10Ni QLT.

Page 21: Microstructural Evolution during Heat Treatment and High ...

5

2.2: Fe-Ni Alloys and Fe-Ni low C Steels

Nickel-containing steels were developed in the 1940s for use in cryogenic applications

due to the strong suppression of the ductile-to-brittle transition temperature (DBTT) of the steel

[2]. The microstructural basis for this increased toughness at low temperature, especially in

relation to heat treatment of the steel, was extensively studied in the 1970’s and 80’s [2–18].

While these studies did not have the spatial or compositional resolution of modern techniques

(e.g. APT), they nevertheless provide a foundation for our understanding of deformation

processes and a framework for our research. The studies considered in this section represent a

steady, incremental progression in understanding of the Fe-Ni system, and key results are

described.

Figure 2.1: Fe-rich end of the binary Fe-Ni equilibrium phase diagram [30]. Note that as temperature decreases, the equilibrium

Ni content in austenite increases rapidly.

The binary Fe-Ni phase diagram is shown in Figure 2.1. Ni is a strong austenite stabilizer

and upon cooling a binary Fe-rich Fe-Ni alloy from the single-phase austenite region, it

decomposes to yield relatively Ni-poor ferrite and Ni-rich austenite. Conversely, if a low Ni

alloy (for example Fe-(6-10) wt.% Ni) is suitably quenched from the austenite region to produce

Page 22: Microstructural Evolution during Heat Treatment and High ...

6

a lath martensite structure, upon tempering in the two-phase region, the martensite will

decompose into Ni-poor ferrite and Ni rich austenite; that is, Ni-rich austenite precipitates out

during tempering. If tempering time is adequate, the volume fraction and the composition of the

austenite can be simply determined from the phase diagram.

However, Ni is a substitutional element in Fe, and so at lower tempering temperatures

(i.e. below 550°C), Ni diffusion can be quite sluggish. For this reason, equilibrium

austenite/ferrite phase fractions may not be reached in practical tempering times, as noted by

Romig and Goldstein [31]. Nevertheless, Ni-alloyed steel (e.g. 5-15 wt.% Ni) can be

austenitized, quenched, and tempered (e.g. between 550-650°C) to produce Ni-rich austenite

[2,3,6,12,13], and if rich enough in Ni, the austenite may be thermally stable even after a

subsequent quench from the tempering temperature [32]. The presence of thermally stable

austenite in the steel has been shown to result in significant improvements in DBTT suppression

and low-temperature fracture toughness [2,4–18]. Understanding the relationship between heat

treatment, the low-temperature mechanical properties of the material, and the thermal and

mechanical stability of austenite were the focus of these studies.

To understand the relevant mechanisms that govern the relationship between heat

treatment and mechanical properties in Ni-alloyed steel, it is first important to describe the

morphology of precipitated austenite in Fe-Ni alloys and how it can be affected by tempering

time and temperature. The Ni-rich austenite precipitates along lath boundaries [2,3,6,13] and

obeys the Kurdjumov-Sachs relationship, as observed by Kim et al. [13] and Fultz et al. [2] in

tempered Fe-6Ni-0.1C (wt.%) steel and tempered Fe-9Ni-0.1C (wt.%) steel respectively.

Furthermore, Kim et al. found that the austenite in Fe-6Ni-0.1C (wt.%) steel that was thermally

unstable reverted to the orientation of the surrounding laths, while mechanically unstable

Page 23: Microstructural Evolution during Heat Treatment and High ...

7

austenite transformed to a different orientation [13]. The morphology of precipitated austenite

also depends on tempering temperature. Kim et al. found that when an Fe-6Ni-0.1C (wt.%) steel

was tempered at 670°C, the precipitated austenite had a lenticular morphology along interlath

interfaces, while the austenite formed at 600°C was blocky and equiaxed [13].

Low-C steels with nominal Ni content under 10 wt.% have been reported to peak in the

volume fraction of thermally stable precipitated austenite for a given tempering temperature.

That is to say, with sufficiently long tempering times, some of the precipitated austenite becomes

thermally unstable in this type of steel. This was observed by Marschall et al. in Fe-9Ni-0.1C

(wt.%) tempered between 590-650°C [33], by Hwang et al. in Fe-12Ni-0.25Ti (wt.%) alloy

tempered between 575-600°C [6], and by Fultz et al. in Fe-9Ni-0.1C (wt.%) steel tempered at

590°C [2]. The onset of thermal instability for extended isothermal tempering was not fully

understood, but Fultz et al. suggested that both chemical and microstructural factors contribute to

the stability of austenite [2,17].

The partitioning of Ni, C, and other minor alloying elements during tempering

contributes to austenite stability [17], and was directly measured using STEM EDS, Mössbauer

spectroscopy, and wet chemical analysis on chemically extracted austenite particles [2,12,17].

The most reliable composition measurements were made by Fultz et al. on an Fe-9Ni-0.1C steel

tempered at 590°C for a range of tempering times [2,16,17]. STEM EDS and Mössbauer

spectroscopy were both used to measure the composition of the precipitated austenite, but the C

content could not be measured using these techniques. Instead, the C content in the austenite was

calculated as a function of austenite volume fraction and ferrite lattice parameter as measured by

Mössbauer spectroscopy. The precipitated austenite was found to be enriched in Ni, Mn, Cr, Si,

and C, as predicted in [10,11]. Additionally, the composition of the austenite was found to be

Page 24: Microstructural Evolution during Heat Treatment and High ...

8

richest in Ni on the outer edges of the austenite particles, which was attributed to its slower

diffusion rate relative to other elements. The C content in the austenite held at 0.7 wt.% for the

first 10 hours of tempering and then decreased, an effect that was attributed to the ferrite

becoming depleted of C after 10 hours, thus diluting the C content in the austenite as it continued

to grow. While the error in measuring the minor alloying elements using these techniques was

relatively high, the measured solute enrichment was estimated to strongly lower the martensite

start (Ms) temperature of the precipitated austenite. However, the estimated Ms of the different

austenite particles was not enough to fully explain their thermal stability. Thus, they concluded

another factor must also contribute to the thermal stability of the austenite.

Hayzelden and Cantor related austenite grain size to thermal stability of the austenite

phase by measurement of the Ms temperature for an Fe-26Ni-0.4C (wt.%) steel [18]. They

showed that for austenite grains around 150µm there was no change in Ms temperature as grain

size decreased until grain size approached 10µm, where there was a suppression of Ms by

approximately 50°C. They did not measure grain sizes smaller than 10µm, but the trends

suggested that this suppression would increase with further grain refinement. Thus, combined

with the results of Fultz et al. [16,16,17], the thermal stability of precipitated austenite is strongly

affected by composition, but other microstructural factors including precipitate size also play a

significant role in controlling the thermal stability of austenite and therefore the deformation

mechanics and mechanical properties of the steel.

In addition to the thermal stability relationship of Hayzelden and Cantor [18], Fultz et al.

demonstrated that the mechanical stability of austenite was also a function of grain size in an Fe-

9Ni-0.1C steel [2,16,17]. The steel was tempered for various times at 590°C to produce a range

in austenite grain sizes, and then the sheets were cold-rolled to induce mechanical

Page 25: Microstructural Evolution during Heat Treatment and High ...

9

transformation. The defect structure surrounding transformed particles was used to qualitatively

estimate the transformation strain energy. Based on their observations, they concluded that larger

particles needed less strain energy to transform, and thus were more mechanically unstable.

Several microstructural factors have been observed to have beneficial effects on low-

temperature mechanical properties such as Charpy fracture energy, fracture toughness, and

DBTT: a reduction in grain size improves DBTT [3,4]; the precipitation of austenite, thermally

stable or not, improves Charpy upper shelf energy because of a gettering effect on interstitial

elements in the martensite [9,10,13]; increasing the volume of thermally stable austenite can

improve the low temperature Charpy fracture energy, fracture toughness, and suppress the DBTT

[5–7,11]; increasing mechanical stability of the austenite tends to lead to increases in fracture

toughness because it delays the onset of stress-induced austenite-to-martensite transformation

during loading [8,11,13]; and lastly the morphology of the austenite plays an important role in

the mechanical properties, with a fine distribution along lath boundaries reducing the effective

grain size by dissuading trans-packet cleavage [12–14]. Each of these findings is briefly further

examined in the following paragraphs.

Jin et al. found that the reduction in grain size in a ferritic Fe-12Ni-0.25Ti alloy from 40-

60µm down to 0.5-2µm through a cyclic annealing heat treatment lowered the DBTT to below

4.2K [3,4]. They argued that by decreasing the grain size, the critical stress for cleavage fracture

is increased above yield for lower temperatures, and thus DBTT is suppressed. Additionally, they

found that the precipitation of austenite through tempering (between 3-5 vol%, as measured by

XRD) further suppressed the DBTT beyond the effects of grain refinement [5]. The

microstructural mechanism behind the improved DBTT from thermally stable austenite was not

conclusively determined, but they suggested two possible hypotheses. First, there may be a

Page 26: Microstructural Evolution during Heat Treatment and High ...

10

beneficial TRIP effect wherein the austenite mechanically transforms to martensite, also

enhancing the overall ductility. Second, the austenite could serve as a sink for deleterious

elements in the ferrite (‘gettering’), thus increasing its ductility and suppressing the DBTT.

Deformation-induced martensite transformation of precipitated austenite and its influence

on Charpy fracture energy has been examined by several researchers. Syn et al. found that the 8

vol% austenite in a tempered Fe-9Ni-0.1C steel had fully transformed to martensite as far as

1mm from the fracture surface in Charpy tests at 77K [8]. Additionally, Fultz and Morris found

that for Charpy tests on an Fe-9Ni-0.1C steel, the transformation depth was similar to the plastic

zone size and thus the austenite had already transformed before it could interact with the

propagating crack front [15]. Lastly, Kim and Schwartz also investigated TRIP effects during

Charpy tests in an Fe-9Ni-0.1C steel [9]. They noticed that an increased volume of thermally

stable austenite improved Charpy fracture energy and found evidence for a TRIP effect as deep

as 300µm from the fracture surface. However, they estimated that the increased toughness from

the additional austenite mechanically transforming via a TRIP effect was insufficient to explain

the observed increase in Charpy impact energy. Thus, while a TRIP effect can contribute to

improved Charpy fracture energy, additional factors enabled by austenite precipitation likely

play a role in improved mechanical properties.

Gettering was directly shown to have a positive effect on Charpy upper shelf energy

[9,10,13,14]. The clearest demonstration was by Kim et al. on an Fe-6Ni-0.1C (wt.%) steel that

was treated with a two-stage temper [14]. The composition of the thermally stable precipitated

austenite and ferrite in the heat treated Fe-6Ni-0.1C was each measured in one of their previous

studies [12], and so, two steels with nominal composition of each phase (austenite and ferrite)

were produced to compare to the two-phase material. The grain sizes of the austenitic steel and

Page 27: Microstructural Evolution during Heat Treatment and High ...

11

the ferritic steel were reduced to be similar to that in the Fe-6Ni-0.1C heat treated steel, and then

the Charpy fracture energies at 77K and DBTT of all three were compared. The ferritic steel had

a similar upper shelf energy to the heat-treated Fe-6Ni-0.1C steel, and both the heat-treated Fe-

6Ni-0.1C steel and ferritic steel had higher upper shelf energy than the as-quenched Fe-6Ni-0.1C

steel. Thus, the removal of C and other elements from the martensite during tempering was

primarily responsible for the increased Charpy upper shelf energy seen in tempered Fe-6Ni-0.1C

steel. However, the single-phase steels had lower Charpy fracture energy than the tempered Fe-

6Ni-0.1C steel, suggesting the DBTT suppression was related to the two-phase morphology

rather than because of the properties of either phase individually.

Strife and Passoja [11] found that the mechanical stability of precipitated austenite played

a role in the low-temperature mechanical properties of low-C, Ni-containing steels. Two steels

were considered, an Fe-9Ni-0.1C steel with a single stage temper and an Fe-5Ni-0.1C steel with

a two-stage temper. While the 9Ni steel had improved fracture toughness with increased

austenite volume fraction, the two-stage tempered 5Ni steel was found to be more brittle with

increased austenite volume fraction. The difference in fracture behavior for austenite-containing

9Ni versus 5Ni suggested that improved mechanical stability in the precipitated austenite (which

can delay the stress-induced phase transformation and extend the strain hardening characteristic),

correlated to improved fracture properties, rather than simply austenite volume fraction.

Kim et al. proposed a morphology-based hypothesis to explain the effect of thermally

stable precipitated austenite on DBTT temperature that took into account the volume, thermal

stability, and mechanical stability of the precipitated austenite [12–14]. While precipitated

austenite does beneficially serve as a sink for deleterious elements like C, the presence of

thermally stable austenite that is resistant to mechanical transformation along lath boundaries is

Page 28: Microstructural Evolution during Heat Treatment and High ...

12

additionally helpful because it decreases the effective grain size by the following mechanism. At

low temperatures, the martensitic material is susceptible to trans-packet cleavage because the

laths are similarly oriented, and so fracture can travel across the similarly-oriented cleavage

planes. By adding mechanically stable austenite along these lath boundaries, the cleavage planes

across laths are broken up, and the effective grain size is decreased. Therefore, increasing the

volume fraction of austenite that resists transformation should suppress the DBTT and improve

the fracture toughness.

Figure 2.2: Schematic representation of the QLT treatment (adapted from [12]). Three single-stage tempered samples (QT2,

QT100, and QL) were also produced to compare to the full QLT treatment.

Page 29: Microstructural Evolution during Heat Treatment and High ...

13

To test their hypothesis, Kim et al. considered an Fe-6Ni-0.1C (wt.%) steel that was

austenitized at 800°C (Q) and then given various heat treatments (Figure 2.2) to produce

different austenite morphologies: one tempered at 670°C (QL), two tempered at 600°C for

different times (QT2 and QT100), and one given combined two-stage temper at 670°C and

600°C (QLT) [12–14]. The isothermal QT (600°C) tempers produced austenite far slower than

the L (670°C) temper and the austenite had a blocky, spherical morphology. The L temper

produced lenticular austenite, but it was thermally unstable due to a lower solute content. The

QLT sample was found to have a lenticular austenite, despite being tempered at T (which

produced blocky austenite in QT2 and QT100), and so it was determined that its lenticular

morphology was due to how it formed. The QL step resulted in lenticular pockets of Ni-rich

martensite, and the additional T step caused an austenite with even richer solute content to

precipitate in these pockets and inherit their morphology [13].

All four tempers had improved Charpy upper shelf energy over the as-quenched

microstructure (Q), which was attributed to the gettering of C from the martensite matrix. This

effect occurs even if the austenite is thermally unstable, as in QL, and thus QL had better Charpy

upper shelf energy than Q. However, the QLT treated sample had a higher Charpy fracture

energy and DBTT than the sample with the long T temper, despite having austenite with similar

composition and volume fraction. Thus, the difference between the DBTT of QT100 and QLT

was attributed to morphology: the lenticular austenite in QLT had better resistance to trans-

packet cleavage than the blocky-austenite in QT100. Additionally, Kim et al. noted that

thermally unstable austenite reverted to the orientation of neighboring laths, while mechanically

unstable austenite (produced during Charpy tests) transformed to a different orientation (as

mentioned previously) [13].

Page 30: Microstructural Evolution during Heat Treatment and High ...

14

In summary, the volume fraction, thermal stability (composition and size dependence),

and morphology of austenite precipitated during tempering of low carbon-(5-10) wt.% Ni steels

is intimately tied to the low-temperature mechanical properties of the steels and these

microstructural aspects rely on tempering schedules (temperature, time, and multi-stage).

However, the role of minor alloying elements, most notably C, has not been adequately

identified and so the extent of their role on austenite thermal and mechanical stability or

mechanical properties of the steels has not yet been rigorously determined.

2.3: Quench and Partition, TRIP, and other Advanced High Strength Steels

While the work on cryogenic Ni-alloyed steels from the 1970’s and 80’s was limited by

the inability to accurately quantify light and minor alloying elements in precipitated austenite,

measure concentration gradients on the nanometer scale, or characterize nano-scale features,

these types of experiments have recently been enabled by advanced experimental capabilities and

performed in more contemporary work on steels with similar compositions and mechanical

properties. A few relevant studies have been isolated, and results from these are discussed in the

following section.

2.3.1: Autotempering

Through the use of high-resolution TEM and atom probe tomography (APT), it has been

observed that C can partition to small amounts of austenite along lath boundaries as martensite

forms during quenching, even in low-C steels [34–36]. This C-enrichment can be high enough

that these austenite films are thermally stable after the quench. This phenomena is called

‘autotempering,’ and depends on quenching rate [34].

Sherman et al. measured the C enrichment in these films as a function of quench rate in

Fe-1.3C-3.2Si-3.2 (Mn, Ni, Cr, Mo, Cu, Al, Ti, V) (at.%) using APT [34]. For samples quenched

Page 31: Microstructural Evolution during Heat Treatment and High ...

15

at a rate of 55K/s, there was an enrichment with a peak of 10 at.% C and a width of 4nm in small

C-rich clusters, and austenite films of 12-18nm were observed using TEM. For samples

quenched at a rate of 560K/s, the films were too fine to identify in TEM, but APT showed 4nm

wide C-rich planes with a peak of 6 at.% C. 90% of the C that remained in the martensite in the

560K/s quenched steel was estimated to be trapped in Cottrell atmospheres at dislocation cores.

Morito et al. quantified C-enrichment from autotempering in an Fe-2Mn-0.2C (wt.%) steel

which was austenitized at 1200°C and water quenched to room temperature [35]. Despite the low

nominal C concentration, APT showed a local C enrichment in 3nm austenite films to the extent

of 4.5 at.% (~1 wt.%). The C content in the martensite matrix was measured to be 0.08 wt.%,

which was lower than the 0.2 wt.% initial C content in the alloy. The decrease in the C

concentration in the bulk martensite was attributed to short-range diffusion of the C to retained

austenite and dislocation structures in the martensite over six months at room temperature.

Morsdorf et al. noted that the extent of autotempering in individual martensite laths

depended on when during the quench the martensite transformed [36]. They analyzed an Fe-5Ni-

0.1C (wt.%) steel that was austenitized at 900°C and water quenched. Two types of martensite

were found: a coarse-lath martensite that formed early in the quench and fine-lath martensite that

formed closer to the martensite finish temperature. The coarse-lath martensite had less interstitial

C than the fine-lath martensite (~0.15 at.% vs 0.3-0.7 at.% respectively) because the early-

forming martensite had more time at high temperature than later-forming martensite, thus

allowing for additional C diffusion during the quench. Additionally, the C in the late-forming,

fine-lath martensite was found to segregate to Cottrell atmospheres. They estimated that even if

the alloy was quenched at 1000K/s, the C in the early-forming laths could diffuse as far as 1.5µm

Page 32: Microstructural Evolution during Heat Treatment and High ...

16

(on the scale of the coarsest lath widths), while late-forming laths were confined to nanometer-

scale local diffusion.

2.3.2: Thermal and Mechanical Stability of Austenite

The idea of Quench and Partition (Q+P) steels was first proposed by Speer et al. [37] as a

novel heat treatment schedule for C-containing steels to produce a large, tunable volume of

thermally stable austenite. The process involves austenitizing and quenching the steel to a

quench temperature between the Ms and Mf temperatures, which retains a controlled volume of

austenite. The steel is then tempered at or above the quench temperature to allow the partitioning

of C from the martensite to the austenite, thus thermally stabilizing it during a final room-

temperature quench. Specifically, Speer et al. [37] noted that there was an optimum austenite

volume fraction given the total available C in the system; further, the volume of austenite during

the temper and the C content in the austenite after the temper are inversely related and depend on

the quench temperature. If the austenite after the temper is rich enough in C, it will be stable after

the final quench. Therefore, there is an optimum quench temperature in which the amount of C in

the retained austenite is just enough to thermally stabilize it at room temperature.

Takaki et al. studied the relationship between austenite grain size and the austenite to

martensite transformation in an Fe-16Cr-10Ni steel [38]. The composition was chosen such that

the Ms temperature was around room temperature. The steels were initially austenitic, cold-

drawn to mechanically fully transform the austenite to martensite, and then tempered at different

temperatures around 630°C (the austenite reversion temperature) for 10 minutes to create initial

austenite grain sizes between 0.8-80µm. The samples were held at room temperature and at 77K

each for up to 700 hours to allow the austenite to transform to martensite. The transformation

behavior could be classified by three grain sizes: coarse grain (~80µm), fine grain (~10µm), and

Page 33: Microstructural Evolution during Heat Treatment and High ...

17

ultra-fine grain (~0.8µm). The coarse grain austenite had 20% transformation to martensite at

room temperature and 50% at 77K, the fine grain austenite had about 4% transformation to

martensite at room temperature and 35% at 77K, while the ultra-fine grain had almost no

transformation at either temperature. Using TEM, they showed that coarse-grain austenite

formed several packets each with multiple blocks of laths about 1µm wide (multi-variant

transformation); the fine-grain austenite formed martensite with a single-variant or single habit

plane; and what little ultra-fine grain austenite that did transform did not have a lath structure and

was single-variant.

The reason for the suppression of the austenite-to-martensite transformation in ultra-fine

grain austenite was rationalized through an estimation of the elastic strain energy barrier for

transformation in multi-variant transformation versus single-variant transformation. The elastic

strain energy for multi-variant transformation can be thought of as isotropic, as there are many

lath orientations present in the original austenite grain. In contrast, single-variant transformation

is highly anisotropic. The difference in these energies is substantial: approximately 1840MJ/m3

for single-variant versus 70MJ/m3 in multi-variant. Therefore, they concluded that a substantial

chemical driving force is necessary for sub-micron austenite to transform to martensite in

thermally metastable austenite.

Size also plays a role in the mechanical stability of austenite, as studied by Wang et al. in

manganese-containing TRIP steels [39,40]. In their first study, an Fe-9Mn-3Ni-1.4Al-0.01C

(wt.%) steel was austenitized, quenched, and then tempered at 600°C for 8 hours to precipitate

Mn- and Ni-rich austenite along lath boundaries. The precipitated austenite was thermally stable

at room temperature, and while the austenite precipitates varied in size, the precipitates had

similar composition. Furthermore, the former austenite grain size was large enough to form

Page 34: Microstructural Evolution during Heat Treatment and High ...

18

multiple martensite packets, and so the austenite that precipitated within a particular packet had

similar crystalline and geometric orientation to each other with respect to the loading axis during

deformation experiments. Thus, the effect of size on deformation mechanisms was isolated from

effects of composition and orientation and their findings are described below.

The mechanical stability of the austenite was analyzed using EBSD in-situ as well as

after tensile and three-point bend tests, and the fine-scale microstructure was analyzed using

TEM [39]. The deformation behavior of the austenite was categorized by their size (surface area)

as small austenite precipitates (0.1-0.3µm2) and large austenite precipitates (0.3-4µm

2). At

relatively low global strains, the large precipitates showed little signs of plastic deformation,

while the small precipitates primarily deformed through stacking fault multiplication and

transformed to martensite. At higher global strains, the larger austenite precipitates were seen to

form subgrains separated by mechanical twins; the small subgrains then deformed through the

multiplication of stacking faults and eventually transformed to martensite. They concluded that

while smaller austenite is thought to be mechanically more stable than large grains, there is also

a size-dependent TWIP effect that can affect the TRIP mechanism. In TRIP steels with rather

homogenously sized and composed austenite, the austenite transforms at low strains and roughly

at the same global strains, which could actually promote crack propagation and failure [41],

suggesting delaying the transformation and spreading it out over a strain spectrum would be

beneficial. Specifically, in their experiment, the austenite-martensite transformation was spread

over a larger range of global strains because of the difference in deformation mechanisms for

small austenite precipitates versus large austenite precipitates. Thus, the mechanical properties of

the material were improved through a ‘spectral TRIP effect’.

Page 35: Microstructural Evolution during Heat Treatment and High ...

19

Wang et al. then designed a thermo-mechanical treatment to produce a wider distribution

of precipitated austenite sizes in Fe-9Mn-3Ni-1.4Al-0.01C [28] than was available in their

previous study [39]. Before tempering, the steel was cold-rolled to increase the density of

nucleation sites for austenite. The associated nucleation energies for defects introduced by cold-

rolling vary; for example, high-angle boundaries have a higher interfacial energy than low-angle

boundaries, and thus lower the energy barrier for austenite nucleation during tempering. By

creating more nucleation sites with varied associated nucleation energies, not only was austenite

precipitation accelerated in comparison to the as-quenched material and tempered material, but

the size of austenite varied as well due to staggered nucleation rates. The austenite that

precipitated early had more time to grow than later nuclei, and so there was a wider range of

austenite size present after tempering. As in the as-quenched material, the cold-rolled steel also

experienced a spectral TRIP effect; furthermore, the mechanical properties were further

improved as a result of the enhanced spectral TRIP compared to as-quenched material that was

tempered to produce a similar volume fraction of austenite.

The spectral TRIP effect was also observed in austenite with varied composition. Yuan et

al. studied an Fe-13.6Cr-0.44C (wt.%) steel that was Q+P treated by austenitizing at 1150°C,

water quenching, and tempering at 300°C, 400°C, and 500°C for 1-30 minutes [41]. Some

austenite remained after the initial water quench (8-20 vol%), and as seen by Sherman et al.,

Morito et al., and Morsdorf et al. [34–36], the C segregated out of the martensite to

austenite/martensite and martensite/martensite interfaces during the quench. With subsequent

tempering, additional C segregated to this new C-rich layer, which reverted to austenite with a

different C content than the retained austenite. After quenching from the tempering temperature,

the final microstructure contained thermally stable austenite with varied sizes and C

Page 36: Microstructural Evolution during Heat Treatment and High ...

20

concentrations. Because the composition of the austenite varied in C, the mechanical stability of

the austenite was also varied. Overall, the composition-aided spectral TRIP effect in these steels

improved the toughness of these steels.

Compositionally distinct layers in austenite have also been produced in low-C Mn

maraging steels. Dmitrieva et al. austenitized and water-quenched an Fe-12Mn-2Ni-0.05C steel

to produce a sample containing a mixture of retained austenite and martensite, both with the

same nominal composition [42]. The steel was then tempered for 48 hours at 450°C to produce

additional Mn-rich austenite. The austenite composition profiles were measured using APT, and

the Mn diffusion kinetics modelled using DICTRA. They found that the austenite in the final

material had two distinct compositions: the nominal composition (retained austenite) and a Mn-

rich region between the retained austenite and the ferrite matrix. The diffusion of Mn at 450°C in

austenite is several orders of magnitude slower than the diffusion of Mn in martensite, which led

to the formation of a Mn-rich boundary at the austenite/martensite interface. This region

continues to enrich until it reaches a local equilibrium, and then the interface grows into the

martensite at the composition dictated by the local equilibrium. After quenching, the final

microstructure consists of two compositionally distinct layers of austenite, the austenite retained

from the first quench and the austenite that formed during the temper.

It is thus seen that in low-C steels there are nanometer-scale features that arise during

quenching and tempering that can have a visible and substantial effect on mechanical properties.

A spectral TRIP effect can be produced in steels with austenite of varied size and composition

that can increase the toughness of the steel. Nanoscale composition variation has been measured

in tempered C-alloyed and Mn-alloyed steels (with advanced experimental capabilities available

in the past two decades), that give rise to this composition and size-based spectral TRIP effect

Page 37: Microstructural Evolution during Heat Treatment and High ...

21

and enable enhanced understanding of the TRIP and TWIP phenomena. Together with improved

computational capabilities, such understanding enables pathways to improved advanced high

strength-high toughness steel design.

2.4: Fe-10Ni-0.1C steels and the QLT Heat Treatment

In this section, we focus on the characterization and optimization that led to the 10Ni

QLT heat treatment and the current understanding of microstructure evolution during the QLT

heat treatment, all of which is directly relevant to the scope of the current effort. Alloy design,

alloy production, and heat-treatment optimization within the family of Ni-containing steels

targeted for improved ballistic resistance for ship structure has resided largely within the U.S.

Navy research lab(s) and has historically relied on a combination of laboratory-scale testing for

static properties and performance in field tests for dynamic properties [25]. Microstructural

analysis has largely been on the optical and scanning electron microscopy level, rather than on

detailed fine structures, although recent US Navy-sponsored research at universities and the

Naval Research Laboratory (NRL) has been filling this gap in knowledge [22,23,27–29].

However, the microstructural understanding has lagged the alloy development and heat treatment

optimization processes.

The nominal alloy composition of the 10Ni steel is Fe-10Ni-1.0Mo-0.08V-0.6Mn-0.6Cr,

and the QLT heat treatment is shown schematically in Figure 2.3a. The QLT heat treatment

begins with an austenitization at 800°C for one hour and water quenching (Q). It is given an

initial ‘lamellarizing’ (L) temper at 650°C for 40 minutes and is then water quenched. It is then

given a final temper (T) at 590°C for an hour and again water quenched. As discussed above, the

lower the tempering temperature, the more Ni-rich is the precipitated austenite (Figure 2.3b).

Page 38: Microstructural Evolution during Heat Treatment and High ...

22

Thus, this treatment significantly decreases the Ms temperature of the austenite through Ni-

enriched austenite (Figure 2.3c).

Figure 2.3: a) The QLT heat treatment- the steel is first austenitized at 800°C for 60 minutes and water quenched; tempered at

650°C for 40 minutes and water quenched; and then tempered at 590°C and quenched. b) The Fe-Ni binary equilibrium phase

diagram with each QLT treatment temperature marked [30]. c) The Fe-Ni Ms temperature diagram with the equilibrium Ni

content in austenite at each stage of QLT marked [43].

The microstructural evolution during this QLT heat treatment is schematically illustrated

in Figure 2.4. The austenitization treatment is sufficient to convert the steel into polycrystalline

austenite with a fraction of undissolved MC and M2C carbides left in it; upon water quenching to

room temperature, a lath martensite microstructure is observed. The L-stage temper produces

moderately Ni-rich austenite which may be thermally unstable upon water quenching. The T-

stage temper produces additional austenite that is more Ni-rich than L because of the lower

Page 39: Microstructural Evolution during Heat Treatment and High ...

23

tempering temperature, and so this generation of austenite is more likely to be thermally stable

than that resulting from the L temper. The end result of this multi-step process is a tempered steel

with a large volume fraction of thermally stable austenite (around 18 vol% as measured by Jain

et al. [28] and Zhang [25]).

Figure 2.4: Overview of the QLT treatment. At the austenitizing temperature, the steel is fully austenitic (with some undissolved

fine carbides). When quenched, it forms a lath martensite structure. The L-temper produces a high volume of precipitated

austenite, though the thermal stability of the austenite is not fully understood. The T-temper produces additional austenite

precipitation that is more Ni-rich than L-generation, and the final result is the presence of an appreciable fraction of thermally

stable austenite.

It is worth noting that prior to heat treating, these steels are hot rolled into plates and

macro-segregation resulting from solidification manifests in the rolled plate as a banded structure

of solute-rich and solute-poor bands [19] that are on a scale too large to be eliminated by any

form of solid-state heat treatment. Therefore, these regions remain distinct through

austenitization and subsequent tempering and will influence austenite precipitation and have

been shown to have an effect on failure mechanisms (e.g. Charpy fracture surfaces [19]).

The 10Ni QLT alloy composition and heat treatment combination was isolated through

ballistic field tests that were conducted on several alloys ranging from 2.5 to 10 wt.% Ni in

composition that had been subjected to heat treatments ranging from single isothermal tempers at

Page 40: Microstructural Evolution during Heat Treatment and High ...

24

different tempering temperatures through various multi-temper heat treatments [25]. Ballistic

resistance in these field tests was quantified by a ‘V50’ metric. Namely, given a particular

caliber projectile, plate specification, and firing distance, the V50 is the velocity at which half of

the projectiles fully penetrate the plate. The results of these ballistic tests showed that the

combination of a 10Ni alloy with the QLT treatment as described above showed superior

ballistic resistance while still maintaining high static tensile strength, elongation, and Charpy

impact energy (Figure 2.5).

Figure 2.5: Ballistic field test results that led to the combination of a 10Ni alloy and the QLT treatment, as reported by Zhang

[25].

Isheim et al. [27] performed microstructural analyses on of some of the precursors to the

10Ni QLT alloy, including a 4.5Ni and a 6.5Ni alloy subjected to a three-temper treatment and a

10Ni alloy subjected to heat-treatment with different L and T temperatures. Comparing APT and

TEM results with DICTRA simulations, they determined that the higher temperature L-tempers

were largely responsible for Ni partitioning and most of the austenite volume in the final state;

Page 41: Microstructural Evolution during Heat Treatment and High ...

25

the lower temperature T-temper merely allowed the precipitation of a new layer of austenite

around L-austenite, but was not long enough to equilibrate the Ni content throughout the

austenite. This correlated with the SEM results of Zhang [25] where the precipitated austenite

volume fraction in the optimized 10Ni QLT treatment was compared to just the microstructure

after QL and after a QT treatment (QLT without the intermediate L). The QT treatment had very

little austenite in comparison to QL and QLT, and so Zhang suggested that the T-temper was

primarily responsible for austenite precipitation within any fresh martensite after L (i.e. in

thermally unstable L precipitates) and their additional growth.

The effects of each stage of the optimized QLT treatment on the microstructure evolution

in the 10Ni alloy was recently examined by Jain et al. by using APT and synchrotron techniques

to compare austenite precipitation in material subjected to the QT, QL, and QLT schedules [28].

Using dilatometry, they determined that some of the austenite precipitated during the L stage was

thermally unstable during the quench and that the final volume fraction was about 8% austenite

by volume. The QT sample had only 3 vol% austenite, and the QLT treatment resulted in 18

vol% austenite. Thus, the increase in austenite volume seen between QL and QLT could not be

ascribed simply to nucleation and growth of austenite precipitates by long-range diffusion of Ni

out of the ferrite during the T treatment, as QT alone had far less austenite than the difference

between QL and QLT. APT showed the composition of the L-generation austenite to be between

12 and 16 wt.% Ni. Two chemically distinct generations of austenite were seen in QLT

specimens, one matching the composition of the L-generation believed to be retained L-

austenite, and a second generation about 21 wt.% Ni that formed during T. Therefore, the

majority of the increased volume in thermally stable austenite between QL and QLT was through

Page 42: Microstructural Evolution during Heat Treatment and High ...

26

the precipitation of new austenite in the L-generation fresh martensite pockets. DICTRA models

supported this result, as seen in Figure 2.6.

Figure 2.6: DICTRA model results compared to APT results, taken from [28], confirm austenite precipitation in Ni-rich fresh

martensite pockets that are formed by the transformation of some of the L-generation austenite to martensite during the quench

from the L temperature.

Additionally, Gupta and Kumar [29], Wang and Kumar [22], Isheim et al. [27], and Jain

et al. [28] each reported the presence of Mo-rich MC and M2C carbides in tempered samples in

these alloys. The steels they considered were compositionally the same but varied in Ni content

and tempering temperature (ranging from 450-650°C). Thus, some of the C in these steels is

removed from the matrix via carbide precipitation and growth. The relationship between C

partitioning and carbide formation kinetics as a function of heat treatment and hence its role in

affecting austenite stability (in particular in the L temper) has not been clarified.

2.5: Dynamic Deformation of Steels and Other Alloys

While the quasi-static mechanical properties of QLT-treated 10Ni steel are relevant, more

important to this effort is the alloy’s response to high strain-rate deformation because of the

intended application as hull and ship deck armor. In particular, we consider the phenomenon of

Adiabatic Shear Banding (ASB), an acute form of shear localization often seen in ballistic

Page 43: Microstructural Evolution during Heat Treatment and High ...

27

applications that implies loss of structural integrity. In high strain-rate deformation, the

temperature that arises from the localized plastic deformation can be quite significant (on the

order of 500°C [44–48]), as there is not adequate time for the heat generated by deformation to

dissipate into the bulk (thus, adiabatic). If the work-hardening from plastic deformation is

overridden by the thermal softening due to this temperature increase, the localization becomes

unstable, concentrating the deformation zone to an increasingly smaller, hotter, and softer band.

Inside this band, the presence of extreme conditions exemplified by a combination of large strain

rate, large strain, and high temperature, enables deformation mechanisms that are otherwise not

active elsewhere in the material and certainly not present in quasi-static deformation. These

include temperature- and/or strain-induced phase transformations (usually diffusionless, as the

duration is of the order of milliseconds), and deformation modes like micro-twinning, carbide

plastic deformation and dissolution, and mechanical recrystallization. Eventually, the material

fractures by cracks forming and linking within this adiabatic shear band. Thus, understanding the

microstructural characteristics that discourage shear localization is central to designing materials

that can better resist ballistic impact. While only limited analysis exists on the microstructural

evolution in 10Ni steels during high strain-rate deformation in general, and even less in 10Ni

steel subjected to the QLT treatment, we can use the extensive literature on high strain rate

deformation in other steels and non-ferrous alloys to guide our thinking in this effort.

Adiabatic shear banding was first described in steel by Zener and Hollomon in 1944 [49],

but it has since been observed and studied in a wide range of materials including various steels,

pure iron, titanium and titanium alloys, aluminum and aluminum alloys, copper, brass, tantalum,

uranium, magnesium, and zirconium [49–57]. Generally, ASBs are seen in materials exhibiting a

low strain-hardening rate, low strain-rate sensitivity, low thermal conductivity, and high thermal

Page 44: Microstructural Evolution during Heat Treatment and High ...

28

softening rates; the easier it is for thermal softening to overcome work hardening for a given

material, the more prone the sample is to localization. In terms of application, shear banding can

occur in ballistic plate armor; self-sharpening projectiles; high-velocity machining, shaping, and

forming; and extrusion and punching [52–54]. For a given alloy, there are several microstructural

factors that can affect shear localization, including defect structure, grain size, and prior strain

hardening [58–62]. ASB formation also depends heavily on specimen geometry, surface defects

(e.g. scratches), initial temperature, stress state, and strain rate [46,53,54,61–63].

Strain rates in the range of 103 (similar to ballistic impact) can be achieved with a

selection of setups, including the Kolsky (Split-Hopkinson) pressure/torsion bars, Taylor impact

experiments, explosive tubes, and ballistic field tests [52,64]. Furthermore, various sample

geometries can be used (e.g. top-hat [60]) to induce shear in predictable regions with

predetermined total strain and strain rates. Each can be paired with high-speed cameras and

thermal sensors to determine global strain, local strain, local temperature, and localization

progression as a function of time. For further reading on experimental setups, refer to Ramesh

[64], Walley [52], and Lindholm [65].

Accurately characterizing local strain, temperature, and microstructural evolution during

high strain-rate deformation is difficult due to the small scale and the rapid nature of shear

localization. Shear bands propagate between 250-1200m/s [44,46,48], and the most developed

portion of the band can be as narrow as 5µm [53]. Experimental work to characterize the

sequence of events and prevailing conditions in the shear band frequently utilizes high-speed

photography and in-situ infrared imaging but is limited by small time-windows and spatial

resolution. Nevertheless, the information obtained from such experiments provides insights into

rationalizing the post-deformation microstructure in the ASBs, for understanding microstructure

Page 45: Microstructural Evolution during Heat Treatment and High ...

29

evolution, and is useful as experimental input for modelling high strain rate deformation and

shear localization.

High-speed cameras have been used to estimate local strain and ASB propagation speed.

Using painted horizontal stripes on a torsional Kolsky sample of HY-100 steel, Marchand and

Duffy [44] were able to measure local shear strains ranging from 500-1900%; they noted that the

strain was inhomogeneous during band formation and during propagation, likely due to

perturbations of the thickness of the cylinder walls [44]. Therefore, they concluded, there is a

strong relationship between geometry and shear band propagation. In addition, they estimated the

propagation speed to be around 500 m/s, though possibly 250m/s, if it propagated along two

fronts. Zhou et al. [46] and Guduru et al. [48] measured with higher confidence a propagation

speed up to 1100 m/s in C300 steel.

Local temperatures as a function of local strain within shear bands have also been

estimated with increasing accuracy through the use of high-speed cameras in conjunction with

arrays of infrared detectors. Marchand and Duffy and Giovanola each measured the increase in

temperature in the shear band in HY100 steel and found it to be between 550-1000°C, although

they were limited by the spatial resolution of their detectors [44,45]. With slightly improved

resolution, Zhou et al. showed that in a C300 maraging steel, the temperature was elevated in a

region between 200-300µm, much larger than the band itself, with a 1400°C maximum increase

in the most concentrated portion of the band [46,47]. Guduru et al. performed similar

experiments on C300 and measured a lower bound of 600°C, but importantly noted that ASBs

are not thermally homogeneous: shear bands can form periodic ‘hot-spots,’ and therefore may

have quite varied internal structure [48]. From a theoretical approach, temperature increases can

Page 46: Microstructural Evolution during Heat Treatment and High ...

30

be estimated using the constitutive equations of the Zerilli-Armstrong model [66], for example,

as Kad et al. have done with ultra-fine grain zirconium [67].

Historically, shear bands have been categorized into two types: transformed bands and

deformed bands [52,53,63,68]. Transformed bands were observed in materials that were capable

of phase transformation upon heating and rapid cooling and were thought to have undergone a

cycle of phase changes during the rapid temperature spike during ASB formation and the rapid

heat extraction that follows as the heat is dissipated to the surrounding material (e.g. ferritic

steels up-quenching to austenite in the band from the temperature spike and then transforming to

martensite following the rapid temperature drop). However, actual evidence for temperature-

induced phase transformation was largely circumstantial and relied on macroscopic observations;

transformed bands were resistant to etching and have been referred to as ‘white bands’ and they

had a measurable hardness increase across the band, which was used as evidence for a phase

change. On the other hand, deformed bands did not etch a different color, and thus were thought

to have had no phase transformation. This historical perspective has been demonstrated to be not

entirely correct: ‘deformed’ bands can in fact have phase transformation, and ‘transformed’

bands may have none at all [51,52,57,61,68]. The increased hardness previously used as an

identifier of a transformed band is now believed to be due to substantial grain refinement

processes [51,57]. As localization continues, shear bands continue to narrow [44,45,58,69]; it is

therefore thought that ‘deformed’ bands may be those that have not progressed so far as to

mechanically recrystallize [57,70].

The mechanical properties and microstructures of Cu, Ta, Al-Li, Ti, and steel are all quite

different, but share a common microstructural feature within the shear band in that they all have

equiaxed grains between 100-300nm in diameter [51,71]. It is thought that each undergoes a

Page 47: Microstructural Evolution during Heat Treatment and High ...

31

similar microstructural process during shear localization: first, a plastic elongation in the shear

direction; next, a breakup of these elongated grains into sub-grains separated by dislocation cell-

walls; lastly, a rotational deformation mechanism that results in a fine-grained equiaxed structure

with low dislocation density [51,57,67,71,72]. This final step of ‘recrystallization’ is thought to

be facilitated by the high temperatures in the shear band, which allows for the small-scale

diffusion necessary for the subgrains to rotate. Note that this model does not require phase

transformations; indeed, in some steels there is strong evidence that no up-quenching/re-

quenching of martensite occurs within these bands [57,58,69,73]. The increased hardness and

white etching seen in these bands are thus attributed to a well-defined region in which the

average grain size is significantly smaller than the edges of localization where grains have not

refined but are elongated by intense shear [57,58,69]. Deformed bands, therefore, are those that

have not undergone this dynamic recrystallization process. However, it is worth noting that

mechanically-induced phase transformations (TRIP) may still occur within the shear band (e.g.

austenite to martensite) [51,57].

Two studies in particular clearly demonstrate this dynamic recrystallization process. Xue

and Gray observed a grain-refinement process in austenitic 316L stainless steel [58,69]. The steel

was deformed using compression Kolsky-bars on top-hat specimens, which were loaded with

strain pulses of varied duration but equal magnitude to produce various stages of shear band

propagation and growth under similar loading conditions. TEM specimens were cut and thinned

from the sheared area for further analysis. The initial stages of shear localization were marked by

dislocation cell structures and a lath-like microstructure elongated along the shear axis. In a well-

developed band sample, there were three distinct microstructural regions with respect to the

band. At the furthest edges of localization, there was a ‘dislocation-avalanche’ structure, where

Page 48: Microstructural Evolution during Heat Treatment and High ...

32

the dislocation cell structures were annihilated parallel to the elongation but grew into thick

tangles at the long ends. Closer to the band, an elongated subgrain region was observed, where

the grain showed a very high aspect ratio along the shear axis (~30). This elongation mechanism

was believed to be facilitated through the activation and multiplication of twins that were along

preferential orientations. Lastly, the center of the shear band was marked by a fine equiaxed

subgrain structure with diameters ranging from 20-100nm. This region was about 10µm in width.

Thus, the full microstructural evolution during shear localization was characterized: first, grains

break into dislocation cells; the cells elongate and further refine; and finally break down into

equiaxed subgrains.

This model was theoretically rationalized by Kad et al. [67] in ultrafine-grain zirconium.

Specimens were cut in the top-hat geometry and deformed using compression Kolsky-bar

experiments; global plastic strain was constrained by changing the protruding hat height. As in

the work of Xue and Gray, they observed a sharply contrasted microstructure within the center of

the shear band denoted by fine, equiaxed grains. To determine if such a dynamic recrystallization

process was feasible (specifically, if equiaxed subgrains could rotate during deformation to form

the high misorientation boundaries seen using SADP and darkfield), they used the Zerelli-

Armstrong model [66] and experimental stress/strain curves to determine the temperature rise in

the band. The temperature was predicted to increase by over 600°C to become about 40% of the

melting temperature. They then estimated the time it would take for enough grain boundary

diffusion at that temperature to rotate hexagonal subgrains of diameter 100-300nm by 30°, and

found it to be between 50-100µs. This time falls within the range of shear band development,

thus reinforcing the hypothesis for dynamic recrystallization.

Page 49: Microstructural Evolution during Heat Treatment and High ...

33

The loss of load-bearing capacity of a material has been shown to be due to void

coalescence that often occurs during ASB propagation and growth [56,57,74]. Xu et al. [56] used

SEM analysis on low-C steel that was deformed using a torsional Kolsky-bar; by interrupting the

test at different stages, they were able to correlate the global stress-strain curve with the

microstructure at different stages of ASB propagation and growth. The sharp drop in load-

bearing capacity more strongly correlated with the formation and growth of micro-voids than

with ASB formation. Cracks were seen to initiate at interfaces, such as grain boundaries,

inclusions, and precipitates. A similar observation was made in Ti-6Al-4V in radially-collapsed

thick-walled cylinders [57]. In this material, ellipsoidal voids were seen to form in the thermally

softened shear band and grow to the edges of the band, creating a perforated structure. The

nucleation and growth of these voids correlated strongly with the loss of load-bearing capacity.

Rigorous microstructural analysis at various stages of deformation has not been

performed on 10Ni QLT and the work on its previous iterations is limited. In total, there are

three principle studies to consider.

Wang and Kumar characterized the microstructural evolution of the Fe-10Ni-1.0Mo-

0.6Mn-0.6Cr-0.08V-0.1C (wt.%) steel subjected to an unspecified heat treatment during quasi-

static and dynamic deformation [22]. The initial microstructure was composed of lath martensite

with a high dislocation density, no discernable retained austenite, and a dispersion of MC and

M2C carbides. High strain rate deformation with strain rates between 1.0x103s

-1 and 4.5x10

3s

-

1were performed on cuboidal specimens using a compression Kolsky-bar. A series of strains and

strain rates were produced, and shear bands were found to occur in samples deformed over 40%

and at strain rates over 2000s-1

. Nano-indentation arrays across the well-developed portion of the

shear band showed that the hardness increased from about 6.5GPa to about 7.5GPa across the

Page 50: Microstructural Evolution during Heat Treatment and High ...

34

width of the band. TEM liftouts from the localized region showed that outside of the band, the

laths were oriented along the shear direction, while the microstructure in the band itself was

comprised of fine, equiaxed grains. Coupled EDS and MDP showed that some of these equiaxed

grains had Ni content close to the nominal concentration yet were austenitic. If these grains had

been austenite in the undeformed microstructure, they would be predicted to be much more Ni-

rich by the Fe-Ni phase diagram. Therefore, their austenitic structure in the band suggests that

they were martensite that reverted to austenite from the high temperature spike during shear

localization, while their fine size was hypothesized as preventing them from transforming back

to martensite during the rapid cooling that followed. Also present were ferritic grains that

appeared to be highly twinned. There were no carbides found inside the band, either because

they were plastically sheared into very thin ligaments that were not recognizable or they re-

dissolved. Another specimen consisted only of ferritic subgrains (i.e. no up-quenched austenite),

suggesting that the temperature in the band was heterogeneous.

Gupta and Kumar examined the microstructural evolution during high strain rate

deformation of the Fe-10Ni-1.0Mo-0.6Mn-0.6Cr-0.08V-0.1C steel that was subjected to two

different heat treatments [29]. In both instances, the steel was austenitized at 840°C for 1 hour

and water-quenched. The first material was then tempered at 450°C for 5 hours, and the second

was tempered at 620°C for three hours, quenched, and then tempered at 540°C for one hour. The

samples were deformed using a compression Kolsky-bar at ~1x103 s

-1 and characterized using

TEM. Both steels initially had a lath martensite structure with MC carbides; the two-stage-

temper steel additionally had a lower dislocation density, M2C carbides, and thermally stable

precipitated austenite. The dynamic stress/strain curves for the single-temper specimen had a

higher yield than the two-stage-temper specimen but exhibited strain-softening indicative of

Page 51: Microstructural Evolution during Heat Treatment and High ...

35

severe shear localization. In contrast, the stress/strain curve from the two-temper specimen did

not have strain-softening, suggesting resistance to shear localization. The single-temper

specimen had well-developed shear bands and an SADP pattern with FCC spots, suggesting that

some of the lightly tempered martensite up-quenched to austenite within the band, similar to that

reported by Wang and Kumar [22]. The shear bands in the two-temper specimen were more

diffuse and so they introduced a notch in the specimen to generate a narrow, well-defined shear

band. TEM analysis using SADP showed no FCC spots suggesting that the retained austenite in

the initial structure had dynamically transformed to martensite during the process of shear

localization. Thus, it was concluded that a TRIP effect provided from precipitated austenite may

help dissuade shear localization.

As discussed earlier in this chapter, Zhang [25] performed numerous field tests to

determine the ballistic resistance of a series of alloy compositions and heat treatment

combinations and those results were presented in Figure 2.5. The 10Ni steel with the QLT

treatment that is the focus of this work was shown to have superior ballistic resistance as

compared to alloys with less Ni but subjected to the same QLT treatment, and other 10Ni alloys

subjected to various single and other multi-stage tempers. Inferior steels were found to form

ASB and fail via plugging; the 10Ni QLT steel was found to resist ASB formation and instead

deformation was distributed by plate bulging. The retained austenite volumes after heat treatment

and in the impact zone were measured using a vibrating sample magnetometer. While the

undeformed material had around 18 vol% austenite, the impact zone was found to have

negligible austenite. Zhang suggested that the mechanical transformation of this large volume of

thermally stable austenite served to suppress shear localization and instead dissipated the impact

energy over a much larger volume of material.

Page 52: Microstructural Evolution during Heat Treatment and High ...

36

2.6: Scope of this Effort

While the ballistic response of an Fe-10Ni-1.0Mo-0.6Mn-0.6Cr-0.08V-0.1C plate

subjected to the QLT heat treatment was clearly superior compared to the other variants tested,

the following fundamental questions remain inadequately unanswered:

1) What is the fine microstructure of 10Ni QLT? How did it evolve in the context of

composition, size, and thermal stability of the precipitated austenite and how do these

relate to the two-stage tempering process?

2) How does the austenite in 10Ni QLT mechanically evolve during high strain-rate

deformation? Specifically, does deformation-induced austenite to martensite

transformation occur in the dynamically deformed material and if it does, how

prevalent is it?

Thus, our experimental effort can be viewed as being composed of two sequential parts:

first, a study of the relationship between heat treatment and microstructure, focusing on

how precipitated austenite size, composition, and morphology relate to thermal stability;

and second, a comparative analysis of the microstructural evolution of 10Ni QLT during

high strain-rate deformation to determine if the thermally stable austenite is mechanically

unstable and if so, to what extent or under what circumstances.

Page 53: Microstructural Evolution during Heat Treatment and High ...

37

Chapter 3: Experimental Procedure

3.1: Materials and Heat Treatment Schedules

Table 3.1: Nominal alloy composition

Fe Ni Mo Mn Cr C V

wt.%

(Given)

87.6 10.0 1.0 0.6 0.6 0.1 0.08

at.% (Calc) 86.7 10.6 0.7 0.7 0.7 0.5 0.1

The nominal composition of the alloy examined in this study was Fe-10Ni-1.0Mo-

0.6Mn-0.6Cr-0.08V-0.1C (wt.%) (Table 3.1). This steel (referred to as the “10Ni steel” later)

was first melted in vacuum using induction heating, cast into billets that were then hot rolled into

plates, and air cooled; samples were then cut from the plate and heat treated. The focus is on the

QLT heat treatment that consists of three stages:

1) An austenitizing step at 800°C for 1 hour followed by a room-temperature water

Quench (the Q step)

2) A high temperature ‘Lamellarizing’ temper at 650°C for 40 minutes followed by a

water quench (the L step)

3) A low temperature Temper at 590°C for 60 minutes followed by a water quench.

To understand the effects of tempering time and temperature on microstructure evolution,

specifically austenite composition, size, dispersion, and thermal stability, we additionally

examined a series of single-stage heat treatments at the L temperature (650°C) and at an

exaggerated low temperature designated T’ (540°C). Specifically, as shown in Figure 3.1, we

Page 54: Microstructural Evolution during Heat Treatment and High ...

38

consider a 40 minute and 25 hour L treatment with water quench (QL and Q25L respectively)

and 5, 25, 125, and 336 hour T’ treatments with water quench (Q5T’, Q25T’, Q125T’, and

Q336T’, respectively). This range of single-temper treatments allows us to examine and

understand the kinetics of austenite precipitation at high and low tempering temperatures. This

also allows us to cumulatively piece the individual stages of QLT together (in the AQ, QL, and

QLT samples) to understand directly the effects of each stage on the final QLT microstructure.

Figure 3.1: Heat treatment schedules considered. All treatments begin with austenitization and quench (AQ). Two single-stage

tempering temperatures (650°C-L and 540°C-T’) with various lengths are analyzed, as well as the two-stage temper QLT

A large piece of QLT-treated plate was supplied by our collaborators at NSWC-CD

(Naval Surface Warfare Center-Carderock Division); in addition, fractured Charpy test

specimens of Q25L and Q5T’, Q25T’, Q125T’, and Q336T’were also provided. AQ and QL

samples were important to our understanding of heat treatment and therefore pieces were cut

from the QLT plate and re-heat-treated at Brown University to obtain the AQ and QL specimens.

Page 55: Microstructural Evolution during Heat Treatment and High ...

39

3.2: High Strain Rate Deformation

The experiments that focused on understanding microstructural evolution during high

strain rate deformation were limited to the material subjected to the QLT heat treatment. A

Kolsky (Split-Hopkinson) compression bar setup (Figure 3.2) was used to generate dynamic

stress-strain data, high strain-rate deformed microstructure, shear localization, and adiabatic

shear banding in the QLT specimens. Such experiments produce strain-rates that are comparable

to ballistic impact and can reliably produce predetermined strain and strain-rates. A brief

description of the experiment follows: the sample is placed between two long steel bars (the

incident bar and the transmitted bar); a variable length projectile bar loaded into a pressurized air

gun with controlled firing pressure can be impacted on one end of the incident bar. The resulting

strain pulse travels through the incident bar to the sample, and the transmitted and reflected

pulses are each recorded using an oscilloscope. These strain pulses, in combination with initial

and final sample dimensions, are then used to calculate the stress/strain response of the

specimen. Detailed explanations of Kolsky-bar setups, including underlying theory and

derivations, can be found in Ramesh [64], Walley [52], and Lindholm [65].

Figure 3.2: Kolsky-bar setup. The sample is sandwiched between two bars; a projectile is fired at the end of one bar, and the

strain pulses through the incident bar, reflected back through the incident bar, and transmitted through the transmission bar are

recorded with strain gauges. A stress/strain curve is thus derived from these pulses.

Page 56: Microstructural Evolution during Heat Treatment and High ...

40

Kolsky-bar experiments have only two input variables, projectile firing pressure and

projectile length, but can achieve a wide range of total strains and strain rates. Therefore, an

extensive amount of calibration is necessary to produce a deformed sample with a desired strain

and strain rate. A wide range of projectile and pressure combinations were thus initially

examined using 4140 tempered martensite to calibrate the Kolsky-bar and to hone in on the

range of parameters that was considered appropriate to achieve the desired strain rate and strain

in the QLT-treated 10Ni-steel samples. 4140 steel was chosen due to its availability, ease of

machinability, and the similar range of static mechanical properties typical of the QLT-treated

10Ni steel. To determine the effects of bar length on strain and strain rate, 15.3cm, 20.3cm,

30.5cm, and 40.6cm projectile bars were used. A wide range of pressures (20-100psi) was used

for each bar that yielded velocities ranging from 8-36m/s. The strain and strain rates for each run

were plotted as functions of the bar and loading pressure used, thus providing a rough estimate of

experimental parameters needed for testing the QLT-heat treated steel.

A few sample geometries were considered, including cuboidal samples, cylindrical

samples, and notched samples. Samples with a cylindrical geometry were selected due to their

radial symmetry that helps maintain experimental consistency. Notched samples were briefly

used because the notch served as an initiation point for shear localization; however, experiments

on un-notched samples were also able to produce shear banding, and so notches were determined

to be an unnecessary complication in sample uniformity.

Cylindrical samples with a diameter of 6.35mm and a height of 9.52mm and subjected to

the QLT heat treatment were cut from bar stock using electro discharge machining (EDM). The

anvils and bars were both of hardened steel and had a diameter of 2.54mm; the projectile bar was

also a steel bar and 30.5cm long. Projectile velocities ranged from 13m/s to 27m/s (which

Page 57: Microstructural Evolution during Heat Treatment and High ...

41

equates to chamber pressures ranging from 47-95psi). Strain and strain-rate were varied on the

QLT specimens to capture different stages of shear localization, including: homogeneous

deformation (i.e. no localization), localization but no developed shear bands, well-developed

shear bands that decayed within the sample, shear bands that ran through the sample, and

fractured samples. Deformed samples were cut in half vertically with EDM or slow-speed

diamond saw and further prepared for characterization.

3.3: Microstructure Characterization

3.3.1:Optical Microscopy:

The microstructure of both undeformed and deformed samples was too fine to fully

define with optical microscopy, though optical microscopy was useful to each of these two

categories for specific reasons. Metallographic samples were prepared by grinding with SiC

paper, polishing with 1.0µm, 0.3µm, and 0.05µm alumina solution, and etching with a solution

of 2-5% nitric acid in methanol. One use for optical microscopy of the undeformed samples was

for analyzing banded structural features, such as those observed by Zhang [25]. During hot

rolling and subsequent treatment, alloying elements may segregate to alloy-rich and alloy-poor

bands, which influences the precipitation of austenite during tempering, and while these banded

features were also observed in SEM, low magnification light microscopy was also used to assess

these features on the millimeter scale. This was most evident in Q5T’, the sample with the least

amount of austenite precipitation. The primary purpose for using optical microscopy to

characterize the deformed specimens was to determine the presence and development of

adiabatic shear bands in deformed QLT samples. Shear bands on etched samples appear as white

bands, and so they can be clearly seen running through the sample if present. Multiple deformed

samples were cut in half to determine the progression of shear bands within, and this information

Page 58: Microstructural Evolution during Heat Treatment and High ...

42

was used to iterate and tune the Kolsky-bar experiments. Furthermore, the presence of a banded

structure in the millimeter scale (described above) in these materials enable an estimation of

localized strain, as they serve as markers and their displacement across a shear band are can be

readily observed. Lastly, optical microscopy acted as a screening tool to isolate suitable

specimens for further analysis using electron microscopy techniques.

3.3.2: Scanning Electron Microscopy (SEM):

SEM was used to qualitatively compare the size and dispersion of austenite as a function

of heat treatment schedules as well as to guide smaller-scale analysis with TEM and APT. SEM

samples were prepared in the same fashion as optical microscopy samples: ground with SiC

paper; polished with 1.0, 0.3, and 0.05µm alumina; and etched with a solution of 2-5% nitric acid

in methanol. Secondary electron and back scattered SEM images were obtained for all heat

treatments to compare differences in: precipitation across the rolling bands, morphology of

precipitated austenite (e.g. spherical, lamellar, etc.), and variations in austenite size. In deformed

material, SEM was used to examine austenite morphology as a function of proximity to the shear

band, as well as for targeted TEM liftouts, which is discussed later in this section in more detail.

The two microscopes used were a LEO 1530VP and a FEI Helios equipped with electron

backscatter diffraction (EBSD) and focused ion beam (FIB) systems.

Electron backscatter diffraction was attempted to quantify the amount of thermally stable

austenite in given heat treatments. Several sample preparation techniques were utilized,

including mechanical polishing with alumina down to 0.05µm; polishing with 1µm alumina

followed by electropolishing in a 10% solution of perchloric acid in acetic acid; and polishing

with a 1µm alumina polish followed by a polycrystalline diamond polish, but the outcome was

Page 59: Microstructural Evolution during Heat Treatment and High ...

43

not satisfactory. TEM provided results with a higher point density and reliability than we could

have hoped for using EBSD, so this aspect was not pursued.

3.3.3: Transmission Electron Microscopy (TEM):

Transmission electron microscopy was extensively used to follow the composition,

structure, and size of the phases present as a function of the various heat treatments examined in

this effort, as well as to examine the microstructure of the dynamically deformed specimens. For

undeformed materials, samples were cut from the bulk into small plates using EDM and

mechanically thinned to widths between 50-100µm by grinding with SiC paper. They were then

polished using 1µm alumina and electrochemically thinned to perforation using a Tenupol 5

twin-jet polisher at 20-25V at -30°C with a solution of 20% nitric acid in methanol. For

deformed QLT specimens, FIB liftouts were made with specific geometries and at distinct

locations with respect to the adiabatic shear band including: far from the band (homogeneous

deformation zone), just ahead of the band (some localization without well-defined ASB), parallel

to the well-defined band, at an angle to (but in) the band, and across the band normal to the shear

band plane. Collectively, these various lift-outs provided snapshots of the microstructure at

various stages of localization.

Initial TEM analysis was done using a Philips CM20 equipped with an Oxford energy-

dispersive X-ray spectrometry (EDS) detector; finer-scale analysis was done using a JEOL

2100F in scanning mode (STEM) with an equipped Oxford EDS detector

Electron diffraction was used to determine the structure/identity of the phases in these

multiphase microstructure (e.g. if a precipitate is thermally stable austenite versus martensite) as

well as to examine texture (or lack thereof) across larger polycrystalline regions (e.g. randomly

oriented). For most treatments, even the smallest selected-area apertures were much larger than

Page 60: Microstructural Evolution during Heat Treatment and High ...

44

individual austenite precipitates, and so selected area diffraction (SADP) was not useful to

determine their thermal stability. However, SADP in conjunction with darkfield (DF) imaging

was useful for comparing the degree of misorientation in undeformed microstructure versus

various stages of shear localization. The primary diffraction technique used to characterize

individual austenite precipitates and martensitic laths was microdiffraction (MDP) - condenser

apertures and spot sizes were reduced until the converged beam was smaller than the feature size,

and a diffraction pattern was taken.

Table 3.2: Example table of calibrated diffraction radii

Ring Calibrated Ratio Lattice Index

1 2.70 1.00 FCC 111 2 2.77 1.00 BCC 110 3 3.15 1.17 FCC 200 4 3.88 1.41 BCC 200 5 4.44 1.65 FCC 220 6 4.73 1.71 BCC 112 7 5.18 1.92 FCC 311

8 5.40 2.00 FCC 222 9 5.51 2.00 BCC 220

10 6.13 2.22 BCC 310 11 6.24 2.31 FCC 400 12 6.69 2.43 BCC 222 13 6.79 2.47 FCC 331 14 6.98 2.58 FCC 420 15 7.26 2.63 BCC 321 16 7.65 2.83 FCC 422 17 7.76 2.82 BCC 400 18 8.10 3.00 FCC 511, 333

19 8.10 3.00 BCC 330 20 8.21 2.98 BCC 411

Patterns were indexed, including measuring diffraction radii and angles. A calibration

table was created for FCC and BCC Fe (austenite versus ferrite and martensite), which was then

used to calculate all diffraction radii for both crystal structures (an example given in Table 3.2).

Page 61: Microstructural Evolution during Heat Treatment and High ...

45

When a crystal was on a zone axis that had similar symmetries for both FCC and BCC (e.g. BCC

and FCC [100] both have a square pattern,), the calibrated table was used to isolate the phase

through the difference in lattice spacing. The table was recalibrated for each individual session to

ensure no mistakes were made.

Lastly, these steels are highly magnetic, which makes tilting the specimen difficult as

significant image shifting occurs. Therefore, most regions of interest were found by looking for

grains already on a zone axis, rather than through using diffraction to tilt to a particular grain's

zone axis. Usually, this meant doing characterization at 20°, the tilt required for EDS analysis.

Very few tilting experiments were performed.

EDS was used extensively in this effort. In undeformed material, single-point EDS was

used to measure Ni content in ferrite and precipitated austenite, allowing for tracking the Ni

content in the precipitated austenite as a function of tempering temperature and time, as well as

understanding the thermal stability of this phase as a function of size and Ni level. EDS was also

used to measure the distribution of Ni by compounding hundreds of single-point measurements

over a large area. These distributions showed on a larger scale the Ni depletion in ferrite as a

function of tempering time and the Ni content approaching the equilibrium phase diagram values

for longer times.

In deformed QLT samples, Ni content served as the link to the undeformed

microstructure. As shear localization during dynamic deformation is essentially a diffusionless

process, a feature with 15 wt.% Ni in the deformed microstructure must have evolved from a 15

wt.% Ni entity in the undeformed state. In the case of QLT, this meant that the information

obtained on the undeformed microstructure could be mapped onto the deformed microstructure

through its Ni content. Thus, by examining multiple stages of shear localization, we could

Page 62: Microstructural Evolution during Heat Treatment and High ...

46

understand the microstructural evolution in the baseline QLT material during dynamic

deformation.

Figure 3.3: a) an example of EDS analysis of QLT using a manually condensed beam (rough size marked by pink circles) in the

CM20. Austenite precipitates include roughly one datum point each. b) STEM EDS on the JEOL 2100F allows for a high enough

data point density to create linear gradient plots within grains

In the early stages of this study, the CM20 was used to measure Ni content with

converged beam EDS. However, the smallest possible spot size was sometimes larger than the

features we sought to characterize- this was especially the case for shear band samples, where

features could be as small as 20nm. Therefore, EDS on the JEOL in the STEM mode was

employed instead. STEM mode has a rastering 1.5nm spot that creates an image, far smaller than

the features of interest, and by selecting points on the interface, the microscope focuses the beam

Page 63: Microstructural Evolution during Heat Treatment and High ...

47

automatically to the desired location. This allowed the number of data points taken, specifically,

the density of data points, to increase from ~20 points per session to ~1000 points per session

(Figure 3.3). STEM EDS thus allowed for two particularly useful analyses: arrays of ~200 data

points across ~2µm2 used to create composition distributions (as mentioned above), as well as a

high enough point density within individual austenite precipitates to create linear gradient plots.

It is worth noting that while automated rastered compositional mapping is possible using STEM

EDS, it was not used in this study due to significant sample drift issues.

STEM EDS analysis however does have its own set of challenges. While it was possible

to create data sets with hundreds of data points, it was difficult to perform the analysis manually.

STEM images needed to be overlaid on BF images, the data points had to be manually placed

(including compensating for drift during the EDS collection), and then the data points had to be

colored according to composition and a binning scheme. Single images would take days to

produce, and it did not allow for secondary analysis like linear gradient plots; adjusting color

schemes from different data sets to maximize contrast was also not feasible. Therefore, a Matlab

script was written and developed to help analyze the large EDS data sets.

The main purpose of the Matlab script (Figure 3.4) was to produce useful figures from

the EDS data sets, including performing secondary analysis such as linear gradient plots.

Experimentally, each individual collection of points was related to three STEM images: an image

taken before data collection, the EDS interface overlay of the ‘before’ image with the data points

locations marked, and an image taken after the data were collected (used to calculate drift during

data collection). First, points were manually selected in order on the second image, which saves

the location of the points in STEM space; simultaneously, the script tags each data point with the

corresponding composition data. Next, the script calculates the total drift (in STEM space) using

Page 64: Microstructural Evolution during Heat Treatment and High ...

48

built-in Matlab image correlation functions comparing the before and after images. The total drift

is linearly interpolated and then applied to each datum point; accuracy in the linear estimate was

mitigated through keeping the individual data sets small enough that they can be collected before

significant drift has occurred.

Next, the STEM image is mapped to the BF image using manual feature mapping- a point

in STEM mode is selected, and then the corresponding point in BF is selected. After several pairs

have been created, a built-in Matlab script determines a best-fit linear transformation from

STEM to BF using the pairs. In order to ensure the map meets a threshold of quality, a

quantifiable metric is assigned to it that describes the deviation between the manual pair

selection and calculated best fit map. The manually selected STEM points are mapped to BF

using the calculated affine transformation and then subtracted from their corresponding manual

BF points; the standard deviations of both X and Y components of these vectors are given, and if

unacceptable, the mapping process can be repeated from feature point selection.

Lastly, the EDS data points are mapped from STEM to BF and the final figure is created

by creating colored boxes on the locations of EDS measurements. The boxes are colored using a

preset color scheme based on their corresponding Ni content in order to visually compare Ni

distribution- for example, low-Ni points may be blue while high-Ni points may be red. The

dimensions of the boxes are the error bars for the point’s exact location, and are a function of

total drift during its particular set’s collection and the standard deviation metric from its

STEM/BF map. As the primary focus of the EDS work related to austenite, ferrite and

martensite, the Mo content was used to filter out data points from carbides as most carbides were

Mo-rich - a Mo threshold was entered, and all data points with Mo above that threshold are not

displayed.

Page 65: Microstructural Evolution during Heat Treatment and High ...

49

Figure 3.4: STEM mode images are mapped to BF using manually selected correlative point pairs- i.e. a location is selected on

the STEM image, and then the same point is selected in BF. Several of these are used to create an affine map. Data is collected in

small sets to avoid significant drift during collection; these sets are uploaded in STEM space, mapped to BF mode, and color

coded by composition. Dozens of individual sets create dense composition maps.

These figures can be made of dozens of data collection sets, each with dozens of

individual EDS measurements; collectively, hundreds of data points can be represented in a final

figure. Because the data are all stored after entry, including position, mapping, and composition,

these figures can be quickly reconfigured to new color binning if so desired (for example, to

maximize contrast between two different treatments). Lastly, linear gradients can thus be

calculated: a gradient box is drawn by manually drawing a line (which will be the plot axis) and

Page 66: Microstructural Evolution during Heat Treatment and High ...

50

providing a width. All EDS data points that intersect the gradient box are then plotted along the

length of the line (with corresponding horizontal error bars related to the EDS point error). An

example of this analysis can be seen in Figure 3.5. By creating a script to analyze raw STEM

EDS data, not only were the data more quickly synthesized, but secondary analysis could also be

done.

Figure 3.5: Sample EDS Ni gradient on a Q336T’ austenite precipitate

3.3.4: Atom Probe Tomography (APT)

In addition to EDS, atom probe tomography (APT) was used to measure composition

adjacent to interfaces as well as track the partitioning of minor alloying elements (Mn, Cr) and

light elements like C as a function of heat treatment. APT volumes are relatively small (e.g.

200nm x 50nm x 50nm) but are reliable down to a fraction of an atomic percent. Therefore APT

is a powerful tool when measuring trace element composition and composition at locations of

Page 67: Microstructural Evolution during Heat Treatment and High ...

51

interest that are tens of nanometers in size such as fine carbides and locations adjacent to

interfaces. By combining APT results with extensive SEM and TEM analysis, a full-scale picture

of microstructure as a function of heat treatment was thus developed.

Our access to APT was limited (APT was performed at the Max Planck Institute für

Eisenforschung in Düsseldorf, Germany), and so the samples examined were carefully chosen.

The QLT steps (AQ, QL, and QLT) were analyzed to provide a comprehensive picture of the

effects of each heat treatment step on the final microstructure. For single stage treatments, Q25L

was chosen to contrast with QL, and Q25T’ and Q125T’ were chosen to represent the T’

treatment.

Tips were extracted from samples ground with SiC, polished with 1µm alumina, and

vibratory polished with 0.06µm colloidal silica. Tips were sharpened using standard focused-ion

beam techniques in a plasma FIB [75–77]. APT was performed using a LEAP 3000 in laser

mode (wavelength = 532 nm) at 75K with 200 kHz 0.5 nJ pulses, and tips were reconstructed

using IVAS software (version 3.8). Analysis of APT reconstructions included composition

measurements within selected volumes, compositional gradients within single-grain regions, and

composition gradients across interfaces. Phase identification was by correlating Ni content with

those determined through TEM EDS+MDP (e.g. regions above 10 wt.% Ni in tempered samples

are precipitated austenite, while regions below 10 wt.% are ferrite). Due to the difficulty with

phase identification with APT techniques, austenite stability was assumed to be what was seen in

TEM (i.e. TEM MDP results were not confirmed or refuted using APT). Importantly, APT

allowed the rest of the chemical makeup of austenite and ferrite, including C and trace elements,

to be determined. Secondarily, it also allowed study of carbides evolution as a function of heat

treatment, as carbides were too small to be characterized with TEM techniques. APT results were

Page 68: Microstructural Evolution during Heat Treatment and High ...

52

complementarily contrasted with EDS results to create a complete picture of composition of

different phases as a function of heat treatment.

Figure 3.6: A pictorial summary of the experimental process utilized in this effort to characterize the undeformed microstructure

as a function of heat treatment.

3.3.5:Nano-Indentation

Nano-indentation experiments were performed on deformed QLT specimens to measure

hardness changes across the shear band. Indents were performed using a Hysitron nanoindenter

with a Berkovich tip and analyzed with Triboscan software; a maximum force of 5000µN (load-

controlled) was applied with a loading time of 250s, held at max load for 5s, and unloaded for

250s. Linear arrays of indentations spaced 2µm apart were obtained perpendicular to the band on

well-polished samples. Exact distance from the band for each indent location was calculated by

etching the sample to reveal the band, imaging the band using SEM, and physically measuring

the shortest distance from the indent to the center of the band.

Page 69: Microstructural Evolution during Heat Treatment and High ...

53

Chapter 4: The Partitioning of Carbon During the Heat Treatment

of Quenched Fe-10Ni-0.1C Steel

4.1: Introduction

C and Ni are both austenite stabilizers and both play a role in affecting the thermal

stability of austenite in Fe-Ni-C steels subjected to quench-and-temper type heat treatments.

While the role of Ni on austenite stability has been examined to some extent [25,27,28], the

influence of C partitioning on austenite formation and stabilization in this alloy or the associated

kinetics have not been examined and may be important. Indeed, C is the primary austenite

stabilizing partitioning element in many TRIP steels [37,41,78–85].

The quantification of C in austenite and ferrite in this steel as a function of tempering is

complicated by the fact that C cannot be quantified using the TEM-EDS techniques that can be

used for Ni (due to the characteristic X-ray energies of C overlapping with the detector

background) and that there are two competing mechanisms for C during tempering- the

partitioning to austenite versus the formation and growth of refractory carbides such as MC and

M2C. Additionally, the solid-state transport mechanism for C is very different from Ni, as C is

interstitial while Ni is substitutional. Therefore, the influence of C on austenite formation and

thermal stability during tempering is likely temporally decoupled from the effect of Ni. Due to

the rapid diffusion of C in Fe relative to Ni, we focus on the redistribution of C from martensite

in the early stages of tempering and its likely influence on austenite precipitation rather than on

subsequent growth [81,85,86].

Quantification of C in the phases present (austenite, ferrite, and carbides) as a function of

heat treatment was performed using APT. The AQ, QL, Q25L, QLT, Q25T’, and Q125T’

Page 70: Microstructural Evolution during Heat Treatment and High ...

54

specimens were all examined by APT, and the results from these specimens are presented in this

Chapter and their implications are discussed. (Lastly, the reader is reminded that in this alloy

system, Fe and Ni content numbers are similar when reported either in atomic percent or weight

percent whereas in the case of C, the atomic percent value is roughly five times that in weight

percent (see Table 3.1). The reason this is mentioned is because EDS data are in weight percent

whereas APT results are in atomic percent and C is only detected using APT and not by EDS).

4.2: Results: L Temper

The samples that best represent C partitioning as a function of tempering time are AQ,

QL, and Q25L. AQ is the original state or the reference state; QL includes the shortest temper

time of all heat treatments and is relevant not only because C transport is rapid at the L

temperature (650°C), but also because it is the first step of the multi-step QLT treatment. Q25L

provides a longer-term isothermal comparison to QL and is anticipated to be long enough that C

partitioning is effectively complete. Therefore, for the purposes of analyzing C behavior as a

function of tempering time, we consider AQ, QL, and Q25L treatments.

Bright field TEM imaging of the specimen following the AQ treatment confirmed a lath

martensite microstructure within which were dispersed a few fine second phase particles, ~20 nm

in diameter. These are highlighted in Figure 4.1a using dashed circles. The lath martensite

structure was verified by bright field TEM imaging coupled with microdiffraction, and EDS

analysis provided an average Ni content of 8.91.3 at.% which corresponds to the nominal alloy

composition in terms of Ni content. The highlighted second phase particles were too small to

reliably assess their chemistry by this technique and are believed to be undissolved carbides.

Microdiffraction confirms that some of these particles have a cubic structure and are thought to

be MC carbides that are known to be resistant to dissolution at high temperatures; MC carbides

Page 71: Microstructural Evolution during Heat Treatment and High ...

55

and M2C carbides have been previously reported in such steels [22,27–29]. The second phase

particle morphology and chemistry were further characterized by APT. Figure 4.1b shows a part

of an APT volume after reconstruction: the levels of Mo, V, and Cr in the alloy are low (less than

1 at.% of each element) and therefore, iso-concentration surfaces with a threshold of 2-3 at.% of

these individual elements are sufficient to highlight the carbides in the matrix. Thus, a 2 at.% Mo

iso-concentration surface in Figure 4.1b shows the location of discrete Mo-rich regions believed

to be carbide particles. Composition measurements of these carbides confirm they are Mo-rich,

although they are alloyed with significant levels of V and some Cr. Some carbides are seen to

additionally have moderate levels of Ti (~10 at.%), although Ti was not specified as an alloying

element and therefore is thought to be an impurity.

Page 72: Microstructural Evolution during Heat Treatment and High ...

56

Figure 4.1: a) TEM brightfield image of an austenitized and quenched (AQ) specimen showing a lath martensite microstructure;

some fine carbides (~20nm) are highlighted with dashed circles. b) APT composition measurement of a carbide shows

predominantly Mo but includes significant amounts of V and Ti. c) C segregation along interfaces in martensite.

Page 73: Microstructural Evolution during Heat Treatment and High ...

57

Retained austenite was not immediately evident in the microstructure although it is

entirely possible that it is indeed present in small quantities as films along prior austenite grain

boundaries, between packets and blocks of martensite as well as between laths, as recently noted

[34,35,87]. In the martensite, APT permits measurement of C in solid solution and any C

segregation that might have occurred during the quench. Thus, the measured average C

composition in the martensitic matrix was 0.230.10 at.% whereas the bulk C level in this alloy

was 0.5 at.%; this implies that roughly half of the C in the alloy is tied up in the carbides that

remained undissolved during the austenitizing treatment. In addition, C segregation/enrichment

is observed in Figure 4.1c (highlighted using a dashed rectangular box) at what is likely an

interface (prior austenite grain boundaries and/or martensite packet/block/lath boundaries) with

peak values of around 2-2.5 at.%, which is an order of magnitude higher than the average C level

in the martensitic matrix. Similar levels of interfacial segregation of C was observed in multiple

locations.

The L tempering of a quenched 9.6 at.% Ni alloy (40 mins at 650°C - the QL treatment)

places it in the two-phase region of the Fe-Ni phase diagram, initiating austenite precipitation

and the transformation of martensite to ferrite. While the solubility of C in ferrite is negligible,

the precipitating austenite can serve as a reservoir for C partitioning out of the martensite

together with additional carbide precipitation. Bright field TEM imaging coupled with

microdiffraction (Figure 4.2a) and chemical analysis by EDS confirmed noticeable Ni

partitioning during tempering into relatively Ni-poor (7.1-9.5 at.%) ferrite and Ni-rich austenite

(16-17 at.%). This new generation of austenite particles is small (50-250nm) and Ni-rich, both of

which contribute to its thermal stability after a water quench [28]. Synchrotron measurements by

Jain et al. [28] estimated the austenite volume fraction to be 8.1% following this heat treatment.

Page 74: Microstructural Evolution during Heat Treatment and High ...

58

The measured Ni level in ferrite following this L temper is higher than what the phase diagram

predicts (~5 at.%), implying that the ferrite is still supersaturated in Ni due to the relatively short

tempering time and the slow diffusion of Ni.

Figure 4.2: a) TEM EDS/MDP shows Ni-depleted ferrite of about 8 wt.% Ni and Ni-rich (15-17 wt.%), thermally stable

austenite precipitates. b) APT composition measurement of a carbide particle in the QL specimen. A Ti- and V-rich core is

surrounded by a Mo-rich but Ti- and V-depleted outer shell (delineated by dashed lines).

APT confirms the ferrite phase regions far from, as well as immediately adjacent to the

austenite interface to be almost completely depleted of C (0.010.01 at.% and 0.040.01 at.%,

Page 75: Microstructural Evolution during Heat Treatment and High ...

59

respectively). The austenite, however, has a C level of 0.290.06 at.% - the solubility of C in

austenite is quite high, and the C is adequately mobile to migrate to it. Unlike the AQ sample, no

C-enriched planes were observed in the APT specimens in this condition; however, there are

many carbides observed in both austenite and ferrite and at the interphase interfaces. Some of the

carbides observed have a core composition that is different from the shell that forms around the

core; in the example shown in Figure 4.2b, a Ti-rich carbide is encased in a Mo-rich outer layer.

There is a noticeable decrease in Mo content in both the ferrite and austenite phases versus the

corresponding value measured in the martensite in the AQ condition (0.260.06 at.% and

0.260.06 at.% versus 0.520.10 at.%), implying that some of the Mo has being depleted from

solid solution to form new carbides.

The Ni distribution profile in the Q25L specimen measured by EDS is also bimodal with

Ni-poor regions (4.80.6 at.%) confirmed by microdiffraction to have a bcc structure (ferrite)

and relatively Ni-rich regions (12.81.5 at.%) that also yielded bcc microdiffraction patterns

(Figure 4.2a), indicative of freshly formed martensite following the water quench from 650°C.

Evidently, the Ni level in these Ni-rich regions is inadequate to stabilize the precipitated

austenite. The somewhat lower Ni level in the Q25L fresh martensite (that was austenite prior to

the quench) relative to the stable austenite in the QL treatment is thought to be related to the fine

austenite particle size in the latter and the related interface curvature-enhanced Ni solubility.

Page 76: Microstructural Evolution during Heat Treatment and High ...

60

Figure 4.3: a) TEM EDS/MDP show the austenite precipitates are moderately Ni-rich (15 wt.%) while the ferrite is close to

equilibrium value of 5 wt.%. The precipitates are thermally unstable. b) Interface segregation of C in fresh martensite in the

Q25L specimen. c) APT measurement of the composition across a carbide particle shows a V-rich, Mo-poor carbide is

sandwiched by Mo-rich, V-depleted carbide.

APT shows that after 25 hours of heat treatment, C is depleted in ferrite and is negligibly

low in the fresh martensite (0.040.01 at.%) as well. Nevertheless, within this fresh martensite,

C-enriched regions (0.30.01 at.%) were present, suggesting some C-segregation to interfaces

(highlighted by the rectangular dashed box in Figure 4.3b), although considerably lower in

concentration than that observed in the AQ condition in Figure 4.1c. The freshly formed

martensite and the ferrite have Mo levels (0.330.06 at.% and 0.330.06 at.%) that are similar to

those observed in QL (0.260.06 at.% and 0.260.06 at.%), suggesting that the carbides have

completely precipitated out during the first 40 minutes of tempering. This would imply that the

Page 77: Microstructural Evolution during Heat Treatment and High ...

61

decrease in the average C level in the freshly formed martensite in the Q25L specimen is

primarily a dilution effect resulting from the increase in the precipitated austenite fraction due to

the long tempering time of 25h. Compositionally, the carbides present in Q25L are similar to

those seen in QL: primarily Mo-based, with some showing a ‘core’ of V or Ti. In Figure 4.3c, a

large V-carbide (with almost no Mo) surrounded by patches of Mo-rich carbide is observed. In

this core region where the V level is ~50 at.%, the C level is around 33 at.%. Carbon is known to

be challenging to quantify by APT [88–91]. The level we report are likely an underestimation of

the actual C level in the carbide as such a V-rich carbide is most likely an MC carbide (V4C3)

rather than an M2C carbide as suggested by the chemistry. Thus, APT-based chemistry alone is

not necessarily adequate to distinguish between MC and M2C carbides as both are expected to be

present. The outer patches of the Mo-rich carbide are more likely M2C carbide type that

precipitates during tempering.

4.3: Results: T’ Temper and QLT

Table 4.1: C content in austenite and ferrite in Q25T’, Q125T’, and QLT as measured by APT (at.%)

Q25T' Q125T' QLT

Ferrite 0.01% 0.00% 0.00%

Austenite 0.01% 0.00% 0.05%

The C content in the austenite and ferrite in Q25T’, Q125T’, and QLT as measured by

APT are shown in Table 4.1. Of the three treatments, QLT has the shortest total tempering time

of 1 hour and 40 minutes, and while some C remains in solid solution in the austenite in this

state, it is significantly lower than that observed in QL austenite (0.290.06 at.%). While one

reason for this decrease is a dilution effect from additional austenite growth during the T temper,

some C is also likely stripped out of the austenite to form additional carbides. Thus, the one hour

at the T temperature of 590oC is adequate to remove all C from austenite and therefore, we

Page 78: Microstructural Evolution during Heat Treatment and High ...

62

conclude that C does not contribute to the austenite stability when quenched from the T

temperature to room temperature. The Q25T’ and Q125T’ samples had much longer tempering

times, and no measurable C remains in austenite.

4.4: Discussion

We now discuss and summarize these findings in the context of the role of C in

influencing the stability of the precipitated austenite. Following water quenching from 800°C,

roughly half the total C in the alloy remains tied up in undissolved carbides while the rest is

present in martensite. The trapped C in martensite shows a tendency to segregate at interfaces

where its local concentration is almost ten times as high as within the martensite laths. Upon

tempering for 40 minutes at 650°C, fine austenite particles precipitate at lath interfaces while

Mo-rich carbides precipitate on undissolved carbides and at lath interfaces. In addition, the

austenite particles are Ni-rich, the Ni content in them being enhanced by particle size and

interface curvature. In contrast, the adjoining ferrite (martensite gradually decomposes to ferrite)

is C-depleted but still Ni-supersaturated relative to phase diagram prediction. The C content in

the austenite, the high Ni content, and the fine austenite particle size together stabilize the

austenite at room temperature following water quenching from 650°C. It appears that 40 minutes

at 650°C is adequate to fully redistribute the previously trapped C in martensite. When the

tempering time is increased to 25 hours, austenite fraction increases primarily by the partitioning

of Ni between ferrite and austenite while the C level in this austenite decreases substantially as

no additional C is available. Meanwhile, the average Ni level in the austenite is not as high, as

the interlath austenite size increases and curvature decreases. The austenite particle size increases

as well. Together, these factors reduce the austenite stability and make it more susceptible to

Page 79: Microstructural Evolution during Heat Treatment and High ...

63

fresh martensite transformation on quenching following the 25 hours exposure to the tempering

temperature.

From the foregoing, it is evident that C does not play a major role in influencing the

thermal stability of precipitated austenite in longer tempering treatments. However, the migration

of C to the cores of dislocations that compose the lath boundaries during quenching and

enrichment of the austenite in early stages of tempering suggest that C could be important for the

nucleation and early stage growth of precipitated austenite during tempering. This early-stage

austenite precipitation can be austenite nucleation, or growth of nano-scale films of residual

austenite (which while we did not see but has been noted by other researchers on similar

materials [34,35,87]).

Thus, it is possible to envision a multi-stage austenite precipitation process from a

composition perspective. An initial stage at short tempering times (say 5 minutes to 30 minutes

at the tempering temperature of the QL material) where C diffusion stabilizes very fine austenite

particles while Ni diffusion lags and the Ni content in the austenite is less than what is predicted

by the phase diagram. In this early stage, as Ni diffusion in austenite is significantly slower than

in martensite, internal Ni concentration gradients may be present within the austenite [42]. A

second mixed stage is possible where Ni and C both stabilize austenite but the role of C

continually decreases as all available C is consumed in the first stage in the austenite and by

precipitation of refractory carbides. In this stage, austenite growth continues due to Ni diffusion

into the austenite, but the increasing volume fraction of austenite leads to C dilution in it. Beyond

this stage, Ni partitioning is all that matters.

Page 80: Microstructural Evolution during Heat Treatment and High ...

64

Chapter 5: The Partitioning of Ni During the Heat Treatment of

Quenched Fe-10Ni-0.1C Steel

5.1: Introduction

In the previous Chapter, it was shown that while C partitioning may be relevant in the

very early stages of the L temper in influencing austenite nucleation, its role diminishes rapidly

within the first one hour at that temperature. In this Chapter, we similarly examine the

partitioning of Ni during the L and T tempers in detail.

The L-temper precipitated austenite is not expected to be particularly Ni-rich according to

the binary Fe-Ni phase diagram (Figure 2.3b) and thus there remains the question whether this

austenite is thermally stable upon quenching after this temper. Our observations in the previous

Chapter showed that austenite in the QL condition is in fact stable whereas that is not the case in

the Q25L condition suggesting that size may play an effect in the thermal stability of austenite.

5.2: The As-quenched Microstructure (AQ)

TEM brightfield images of the AQ treatment reveal a predominantly lath martensite

structure (as also mentioned in the previous Chapter), and selected area diffraction and MDP

confirm that it is indeed martensite (Figure 5.1). Arrays of EDS measurements show a narrow Ni

distribution of 9.3±1.2 wt.% Ni, which is close to the bulk composition of the alloy as did APT

results (10.6±1.9 wt.%Ni). It is possible that there are nano-sized films of retained austenite at

the lath boundary [34–36], but they are not readily observed in BF mode.

Page 81: Microstructural Evolution during Heat Treatment and High ...

65

Figure 5.1: TEM brightfield image with SADP of the AQ treatment. Note the lath martensite structure.

5.3:The L Tempers Microstructure (QL and Q25L)

The L-tempering temperature of 650oC places the 10 Ni alloy in the two-phase region of

the Fe-Ni binary phase diagram implying the decomposition of martensite into moderately Ni-

rich austenite and a Ni-poor ferrite phase (refer to Figure 2.3b). An SEM image of a QL

specimen (Figure 5.2a) shows that 40 minutes at 650°C is long enough to enable widespread

precipitation of austenite, though it is much less than the ~0.5 volume fraction predicted by the

equilibrium Fe-Ni phase diagram. After 25 hours of tempering at 650oC, the austenite

precipitates have grown substantially and the small austenite particles have become micron-long

lamellae that cover just about half of the micrograph, consistent with phase diagram predictions

(Figure 5.2a). The austenite morphology in Figure 5.2a after a 25h exposure to 650oC suggests

austenite precipitation along interlath interfaces, and possibly along packet and block boundaries.

Page 82: Microstructural Evolution during Heat Treatment and High ...

66

Figure 5.2: a) SEM images of QL and Q25L. After only 40 minutes of tempering, austenite has precipitated out.; after 25 hours,

the precipitates have evolved and occupy more than 50% of the total area. b) TEM-EDS arrays start around bulk composition (10

wt.%) for short tempering, but a bi-modal distribution is visible after 25 hours.

Arrays of EDS measurements were obtained in the TEM for the QL and Q25L specimens

and the results are compared in Figure 5.2b. In the QL condition, a Ni content distribution in the

microstructure with a peak centered at 9.3±0.9 wt.% Ni is noted although Ni levels as high as

17% and as low as 6.5% were observed. The peak is believed to be ferrite whose composition for

the most part is close to the Ni content of the alloy rather than the ferrite composition of ~5% Ni

predicted by the phase diagram. Indeed, EDS coupled with microdiffraction measurements show

that the ferrite for the most part is still supersaturated in Ni and has a composition ranging from

Page 83: Microstructural Evolution during Heat Treatment and High ...

67

7.5-10 wt.% Ni. Locations exhibiting high Ni content are associated with precipitated austenite

and discussed further below. The observed response clearly reflects the sluggish diffusion of

nickel at this temperature. In contrast, in the Q25L specimen, TEM/EDS measurements show

two distinct peaks in Ni content distribution in the microstructure: a ferrite peak at 5.1±0.6 wt.%

Ni and a wider austenite peak centered at 13.3±1.5 wt.% Ni, reflecting the evolution of the

tempering process and the gradual partitioning of Ni between ferrite and austenite.

Returning to the QL specimen, a TEM BF image (Figure 5.3a) shows that the precipitates

are small, around 100-200nm in size, and form at the boundaries between martensite laths. MDP

shows that the precipitate is thermally stable. EDS analysis of the austenite particle shown in

Figure 5.3a is presented in Figure 5.3b confirming its composition. The austenite precipitates

show a moderately high-Ni core between 15-18 wt.%, consistent with the austenite composition

expected from the phase diagram at this temperature (Figure 2.3b). The Ni content in the

austenite however appears to decrease gradually across the austenite/ferrite interface rather than

abruptly (Figure 5.3b), but this is attributed to an overlap through the foil thickness of the two

co-existing phases, one Ni-rich and one Ni-poor; the highest values in the center of the

precipitate are believed to be accurate.

Page 84: Microstructural Evolution during Heat Treatment and High ...

68

Figure 5.3: a) TEM BF/MDP with corresponding EDS measurements (b) of the QL treatment. Note that the center of the

precipitate is Ni-rich, around 17 wt.% Ni, and radially decreases in composition. This is an artifact of overlap between the

austenite and ferrite, which causes a dilution effect where the austenite is thinnest. This is confirmed from interfacial composition

measurements using APT (c), where the composition measured near the phase boundary matches the composition measured by

EDS at the center of the austenite particles.

Page 85: Microstructural Evolution during Heat Treatment and High ...

69

APT measurements (Figure 5.3c) confirm that the variation seen in TEM is indeed

artifact from overlapping phases: within nanometers of the boundary, the austenite composition

is measured to be 17.5±3.0 wt.% Ni, consistent with the values measured in the center of the

austenite precipitates using EDS. Adjacent to the austenite/ferrite interface, the Ni content in the

ferrite is somewhat higher (~6.5% Ni) than that predicted by the phase diagram (~5% Ni)

although it is higher in the bulk ferrite as previously noted.

In summary, APT and EDS together determine that QL produces small 100-200 nm

austenite precipitates with Ni content of about 15-17 wt.%, though the ferrite is still

supersaturated in Ni; this is consistent with the notion that the volume fraction of austenite is less

than that predicted by the phase diagram and the tempering kinetics are limited by Ni diffusion.

MDP show that these precipitates are thermally stable austenite.

After 25 hours of tempering at 650oC, the austenite precipitates have grown substantially,

predominantly along the martensite lath interfaces (Figure 5.2a). Ferrite and austenite were

distinguished using MDP and their Ni content was measured by EDS (Figure 5.4a and b).

Results show that the ferrite is uniformly depleted in Ni to the predicted equilibrium value of 5

wt.% Ni, and this is matched well by APT measurements (4.8±1.0 wt.% Ni). However, the Ni

content as measured by EDS within the austenite varied quite substantially, between 12 and 17

wt.% (Figure 5.4b). Unlike the EDS measurements of QL precipitates, the highest values

measured do not appear at the center of the precipitates, but at rather the ends and edges of the

austenite lamellae, and thus, this variation cannot entirely be attributed to overlapping phases.

APT (Figure 5.4c) measurements on the other hand fall in the lower end of the range obtained by

EDS (12.0±2.2 wt.% Ni) and showed no measurable composition gradient. It is worth noting that

EDS measurements were carried out along the length of the lamellar austenite precipitates that

Page 86: Microstructural Evolution during Heat Treatment and High ...

70

decorate the prior martensite lath interfaces and over a distance of ~1 m, whereas the FIB lift-

out APT specimen likely traversed the austenite lamella and the measurement is typically only

over 100-200 nm. The origin of Ni content fluctuation in austenite in the EDS measurements is

not clear. As we know that in even in the early stages such as the QL treatment (40 minutes at

650oC), it is possible to get fine austenite particles with 17% Ni, and thus Ni pile-up at the

advancing austenite/ ferrite interface (due to reduced diffusivity of Ni in austenite as compared to

ferrite/martensite) appears to be an inadequate explanation. Combined, EDS indicates that there

are pockets of higher Ni similar to QL composition (17 wt.%), while APT and EDS show that

the composition is in fact not uniform and drops as low as 12 wt.% in many regions. Importantly,

MDP along these lamellae confirm that they are thermally unstable and have transformed to

martensite during the quench following this tempering step.

Page 87: Microstructural Evolution during Heat Treatment and High ...

71

Figure 5.4: a) TEM BF/MDP with corresponding EDS (b) of Q25L shows Ni content varies throughout, with some pockets of

~17 wt.% Ni and large swaths of 12-15 wt.% Ni. Some of the moderate Ni values are believed to be true, as they are from the

thickest parts of the austenite, and thus are unlikely to be an average of a higher Ni content averaged with low-Ni ferrite. b) APT

confirms that some of the low-Ni readings are not an artifact from overlapping phases, as the boundaries can be as low as 12

wt.% Ni. Together, we can conclude that some regions within Q25L austenite are quite Ni-rich, while others are very Ni-lean.

MDP shows the austenite of Q25L is thermally unstable.

Page 88: Microstructural Evolution during Heat Treatment and High ...

72

5.3: The T’ Temper Microstructures (Q5T’, Q25T’, Q125T’ and Q336T’)

The T’ tempering temperature is 110°C lower than L (650oC), and thus results in

significantly slower Ni diffusion. After 5 hours of tempering, the precipitates seen by SEM are

smaller and sparser than just 40 minutes of L (compare Figure 5.5a with Figure 5.2a); only after

125-336 hours of tempering do we begin to see the austenite take a coarser, lamellae-like

morphology. From the Fe-Ni phase diagram (Figure 2.3b), we expect the T’ to have ferrite

composition of ~6 wt.% Ni and a significantly richer Ni content of ~25 wt.% in austenite. TEM

EDS arrays show the slow depletion of Ni in ferrite, starting from 9.5±1.3 wt.% Ni after 5 hours

(Q5T’), going to 8.7±1.1 wt.% Ni in Q25T’, and then to 6.9±0.7 wt.% in Q125T’, and finally to

6.3±0.6 wt.% Ni in Q336T’ (Figure 5.5b); this progression is illustrative of the sluggish Ni

diffusion kinetics and the final value is in excellent agreement with the Fe-Ni phase diagram

(Figure 5.5b).

Page 89: Microstructural Evolution during Heat Treatment and High ...

73

Figure 5.5: a) SEM comparison of T’ treatments. Ni diffusion at the T’ temperature is much slower than at L, and so even after 5

hours there is not a significant volume of precipitates. After 125 hours, precipitates now line most lath/packet/block/grain

boundaries, and after 336 hours, coarsened globules are seen along high-angle boundaries. b) EDS measurements of the Ni

content in ferrite show a progressive shifts towards its equilibrium value with longer tempering times

Page 90: Microstructural Evolution during Heat Treatment and High ...

74

The austenite precipitates in Q5T’ were examined in TEM BF mode and measured

between 50-100nm in diameter; after 25 hours, they had grown to around 200nm in diameter;

after 125 hours, they begin to show directional growth along interfaces and approached lengths

of 0.5µm. Q336T’ included two austenite morphologies: lamellae of the order of 1µm along lath

boundaries and large, equiaxed precipitates with radius between 0.5-2µm at high-angle

boundaries.

Austenite composition measurements using EDS on the Q5T’ specimen were not reliable

due to the fineness of the precipitate size, and hence the ratio of precipitate size to TEM foil

thickness. While some measurements were around 22 wt.% Ni, the Ni content often varied from

10-15 wt.%, likely due to overlap of ferrite and austenite in the foil thickness. As the precipitates

grew in size through further tempering (Figure 5.6a and d), EDS data became more reliable, and

a consistent austenite composition of 24-26 wt.% Ni emerged as the ceiling for Ni content in the

core (Figure 5.6b and e). As in QL, these precipitates showed a Ni gradient (radially for equiaxed

precipitates, lengthwise for high aspect-ratio precipitates), due to an overlap with the lesser Ni-

containing ferrite and hence an averaging effect. In contrast, APT confirmed that these austenite

precipitates were in fact uniformly high in Ni, with austenite compositions near the interface of

24.1±3.9 wt.% in Q25T’ and 24.8±4.0 wt.% Ni in Q125T’, both closely matching the EDS

values obtained from the core of the austenite precipitates (Figure 5.6c and f). We therefore

conclude that Q25T’ and Q125T’ precipitates have uniform Ni content between 24-26 wt.% Ni;

it is reasonable to assume that the precipitates in Q5T’ are similarly composed. MDP of these

relatively small and Ni-rich T' precipitates confirmed them to possess an FCC structure (Figure

5.6a and d), implying that the austenite particles were thermally stable and did not transform to

martensite on cooling.

Page 91: Microstructural Evolution during Heat Treatment and High ...

75

Figure 5.6: a) TEM BF with corresponding EDS (b) of Q25T’ shows the core of precipitates have a composition between 24-26

wt.% Ni with radially decreasing values. As in QL, APT (c) confirms that this decrease is an artifact of phases overlapping

through the foil thickness – APT composition measurements near austenite/ferrite interface closely matches those measured with

EDS at the austenite core. As precipitates grow with increasing tempering time (Q125T’, d), the overlap effect in EDS diminishes

while the core value remains constant (e). APT again corroborates EDS measurements(f). All precipitates in Q5T’-125T’ were

found to be thermally stable (FCC).

Page 92: Microstructural Evolution during Heat Treatment and High ...

76

Figure 5.7: Q336T’ has large, coarsened, equiaxed austenite precipitates on the scale of 1-2µm. They have a leaner composition

than smaller T’ treatment precipitates (20-22 wt.% Ni vs 24-26 wt.%), and are sometimes thermally stable (a, c) or have

transformed to martensite during the quench (b, d). Note the thermally stable austenite (c) has a low defect structure, while the

thermally unstable austenite (d) has a lath-like internal structure.

Page 93: Microstructural Evolution during Heat Treatment and High ...

77

In the Q336T’ condition, small, thermally stable austenite particles with 24-26 wt.% Ni

are also observed. However, the lengthy tempering time also allows some precipitates to

significantly grow and coarsen; equiaxed austenite precipitates with diameter up to 2µm can be

observed (Figure 5.7). In these relatively coarse precipitates, overlap effect if present is

negligible even within tens of nanometers of the interface, giving reliable, uniformly high EDS

measurements throughout of 20-22 wt.% Ni. Some of these globular austenite precipitates are

thermally stable, and show a very low-defect structure within them; however, other austenite

precipitates with a similar Ni content are unstable (MDP confirms bcc structure), with internal

lath-like sub-structures and a high dislocation density.

Table 5.1: Composition of austenite in the four isothermal heat treatments as measured by APT (at.%)

AQ QL Q25L Q25T' Q125T' Avg. Err.

Fe 87.1% 79.5% 86.3% 72.6% 73.7% 3.7%

Ni 10.1% 16.7% 11.5% 23.7% 23.0% 2.8%

Mn 0.67% 1.77% 0.92% 2.46% 2.06% 0.30%

Cr 0.66% 0.79% 0.63% 0.60% 0.66% 0.14%

Mo 0.52% 0.26% 0.33% 0.30% 0.28% 0.07%

Si 0.48% 0.29% 0.15% 0.19% 0.12% 0.06%

C 0.26% 0.29% 0.05% 0.01% 0.00% 0.04%

Cu 0.17% 0.29% 0.07% 0.18% 0.04% 0.03%

V 0.07% 0.03% 0.05% 0.02% 0.04% 0.01%

Al 0.02% 0.01% 0.03% 0.02% 0.06% 0.01%

In addition to Ni, it is relevant to consider the possible role of other minor alloying

elements present in the alloy in affecting the thermal stability of austenite. The composition of

austenite measured in several heat treatments using APT is summarized in Table 5.1. However,

while many of these elements also partition during tempering (e.g. Mn), the difference in their

content in austenite between the L and T’ tempering treatments is not as significant as the

Page 94: Microstructural Evolution during Heat Treatment and High ...

78

partitioning of Ni. As discussed in the previous chapter, C is notably present in QL austenite;

however, it is still relatively low (0.3 at.%, or 0.05 wt.%) and is therefore not as significant in

relation to Ni partitioning; no other treatment has measurable C due to carbide precipitation and

growth in the early stages of tempering.

5.4: Discussion of Isothermal Tempering

Table 5.2: Size and composition of precipitates in isothermal treatments

Treatment Time (h) Morphology Size (µm) Austenite Comp

(wt.% Ni) Stability

L (650°C) 0.67 Equiaxed 0.1 17 Stable

25 Lamellar ~1 12-17 Unstable

T' (540°C)

5 Equiaxed 0.05 10-22* Stable

25 Equiaxed 0.25 25 Stable

125 Lamellar 0.5 25 Stable

336 Lamellar ~1 25 Stable

Equiaxed 0.5-2 22 Mixed

* Q5T’ austenite composition measurements are believed to be unreliable due to their size, true composition is assumed to be

25 wt.% Ni.

A summary of the austenite precipitate shape, size, composition in terms of Ni content,

and stability following quenching from the different isothermal tempering conditions examined

in this chapter are presented in Table 5.2. For both tempering temperatures (L and T’), short

tempers produce small, equiaxed, thermally stable austenite with Ni content that closely matches

the Fe-Ni phase diagram. However, as the austenite grows and coarsens during extended

tempering, both isothermal treatments produce thermally unstable austenite. There are two

possible explanations for this onset of instability: their relative decrease in Ni content and

increase in size.

Larger austenite precipitates in both treatments appear slightly leaner in Ni than the

precipitates seen in shorter treatments, which should lessen thermal stability. The austenite in

Page 95: Microstructural Evolution during Heat Treatment and High ...

79

Q25L is on average roughly 3 wt.% Ni poorer than QL, although it is highly inhomogeneous; the

homogenous, globular Q336T’ precipitates are also about 3 wt.% poorer than lamellar or smaller

equiaxed T’ austenite. While this slight decrease in Ni undoubtedly decreases the thermal

stability of the austenite precipitates relative to the richer, short-temper ones, the stability of

small, moderate-Ni QL austenite juxtaposed with the instability of globular, Ni-rich Q336T’

austenite suggests that size is likely the dominant factor in thermal stability.

Takaki et al. showed that there is a strong Ms temperature suppression for sub-micron-

sized austenite, as an increasingly significant chemical driving force is required to overcome the

relatively massive elastic energy barrier for transformation [38]. While Q336T’ is significantly

more Ni rich than Q25L, both have micron-scale dimensions and are thermally unstable; despite

being much more Ni-lean than unstable Q336T’ austenite, the 100nm QL austenite precipitates

are thermally stable. Therefore, we can conclude then that within the range of L and T’

tempering temperatures, that is to say, between 15-25 wt.% Ni, austenite precipitates are only

thermally stable if sufficiently small. Moreover, if allowed to coarsen to micron-scale sizes

through extended tempering, austenite in all isothermal treatments within this temperature range

should transform to martensite upon a room temperature quench. This size effect is schematically

superimposed on the Ms temperature versus Ni plot in Figure 5.8, while Figure 5.9 summarizes

the above discussion relating to size and Ni content on the stability of the austenite precipitates.

Page 96: Microstructural Evolution during Heat Treatment and High ...

80

Figure 5.8: The Ms temperature diagram for a binary Fe-Ni alloy [43] predicts both L and T’ austenite to be thermally unstable

at room temperature. The observed austenite stability in short tempering times is thought to be due to a size-effect.

Figure 5.9: For both L and T’ tempers, the small austenite from short tempers is thermally stable. However, even the very Ni-

rich T’ austenite becomes unstable after 336 hours, suggesting both small L and T’ tempers are only stable due to their small size.

If let coarsen long enough, presumably all T’ austenite should be unstable.

Page 97: Microstructural Evolution during Heat Treatment and High ...

81

However, effect of composition on thermal stability may still be relevant in certain

circumstances. That is, a sufficiently large, moderate-Ni L austenite may be thermally unstable

while a similarly sized, Ni-rich T’ austenite would not be, as the chemical driving force for the

transformation would be higher for L than T’. In principle then, lower temperature tempering in

this alloy system should be able to produce thermally stable austenite of a larger size than high

temperature treatments due to a composition-enhanced stability.

There are other considerations that go into selecting a tempering schedule, even within a

range capable of producing thermally stable austenite of a desired size. The kinetics of austenite

nucleation and growth are substantially different between high and low temperature tempers as

evidenced by comparing QL with Q25T’. The Q25T’ needed to be tempered roughly 35 times

longer than QL to produce austenite precipitates of similar sizes. In addition, there appear to be

morphological differences between L-austenite and T’-austenite, the former appearing more

lamellar, and forming at inter-lath interfaces whereas T’ austenite appears globular. As shown by

Kim et al. [13], a lamellar austenite morphology is more effective at breaking up trans-packet

cleavage. But beyond these considerations, desired mechanical instability is another factor that

can affect selection of tempering temperature. TRIP mechanisms are also a function of

composition, size, and morphology, and so choosing a composition that produces austenite that

can be readily rendered mechanically unstable may be desirable. So while precipitated austenite

from a wide range of short tempering temperatures would be stable due to its small size, there are

additional considerations that influence selecting a tempering schedule.

Page 98: Microstructural Evolution during Heat Treatment and High ...

82

5.5: The QLT Treatment – Results and Discussion

Figure 5.10: SEM of QL (shown here again for convenience) and QLT. Note that there is significantly more austenite than QL,

which suggests that the T temper contributes heavily to additional austenite nucleation and growth.

The microstructure of the multistep QLT temper is the result of a summation of the

microstructure resulting from the QL temper and a temper in between T’ and L; by comparing

QLT to AQ, QL, and the QT’ samples, we can isolate the effect of each tempering step on the

final microstructure. SEM of QLT (Figure 5.10) shows noticeably more precipitation than QL,

which suggests that additional austenite continues to form during the T treatment. The Fe-Ni

phase diagram (Figure 2.3b) predicts this second generation austenite to have a higher Ni content

than austenite formed during L; however, it is unclear if the L-generation of austenite that was

thermally stable will have adequate time to acquire Ni to attain the new equilibrium value at the

T temperature, as Ni diffusion within austenite is slower than Ni diffusion to the austenite/ferrite

interface through ferrite. Brightfield images show a mixture of lamellar and equiaxed

precipitates; lamellar austenite can have a length as much as 1µm, though narrower than Q25L

and Q336T’ lamella (100nm), while equiaxed precipitates are roughly 100nm in diameter

(Figure 5.11a). TEM EDS (Figure 5.11b) shows a scattered range of Ni level between 15-20

wt.%, even at the core of the austenite. Like in QL and T’ treatments, this may be simply due to

Page 99: Microstructural Evolution during Heat Treatment and High ...

83

the Ni-rich austenite and Ni-poor ferrite phases overlapping through the foil thickness, but it also

may be attributed in part to two distinct generations of austenite with different Ni content. APT

confirms that some of this variation is in fact real, as austenite as low as 14.8±2.6 wt.% Ni and as

high as 22.1±3.6 wt.% Ni is present (Figure 5.11c). MDP of QLT precipitates confirm they are

thermally stable.

Figure 5.11: a) TEM BF of QLT with corresponding EDS (b) shows mixed composition of 15-20 wt.% at the center of

precipitates. This is not an artifact of phase overlap, but rather is a result of two distinct generations of austenite with different

growth compositions. c) APT tips confirm that both L-generation and T-generation austenite are present in QLT.

Page 100: Microstructural Evolution during Heat Treatment and High ...

84

SEM images corroborate the austenite volume increases seen by Zhang [25] and by Jain

et al. [28]. The T treatment measurably increases the volume fraction of austenite precipitates.

However, the reason for this increase appears to be somewhat different from the theory they

proposed; they argue that many of the precipitates seen in the SEM image following the QL

temper are in fact fresh martensite, and thus the thermally stable austenite content is less than

what it appears to be visually in such images. They postulate that much of the stable austenite

found in QLT is thus a consequence of precipitation of a new generation of austenite within these

Ni-rich fresh martensite pockets. However, as discussed previously, we did not find evidence for

unstable austenite in QL and we attributed the stability to size. Therefore, our view is that the

increase of stable austenite content from the second lower temperature T temper is due to

additional precipitation on top of existing L-generation austenite as well as nucleation and

growth. This two-step austenite evolution is evidenced by a combination of EDS and APT,

which show two compositionally distinct generations of thermally stable austenite; in the case of

co-precipitation, because the T temper is so short, the residual L-generation austenite does not

have time to equilibrate to the new austenite composition. Thus, we can separate the L- and T-

generation of austenite precipitation through Ni content; 15-17 wt.% L-generation austenite and

20-22 wt.% T-generation austenite. Schematically, this two-stage austenite core-and-shell

morphology evolution is shown in Figure 5.12. In addition, we can now provide a more complete

schematic overview of the QLT process compared to that presented in Chapter 2 (Figure 2.4),

complete with austenite composition and identification of phases for each stage of tempering

(Figure 5.13).

Page 101: Microstructural Evolution during Heat Treatment and High ...

85

Figure 5.12: QLT process produces thermally stable austenite around 15-17 wt.% Ni during the L treatment. During T, growth

resumes, resulting in additional growth in the range of 20-22 wt.% Ni.

Figure 5.13: QLT Process, understood in terms of austenite size and composition.

Page 102: Microstructural Evolution during Heat Treatment and High ...

86

The thermal stability of both generations of QLT austenite appears to be due to their

small size: QLT austenite is still in the sub-micron regime, and the composition range falls

within the boundaries of the isothermal L and T’ treatments- both of which were found to be

unstable when sufficiently large. While the Ni-enriching T-step does provide additional thermal

stability to some of the austenite, it is perhaps not the primary factor for an increase in thermal

stability of the austenite as a whole. Perhaps a longer L-step could also produce the same volume

fraction of thermally stable austenite with similar morphology.

So why does QLT perform better in mechanical tests if similar volume fractions can be

produced isothermally? Zhang [25] is not the first to describe the mechanical superiority of a

QLT treatment in Ni steels; Kim et al. [13] in 1983 found that QLT treatments on a 5.5Ni steel

also produced superior Charpy toughness versus a range of isothermal treatments. It seems that

there must be microstructural reasons that a two-stage temper has superior mechanical properties

during dynamic deformation, be it Charpy impact energy or ballistic resistance. In both studies,

QLT-treated Ni-containing alloy is believed to undergo a TRIP mechanism wherein the austenite

dynamically transforms to martensite, and like thermal stability, the mechanical stability of

austenite is also a function of austenite composition and size. Recently, Yuan et al. [41]

suggested that a distribution of austenite in the microstructure with varied composition can have

a ‘spectral TRIP effect’, where the critical strains to transform individual austenite particles to

martensite are then spread over a large range, thereby increasing the work hardening capacity

and toughness of the material. In contrast, homogeneously composed austenite from that same

alloy was found to transform within a narrow strain window and was thus prone to failure with

lower toughness. A similar effect was also noted by Wang et al. [39,40] with respect to varied

austenite size- a distribution of austenite grain sizes with the same composition also helped

Page 103: Microstructural Evolution during Heat Treatment and High ...

87

postpone failure and improved overall toughness through staggered critical strains. The QLT

treatment that the 10Ni steel has been subjected to may combine both of these mechanisms

through its two stage temper: L- and T-generation austenite are chemically distinct, providing a

composition-based spectral TRIP; while the mixed sizes and morphologies arising from two

stages of growth provide a size-based spectral TRIP. Thus, the ballistic superiority of 10Ni QLT

can be explained through a combination of spectral TRIP effects, which are enabled through its

two stages of thermally stable austenite growth that can at least in part be rendered mechanically

unstable.

Page 104: Microstructural Evolution during Heat Treatment and High ...

88

Chapter 6: Microstructural Evolution in an Fe-10Ni-0.1C Steel

During Dynamic Deformation

6.1: Introduction

It is now established that Fe-10Ni-0.1C steel subjected to the QLT heat treatment results

in a microstructure containing two compositionally distinct, thermally stable generations of

austenite precipitates: the 15-17 wt.% Ni L-generation and the 20-22 wt.% T-generation.

However, the question remains- how does the microstructure evolve during high strain-rate

deformation? Specifically, what happens to each generation of austenite as deformation

commences, progresses, and then localizes, initiating adiabatic shear bands (ASB) that then

propagate through the material? This chapter examines these aspects using the compression

Kolsky-bar set-up to obtain the high strain rates needed to study the problem.

6.2: Kolsky-Bar Calibration using 4140 Steel

The supply of Fe-10Ni-0.1C steel with the QLT heat treatment steel was limited, and

therefore, Kolsky-bar experiments were first performed on 4140 tempered martensite for

purposes of calibration. A wide range of firing pressures was examined for each projectile bar

length, from low pressure to a regime where specimen fractured consistently or the maximum

permitted chamber pressure was reached, whichever occurred first. Strain/strain-rate pairing was

recorded in each instance and plotted for different projectile bar lengths (Figure 6.1) so that the

test parameters for specific predetermined strain/strain-rate regime could be selected a priori for

the 10Ni QLT experiments. While the 15.3cm bar achieved the greatest strain-rates, the total

height strain was limited by the upper allowable firing pressure; alternatively, the 30.5cm bar

could achieve much higher height strains while still reaching relatively high (2000s-1

) strain-

Page 105: Microstructural Evolution during Heat Treatment and High ...

89

rates, and so it was selected for the QLT experiments. Etched 4140 smooth cylindrical samples

were observed using SEM, and the presence of ASBs were confirmed; thus, this sample

geometry was determined sufficient for producing ASB in 10Ni QLT experiments.

Figure 6.1: Strain/strain-rate pairs for Kolsky experiments on 4140 tempered martensite. Each datum point correlates to a

specific firing pressure with the noted bar.

6.3: Kolsky-Bar Testing of 10Ni-QLT

As the focus of this study was the microstructural evolution during shear localization, low

strain and strain-rate Kolsky-bar experiments were not performed on QLT because they were

considered unlikely to localize and produce ASBs. Instead, gun pressure was increased up to and

just short of fracture in order to attempt to consistently induce shear localization. A variety of

samples in various stages of shear localization were produced, and the extent of shear

localization in each was determined using optical microscopy (examples are shown in Figure

6.2). The appearance of the white bands classified them as ‘transformed bands,’ suggesting they

had undergone dynamic recrystallization in the regions of most intense localization

[51,52,57,61,68]. No cracks or voids were observed in the samples containing these white bands.

0

500

1000

1500

2000

2500

3000

0 0.05 0.1 0.15 0.2 0.25

Stra

in-r

ate

(s-1

)

Height Strain

15.3cm

20.3cm

30.5cm

40.6cm

Page 106: Microstructural Evolution during Heat Treatment and High ...

90

Figure 6.2: A diffuse and a well-developed adiabatic shear band

One particular sample was chosen for further investigation because it had two well-

developed shear bands in opposite corners which dissipated towards the center of the sample

(Figure 6.3). Thus from this single sample, several TEM specimens were FIB-lifted-out from

specific locations that corresponded to various stages of shear localization; a location far from

the shear band was included as a baseline and represented what might be considered as a

homogeneous plastic deformation region prior to the onset of localization. The nominal global

strain-rate for this particular sample was around 1900s-1

and the total height strain about 20% (It

should be noted however that the strain rates and strains experienced after localization

commences, and particularly within the ASB, will be significantly higher).

Page 107: Microstructural Evolution during Heat Treatment and High ...

91

Figure 6.3: Sample chosen for microstructural analysis. A well-developed shear band initiates at the bottom left, propagates

towards the center of the sample (top right) and eventually dissipates. Another band initiated in the opposite corner (not visible in

this image) that also propagated towards the center of the sample, though it was smaller and less developed than the one seen in

the montage above. The total height strain was about 20% and the strain-rate was about 1900s-1. Specific TEM specimen lift-out

locations are marked using numbers from 1 to 10. The relative displacements of the vertical etching bands (banding in the rolled

plate) provide a sense of the large shear experienced within the ASB.

6.4: Nanoindentation across the ASB

An array of nanoindentation hardness measurements was made across the well-developed

ASB (Figure 6.4). There is a measurable increase in hardness across the band- from 5.5GPa to

7GPa, in agreement with the increase from 6.5GPa to 8GPa previously measured by Wang and

Kumar across a shear band in this alloy that had been subjected to a different heat treatment [22].

In addition, the region with increased hardness extends well beyond the boundaries of the white

portion of the band as was also observed by Xue and Gray in 316L stainless steel [58,69]. They

Page 108: Microstructural Evolution during Heat Treatment and High ...

92

noted that the white band did not signify the boundary of the shear localization but that the

microstructure in the white band has dynamically recrystallized, whereas the regions flanking it

was in the process of forming subgrains.

Figure 6.4: Nanoindentation across the shear band (marked in purple). There is an increase in hardness (~5GPa to ~7.5GPa) that

extends beyond the boundary of the unetched portion of the shear band.

Page 109: Microstructural Evolution during Heat Treatment and High ...

93

6.5: Microstructural Analysis of the dynamically deformed Specimen

As noted above, the strain localized region extends in width beyond the unetched portion

of the band, the extent of which can be seen in SEM micrographs (Figure 6.5). The center of the

band itself appears featureless and corresponds to the ‘white band’ observed by optical

microscopy; adjacent to this region, the extent of shearing is evident through the highly

elongated microstructure parallel to the ASB. About 10-15µm away from the center of the band

and on either side of it the microstructure is significantly less aligned. The unetched portion of

the band is approximately 4µm wide. Inside this central portion of the band, BF TEM

micrographs show that significant microstructural changes have occurred: no trace of the former

lath martensite structure remains, and is instead replaced by a fine-grained, equiaxed structure

(Figure 6.6).

Page 110: Microstructural Evolution during Heat Treatment and High ...

94

Figure 6.5: SEM micrographs of QLT at increasing magnification (top to bottom), undeformed (left) and near the shear band

(right). Note the lack of features in the band, which gives it the unetched appearance. Outside of this region, the austenite

precipitates can be seen to be sheared parallel to the shear plane.

Page 111: Microstructural Evolution during Heat Treatment and High ...

95

Figure 6.6: SADP in undeformed QLT compared to in the band using the same aperture. On the left, the lath martensite structure

produces a highly textured SADP; on the right, the highly misoriented, equiaxed, mechanically recrystallized grains in the band

produce a ring-like SADP.

To characterize the microstructural evolution during adiabatic shear banding, we consider

four locations in particular: i) far from the band in a homogeneously deformed region; ii) ahead

of the white band, in the band plane, where localization had occurred but was not well-developed

(labelled ‘7’ in Figure 6.3); iii) adjacent to the developed band, and an orientation that was

normal to the band plane (labelled ‘5’ in Figure 6.3); and iv) in the band, normal to the band

plane (labelled ‘10’ in Figure 6.3). As this deformation mode is extremely rapid, any associated

phase transformation is expected to be diffusionless, and thus the Ni content can be used to link

deformed microstructure to undeformed microstructure. Therefore, we particularly look to

characterize regions with Ni content that ties them to the two generations of austenite in

undeformed material, the 15-17 wt.% L austenite and the 20-22 wt.% T austenite.

Page 112: Microstructural Evolution during Heat Treatment and High ...

96

6.5.1: Location far from the shear band:

Figure 6.7: TEM lift-out from a location that is far from the shear band (circled in blue dashed line).

The ‘striations’ seen in Figure 6.7 that etch light and dark result from rolling and

originate likely from macrosegregation but conveniently serve as markers for demarcating

localized deformation. Far from the ASB, these striations remain relatively parallel, suggesting

there is not significant local shearing in this region as opposed to the region flanking the ASB

where significant shearing is evident (Figure 6.7). Indeed, the microstructure seen in TEM BF in

a specimen obtained from the location circled in Figure 6.7 appears very similar to the

undeformed material, and the lath microstructure is still intact (Figure 6.8). Notably, precipitates

with Ni content ranging from 15-20 wt.% (indicative of both, L-generation and T-generation

austenite) were found to retain the FCC structure implying that the L- and T-generation austenite

are mechanically stable in this location (Figure 6.8).

Page 113: Microstructural Evolution during Heat Treatment and High ...

97

Figure 6.8: Far from the shear localized region, the lath martensite seen in the undeformed microstructure is still intact. Here, the

composition of two austenite precipitates are measured and found to be between 15-20 wt.% Ni. Thus these precipitates are a

mixture of L- and T-generation austenite, and MDP shows that they are mechanically stable.

6.5.2:Location ahead of shear band:

The differential etching response manifesting as bright and dark striations in the optical

micrograph shows that there is intense shear just ahead of the ‘white band’ (Figure 6.9a); in the

TEM, BF imaging confirms that the martensitic laths and/or austenite precipitate morphology

noted in the undeformed material is no longer visible and the microstructure looks more

‘recrystallized’ (Figure 6.9b). SADP over large areas shows a more or less continuous ring

Page 114: Microstructural Evolution during Heat Treatment and High ...

98

pattern implying the structure consists of multiple grains/subgrains that are significantly

misoriented with respect to each other, suggesting a good level of subgrain rotation has occurred,

likely due to the intense shearing that prevails in this location. However, since this process can

be considered diffusionless, the relationship between these newly-developed substructures and

their parent should still be discernable through Ni content measurements.

Figure 6.9: a) TEM specimen lift-out from a location ahead of the shear band. Notice the stark white band on the left, which

dissipates to the right. b) SADP from the location ahead of the band microstructure (aperture marked in orange). Ring pattern

suggests significant grain refinement and grain rotation and recognized from the relatively low magnification bright field image.

The fine austenite precipitates in the undeformed QLT condition is expected to have a

low-defect structure, but these subgrains show dislocation contrast within them implying

significant plastic deformation had occurred and enough to break these precipitate down into

subgrains. Figure 6.10 shows one such related event: beyond what is in contrast initially in

Figure 6.10b, there is a high-Ni region that extends into a wider area that is not in contrast- this

continuous, unbroken high-Ni region represents the trace of a single former austenite precipitate.

Through tilting (Figure 6.10a), its neighboring subgrain (Figure 6.10c) can be brought into

contrast and similarly indexed. The first subgrain is entirely from T-generation precipitation, as

demonstrated by its homogeneous Ni content of ~22 wt.%; MDP shows it is mechanically stable

austenite. Similarly, its neighboring subgrain is also from T-generation growth (20-22 wt.%) and

Page 115: Microstructural Evolution during Heat Treatment and High ...

99

is mechanically stable. Together, they make up a T-generation austenite precipitate that has

broken up through mechanical recrystallization into mechanically stable austenite subgrains.

This implies that the T generation austenite is stable despite the intense shear experienced in this

region locally and that if it is going to mechanically transform to martensite, it is going to require

even more intense deformation (i.e. much ‘later’ in the dynamic deformation process). The

microstructure also shows that the austenite precipitates from the QLT treatment are further

segmented during the local intense shearing process that occurs ahead of the propagating ASB,

so that a reduction in austenite subgrain size prior to TRIP can further discourage the TRIP

process and may even preclude it.

Figure 6.11 shows another austenite precipitate dynamically broken down into subgrains.

Similar to the previous example, the T-generation subgrain (Figure 6.11c) is mechanically stable.

However, its neighbor is comprised of two generations of austenite growth- the 15-17 wt.% L-

generation and 20-22 wt.% T-generation (Figure 6.11b). This mixture suggests that the parent

precipitate originally had a core/shell structure resulting from T-generation precipitation on top

of an existing L-generation precipitate. The L-generation austenite should be more mechanically

unstable due to its lower Ni content. However, this subgrain is mechanically stable, possibly

stabilized by the fragmentation of the original grain that reinforces the size effect.

Page 116: Microstructural Evolution during Heat Treatment and High ...

100

Figure 6.10: a) Two adjacent subgrains, which are brought into contrast by tilting. b) EDS on the first subgrain gives a

composition of about 22 wt.% Ni (T-generation), and MDP shows it is mechanically stable FCC. c) Its neighbor is also from T-

generation austenite (20 wt.% Ni), and is also mechanically stable austenite. Together, they show a T-generation austenite

precipitate has mechanically broken down into subgrains and these are mechanically stable.

Page 117: Microstructural Evolution during Heat Treatment and High ...

101

Figure 6.11: An L- and T-generation austenite precipitate has mechanically broken down into two subgrains (b and c) and can be

brought into contrast through tilting (a). The subgrain before tilting (b) is mechanically stable and has Ni content varying between

L- and T-generation, and was likely part of a core/shell structure before deformation. Its neighboring subgrain (c) is also

mechanically stable, but its composition shows that it is entirely growth from the T temper.

Page 118: Microstructural Evolution during Heat Treatment and High ...

102

6.5.3: Location adjacent to the shear band:

Next, we examine the deformed microstructure in the regions adjacent to the shear band

(Figure 6.12a). The region normal to the shear plane about 20µm from the edge of the white

band has a microstructure that appears lamellar with lamellae aligned parallel to the plane of the

shear band (Figure 6.12b). This morphology is seen throughout the TEM lift-out specimen,

which was dimensions that are approximately 20µm x 30µm and is larger than the original

martensite packet and block size. Thus, this lamellar microstructure is not the original

martensitic lath structure, but rather a consequence of microstructural evolution during shear

localization- the martensite structure has broken down and has been sheared parallel to the ASB

propagation direction. EDS analysis confirms thin strands of Ni-rich regions aligned parallel to

these lamellae, and an example is illustrated in Figure 6.13, where the strand is composed of

subgrains, some of which are in contrast while others are not. The subgrain that is in contrast has

Ni content ranging from 15-22 wt.%, suggesting it is a mixture of both L- and T-generation

austenite. MDP confirms that the subgrain is mechanically stable austenite (Figure 6.13).

Figure 6.12: a) Lift-out location with ASB demarcated by the two orange dashed lines and axes defined (shear plane normal is X

axis, and shear band is propagating in the Z direction). b) TEM BF image of area about 20µm from the band with rotated axes

marked. Microstructure is elongated in the shear plane (Y-Z).

Page 119: Microstructural Evolution during Heat Treatment and High ...

103

Figure 6.13: Both, the ferrite and the austenite precipitates adjacent to the band are aligned parallel to the shear plane (axes

marked) and have broken up into subgrains. The subgrain that is in contrast here has Ni content ranging from 15-22 wt.%,

suggesting it originates from a mixture of L- and T-generation austenite. MDP shows the subgrain to have the FCC structure.

6.5.4: Location inside shear band:

TEM specimens that were initially lifted out from within the band but aligned parallel to

the band plane provided EDS Ni distributions that narrowly peaked around the nominal

composition of the alloy. This is a direct consequence of an averaging effect due to the grain’s

extremely small dimension along the axis normal to the shear plane (schematically illustrated in

Figure 6.14). Therefore, an in-band sample was milled normal to the shear band to increase

Page 120: Microstructural Evolution during Heat Treatment and High ...

104

single-grain-depth through the foil thickness for TEM analysis, as the grain dimension along the

band propagation axis was comparable to the sample thickness (Figure 6.15). To verify the lift-

out sample thickness was in fact single-grain, EDS Ni distribution plots were compared to those

from undeformed material and were found to be closely matched (Figure 6.16- compare green

and red curves as opposed to the blue curve which was obtained from a TEM lift-out oriented

parallel to the plane of the shear band). To ensure the area thinned to electron transparency was

indeed within the band, the location of the white band relative to the sample was carefully

measured before milling to approximate its location in TEM.

Figure 6.14: Exaggerated schematic of TEM specimen lift-out geometries. The austenite precipitates are stretched parallel to the

shear band- when lift-outs are taken along this plane, there is an averaging effect in the EDS measurements. Lift-outs are taken

normal to the band to mitigate this effect.

Figure 6.15: Left: Location of in-band lift-out prior to FIB milling. The unetched portion of the band is approximately 4µm from

the lift-out edge, while the other boundary of the band is about 10µm into the lift-out. Right: The measurements taken on the

SEM are overlaid on a BF image of the lift-out specimen (which extends beyond what is shown, as marked by the black arrows).

Page 121: Microstructural Evolution during Heat Treatment and High ...

105

Figure 6.16: Normalized EDS Ni distribution of undeformed QLT (Green) in comparison to data obtained from a TEM specimen

that was lifted out parallel to the band (Blue) and another that was oriented normal to the shear band (Red). The parallel geometry

has a strong, narrow peak around the nominal Ni content due to an averaging effect rising from the ratio of grain depth to sample

thickness; the normal-to-band sample has a distribution that closely matches the undeformed material, suggesting it largely has

single-grain depth.

Figure 6.17: BF and corresponding DF image of the region of the deformed specimen that is estimated to include the white band

and the region immediately outside it. There is no observable difference in microstructure, but note that the demarcation of the

boundary is a best estimate based on markings in Figure 6.15.

0

0.05

0.1

0.15

0.2

0.25

0 5 10 15 20

Co

un

ts (

No

rmal

ize

d)

wt.% Ni

Undeformed QLT

In-Band Liftout (Parallel)

In-Band Liftout (Normal)

Page 122: Microstructural Evolution during Heat Treatment and High ...

106

As seen in Figure 6.17, the microstructure estimated to be inside the white band has

mechanically recrystallized into a fine-grain, equiaxed microstructure with grains diameters

between 50-200 nm. Darkfield images confirm the grains have adequately rotated to be

differentiated from their neighbors. The grain size is much smaller than the subgrains observed

ahead of the band (Figure 6.10 and Figure 6.11), demonstrating further grain refinement has

occurred within the band as the shear band developed. The microstructure is also distinctly

different from the regions adjacent to the band (further away from the region shown Figure 6.17

and illustrated in the next sub-section), as no elongated lamellar grain morphology is visible.

There is no sharp, delineating microstructural features that separates the portion of the sample in

the unetched (white) portion of the shear band (labelled ‘inside white band’ in Figure 6.17) from

the region immediately outside of it (labelled ‘outside white band’), although the boundary itself

is only located based on best estimate from the images shown in Figure 6.15.

The low-Ni (previously ferrite lath) grains are between 50-300 nm in diameter, and some

have features that appear to be BCC twins (Figure 6.18). The high-Ni, former austenite

precipitates also appear to be extensively deformed and then mechanically recrystallized with a

morphology that resembles 100-200 nm-long strips within the ferrite and composed of small,

misoriented, equiaxed grains of diameter 20-50nm lined single-file along the strip. Figure 6.19

shows one-such high-Ni region located in the area of the sample believed to be just outside the

white band; the high-Ni strip is narrow and elongated and is surrounded by low-Ni ferrite. The

portion of the strip that is in contrast in Figure 6.19 has a Ni content of about 22 wt.% Ni,

suggesting it is likely T-generation austenite. Its MDP gives an FCC pattern with additional

twinning spots – thus we conclude it is heavily deformed, mechanically stable, T-generation

austenite.

Page 123: Microstructural Evolution during Heat Treatment and High ...

107

Figure 6.18: A small grain is in contrast. It is low in Ni (boundaries marked in orange on EDS gradient plot, 5-7 wt.%Ni). Inside,

a band-like substructure suggests twinning.

Page 124: Microstructural Evolution during Heat Treatment and High ...

108

Figure 6.19: The high-Ni region, formerly an austenite precipitate, is spread over a very narrow strip (about 20nm wide by

200nm in length). The portion that is in contrast is marked in the plot in orange and has a Ni content of 20 wt.%, suggesting it is

T-generation austenite; an MDP from this location confirms it to be FCC and additionally reveals twinning- it is mechanically

stable austenite.

In other regions within the shear band, there is evidence for the austenite being

mechanically unstable. Figure 6.20 shows a narrow strand along which the Ni content was high

but is surrounded by low-Ni regions in the portion of the sample along the border of the

approximated white band. The in-contrast grain in the strand is about 20nm in diameter and has

Page 125: Microstructural Evolution during Heat Treatment and High ...

109

composition of around 20 wt.% Ni, suggesting it had evolved from T-generation austenite.

However, it has a BCC MDP implying it had mechanically transformed to martensite. Since the

size of this particle is of the order of 20nm, a plausible scenario is that the parent more-or-less

equiaxed austenite particle that was deformed into the long strand-like configuration transformed

to the BCC martensitic structure (TRIP) and then fragmented by mechanical recrystallization

into multiple fine subgrains, one of which is in contrast in Figure 6.20. Similarly, an L-

generation grain (15-17 wt.% Ni) from the white-band section of the lift-out sample which is in

contrast is shown in Figure 6.21. Given its relatively lower Ni content, it is should be less

mechanically stable than a T-generation particle; indeed, it has also mechanically transformed to

martensite.

Due to the fine, highly-defective microstructure, it is rather difficult to determine

definitively whether the carbide particles are still present within the shear band. However,

occasionally, local measurements which indicate high Mo levels were detected by EDS. If the

carbides are plastically elongated and have indeed wholly or partially dissolved, the Mo is still

locally enriched.

Page 126: Microstructural Evolution during Heat Treatment and High ...

110

Figure 6.20: The region presented in this image falls along the border of the white band. On the left, the narrow strand of high-Ni

(running left to right) is shown to be over 100nm long with only a small grain of about 20nm in diameter in contrast. On the right,

the composition of the grain is isolated and highlighted in orange: it is about 20 wt.% Ni, and therefore came from T-generation

austenite. The MDP is BCC [111]- this grain is very fine, high in Ni but was mechanically unstable.

Page 127: Microstructural Evolution during Heat Treatment and High ...

111

Figure 6.21: Another region close to the border of the white band; here, a small grain that dynamically recrystallized from L-

generation austenite (15-17 wt.%, marked in orange) has mechanically transformed to martensite.

Page 128: Microstructural Evolution during Heat Treatment and High ...

112

6.6: Discussion

By comparing the microstructure at various locations relative to the shear band in

addition to using the baseline microstructure in the undeformed QLT condition, the

microstructural evolution during high strain-rate deformation and adiabatic shear banding can be

reconstructed. Far from the shear localized region, the lath microstructure is intact and there was

no evidence for deformation-induced phase transformation of austenite as precipitates with both

L- and T-generation compositions were found to be mechanically stable. Ahead of the shear

band, austenite precipitates are extensively sheared and begin to divide into subgrains (by

‘mechanical recrystallization’), and despite the large strains needed to form these subgrains, both

L- and T-generation austenite are still observed to be mechanically stable.

Adjacent to the most intense shear localization, SEM images show that the precipitates

are stretched along the axis of the plane’s propagation. TEM of this region from the perspective

normal to the shear plane shows a lath-like subgrain microstructure parallel to the shear plane,

similar to the microstructure observed by Xue and Gray in 316L stainless steel [58,69] and Wang

and Kumar in a steel with the same alloy composition but subjected to a different heat treatment

[22]. Together, these two microstructural perspectives show that the austenite precipitates are

pancaked parallel to the shear plane and are broken into finer subgrains. Furthermore, EDS and

MDP together show that subgrains from both L- and T-generation austenite are stable in this

region.

The unetched portion of the band appears featureless in SEM and is approximately 4µm

thick. TEM analysis shows there is significant grain refinement through mechanical

recrystallization, again corroborating Gupta and Kumar’s [29] and Wang and Kumar’s [22]

observations on steels with this alloy composition. The grain size is significantly smaller than

Page 129: Microstructural Evolution during Heat Treatment and High ...

113

those ahead of the band, which suggests that the grains continue to refine as the band develops.

The ferrite shows evidence for twinning, the L-generation austenite has transformed to

martensite, some T-generation austenite has transformed to martensite, and some T-generation

austenite appears to be stable and twinned.

The overall microstructural evolution matches the dynamic recrystallization theory put

forth by Xue and Gray [58,69] and Meyers et al. [51,71]. As the shear localizes, grains break into

finer and finer subgrains; with increased strain and high temperature in the most developed

portion of the band, the subgrains rotate to form fine, equiaxed, dynamically recrystallized grains

on the scale of 50-200nm. In particular, we have shown that if the original microstructure is

inhomogeneous (here a mixture of relatively small precipitates in a lath martensite matrix), the

grain refinement mechanisms for both phases appear to operate in parallel. Additionally, the

smaller austenite precipitates refine to smaller recrystallized grains than the ferrite. However,

unlike the austenite in the ballistic samples from Zhang’s experiments [25], the majority of the

austenite throughout this sample is mechanically stable. It is only within the band itself, which

represents a small percentage of the total volume, do we observe instability, and even then, some

T-generation austenite remains stable.

Zhang showed that quasi-static, tensile loading transformed half of the austenite in QLT

(9.5vol% austenite down from 18.98vol% austenite), while ballistic testing left no detectable

austenite in the impact region [25]. However, there are significant differences between Kolsky

experiments and quasi-static or even ballistic loading experiments. Talonen et al. [92] showed

that austenite strained in quasi-static tension tests was more mechanically unstable than high

strain-rate (200 s-1

) tensile loading where the austenite particle size is substantially reduced by

mechanical recrystallization. Therefore, it is reasonable to expect the austenite in QLT to more

Page 130: Microstructural Evolution during Heat Treatment and High ...

114

readily transform in tensile tests than in high-strain-rate experiments. Zhang’s ballistic field tests

have similar strain-rates as Kolsky experiments, but the loading conditions are quite different-

ballistic impact tests affect an area much larger than the projectile size, whereas the Kolsky anvil

is over four times larger than the sample. The ballistic plate can thus dissipate the impact energy

over a much larger area than the Kolsky sample. Therefore, while the strain-rate experienced in

both types of testing may be in the dynamic regime, energy dissipation is likely very different in

the two cases and direct comparison is tenuous.

Nevertheless, the results from this effort provide microstructural justifications for the

superior high strain-rate deformation response of 10Ni QLT. Unstable shear localization occurs

when thermal softening overrides work hardening. Here, several strain-hardening mechanisms

enabled by the thermally stable austenite content in the steel are observed in this material: the

precipitated austenite in this alloy which begins with varied composition (Ni content), size

distribution, and a low defect morphology has been shown to multiply dislocations, form twins,

and eventually experience deformation-induced phase transformation. Each of these

mechanisms, observed in the Kolsky sample, increases the overall work-hardening capacity of

the material, thereby increasing the resistance to ASB formation and propagation.

Page 131: Microstructural Evolution during Heat Treatment and High ...

115

Chapter 7: Conclusions

7.1: Microstructural Evolution during Heat Treatment of an Fe-10Ni-0.1C Steel

C Partitioning During Tempering

i) In the austenitized and quenched condition, roughly half of the C content in the alloy

is in the martensitic matrix; the rest remains tied to the refractory elements in the

form of undissolved Mo- and V-rich refractory MC and M2C carbides. C-segregation

to interfaces was noted in APT specimens of the as-quenched material.

ii) After 40 minutes of L (650°C) temper, the ferrite was entirely depleted in C. Some of

this C partitioned to austenite, while the rest partitioned to Mo-rich carbide particles.

iii) Negligible C was found in both austenite and ferrite in specimens subjected to heat

treatments such as QLT, Q25L, Q25T’ and Q125T’. The decrease in C content in

austenite is believed to be due to a combination of a dilution effect as the austenite

grows and the further precipitation and growth of carbide particles during extended

tempering.

Together these findings imply that the role of C in influencing precipitated austenite stability

in this alloy, if any, is restricted to the early stages of the L temper. In addition, it likely plays

an important role in the nucleation of the austenite particles when tempering commences.

Page 132: Microstructural Evolution during Heat Treatment and High ...

116

Ni Partitioning During Tempering

i) Ni-rich austenite is found to precipitate during isothermal tempering. At 650°C, the

austenite is found to contain 15-17 wt.% Ni; at 540°C, the austenite is more Ni-rich

and incorporates 22-24 wt.% Ni.

ii) For both isothermal tempering treatments, L and T’, the austenite is found to be

thermally unstable after extended tempering times. The onset of thermal instability is

believed to be due to the increased precipitate size, and so precipitates from shorter

tempers are presumed to be primarily stabilized by their small size.

iii) The two-stage QLT treatment is found to have two chemically distinct generations of

austenite. The Ni content in thermally stable austenite from the L-generation does not

equilibrate during the T temper (within the allocated time at this temperature), and so

both the 15-17 wt.% Ni austenite from the L temper and the 20-22 wt.% Ni austenite

from the T temper are present in the final microstructure.

iv) The austenite in QLT is also believed to be primarily thermally stabilized by the fine

particle size.

7.2: Microstructural Evolution during High Strain-Rate Deformation of an Fe-10Ni-0.1C

Steel Subjected to a Two-Stage Heat Treatment

i) The Fe-10Ni-0.1C steel was subjected to a range of strains and strain rates using a

compression Kolsky-bar testing facility. Different extents of shear localization were

generated by varying the test parameters, including no localization, diffuse ASB, and

well-developed ASB.

ii) Well-developed shear bands were found to be of the ‘transformed’ type and had

increased hardness across their width.

Page 133: Microstructural Evolution during Heat Treatment and High ...

117

iii) Microstructural analysis at various locations relative to a well-developed shear band

was performed, including: far from the shear band (homogenous plastic deformation

zone), ahead of the shear band, adjacent to the shear band, and in the shear band.

a. Far from the shear band, there was no evidence of deformation-induced

transformation of precipitated austenite to martensite.

b. Ahead of the band, the austenite precipitates were found to have broken into

subgrains. Both the L-generation and T-generation austenite were found to be

mechanically stable after subgrain formation.

c. Adjacent to the shear band, the ferrite had a lamellar morphology parallel to the

shear plane. Former austenite precipitates are also seen to be sheared parallel to

the band and had broken into fine subgrains.

d. In the shear band, the ferrite was found to have dynamically recrystallized as

equiaxed grains between 50-300nm in size. Twinning was observed in some

ferrite grains. The austenite precipitates were found drawn into strands between

100-200nm in length, and within individual strands, to have recrystallized into

equiaxed subgrains of about 20-50nm in diameter. L-generation austenite was

found to be mechanically unstable; twinning was observed in some stable T-

generation austenite grains; other T-generation grains were observed to have

mechanically transformed to martensite.

Page 134: Microstructural Evolution during Heat Treatment and High ...

118

Chapter 8: Recommendations for Future Work

There are still some open questions, answers to which would further improve our

understanding of the response of this alloy to heat treatment and deformation. These are

identified below.

It was observed that the as-quenched Fe-10Ni-0.1C steel had C enrichment at martensite

interfaces. Similar observations have been previously made in other steels [34–36]; specifically,

if these interfaces are rich enough in C, they may remain as stable austenite after the quench [34–

36]. The presence of nanoscale residual austenite films was not examined in this study, but if

present, may play an important role in subsequent tempering, for example as a nucleation site for

Ni-rich austenite. This aspect requires further attention.

Furthermore, it was observed that for a short tempering time (40 mins at 650°C), the Ni-

rich precipitated austenite appeared somewhat richer (or equivalent) in C relative to the

untempered martensite, suggesting that C quickly partitions from the martensite to the

precipitated austenite in the early stages of tempering. However, the rate and extent to which C

partitions in the early stages of tempering was not addressed in this study in detail.

A study of microstructure evolution using short tempering times (e.g. 1-30 minutes)

could address the above points, namely: i) the potential role of C-rich retained austenite films on

subsequent Ni-rich austenite evolution, and ii) quantify the rate and extent of C partitioning in

early tempering as a function of tempering time and temperature.

It was found that the Ni content in the austenite phase was spatially uniform for the T’

tempers but was quite varied in the Q25L sample. The origin of this behavior is not currently

Page 135: Microstructural Evolution during Heat Treatment and High ...

119

understood. Analysis of a wider range of L tempering times could conceivably provide further

insight.

The relationship between austenite size and thermal stability was qualitatively determined

in this study but has not been quantitatively assessed. It was determined that the austenite

examined in this study (i.e. with Ni content between 15-25 wt.%) is unstable if sufficiently large,

but the critical size as a function of Ni content remains unknown. This aspect is relevant if the

volume fraction of austenite is to be controlled through modified QLT-type heat treatments.

Additionally, the composition of austenite in the various tempers was measured using APT and

STEM EDS, but a size-independent, composition-based model for predicting Ms temperature

was not used (e.g. as done by Jain et al. [28]). Similar analysis on the compositions measured in

the isothermal samples in this study may provide further insight for future alloy design.

Several aspects of the microstructural evolution of 10Ni QLT and similar alloys during

deformation are still not well understood. Zhang showed the austenite volume fraction in 10Ni

QLT decreased by about 50% when deformed by quasi-static tensile tests [25]. However, in light

of the detailed microstructural analysis done by Wang et al. on Mn-containing TRIP steels

[39,40], similar experiments could be done on 10Ni QLT. Specifically, they determined that

different sizes of austenite particles with similar composition undergo different deformation

mechanisms (TWIP vs TRIP) during quasi-static deformation in a ‘spectral TRIP effect’ [39,40].

Additionally, Yuan et al. also proposed a composition-based spectral TRIP effect in C-alloyed

steel [41]. This study has shown that the austenite in 10Ni QLT varies both in size and

composition, and therefore may also undergo a spectral TRIP effect during quasi-static

deformation. However, detailed analysis of microstructural evolution during quasi-static

Page 136: Microstructural Evolution during Heat Treatment and High ...

120

deformation has not been performed and is likely to yield more fundamental insights into the

deformation behavior of this family of steels.

This study only examined the microstructural evolution in 10Ni QLT during dynamic

deformation. A comparative Kolsky-bar experiment with similar in-depth microstructural

analysis of the alloy subjected to single-stage L, T, and T’ temper treatments could help refine

and isolate primary deformation mechanisms that affect performance improvements.

Page 137: Microstructural Evolution during Heat Treatment and High ...

121

References

[1] G. Spanos, A.B. Geltmacher, A.C. Lewis, J.F. Bingert, M. Mehl, D. Papaconstantopoulos,

Y. Mishin, A. Gupta, P. Matic, Mater. Sci. Eng. A 452–453 (2007) 558–568.

[2] B. Fultz, J.I. Kim, Y.H. Kim, H.J. Kim, G.O. Fior, J.W. Morris, Metall. Trans. A 16 (1985)

2237–2249.

[3] S. Jin, J.W. Morris, V.F. Zackay, Metall. Trans. A 6 (1975) 141–149.

[4] S. Jin, S.K. Hwang, J.W. Morris, Metall. Trans. A 6 (1975) 1569–1575.

[5] S. Jin, S.K. Hwang, J.W. Morris, Metall. Trans. A 6 (1975) 1721–1726.

[6] S.-K. Hwang, S. Jin, J.W. Morris, Metall. Trans. A 6 (1975) 2015–2021.

[7] C.K. Syn, S. Jin, J.W. Morris, Metall. Trans. A 7A (1976) 1827–1832.

[8] C.K. Syn, B. Fultz, J.W. Morris, Metall. Trans. A 9 (1978) 1635–1640.

[9] K.J. Kim, L.H. Schwartz, Mater. Sci. Eng. 33 (1978) 5–20.

[10] J.I. Kim, J.W. Morris, Metall. Trans. A 11 (1980) 1401–1406.

[11] J.R. Strife, D.E. Passoja, Metall. Trans. A 11 (1980) 1341–1350.

[12] J.I. Kim, J.W. Morris, Metall. Trans. A 12 (1981) 1957–1963.

[13] J.I. Kim, C.K. Syn, J.W. Morris, Metall. Trans. A 14 (1983) 93–103.

[14] J.I. Kim, H.J. Kim, J.W. Morris, Metall. Trans. A 15 (1984) 2213–2219.

[15] B. Fultz, J.W. Morris, Metall. Trans. A (1985) 173–177.

[16] B. Fultz, J.W. Morris, Metall. Trans. A 16 (1985) 2251–2256.

[17] B. Fultz, J.I. Kim, Y.H. Kim, J.W. Morris, Metall. Trans. A 17 (1986) 967–972.

[18] C. Hayzelden, B. Cantor, Acta Metall. 34 (1986) 233–242.

[19] X.J. Zhang, E.J. Czyryca, in: W.O. Soboyejo, J.J. Lewandowski, R.O. Ritchie (Eds.), Mech.

Mech. Fract. John Knott Symp., 2002, pp. 295–300.

[20] D. Jain, D.N. Seidman, E.J. Barrick, J.N. DuPont, Metall. Mater. Trans. A 49 (2018) 1031–

1043.

[21] D. Jain, D. Isheim, D.N. Seidman, Metall. Mater. Trans. A 48 (2017) 3205–3219.

[22] P. Wang, K.S. Kumar, Mater. Sci. Eng. A 519 (2009) 184–197.

[23] R.W. Fonda, G. Spanos, Metall. Mater. Trans. A 45 (2014) 5982–5989.

[24] E.J. Barrick, J.N. DuPont, Mater. Sci. Eng. A 748 (2019) 189–204.

[25] X.J. Zhang, Mater. Sci. Technol. 28 (2012) 818–822.

[26] D. Jain, D. Isheim, A.H. Hunter, D.N. Seidman, Metall. Mater. Trans. A 47 (2016) 3860–

3872.

[27] D. Isheim, A.H. Hunter, X.J. Zhang, D.N. Seidman, Metall. Mater. Trans. A 44 (2013)

3046–3059.

[28] D. Jain, D. Isheim, X.J. Zhang, G. Ghosh, D.N. Seidman, Metall. Mater. Trans. A 48 (2017)

3642–3654.

[29] R. Gupta, K.S. Kumar, in: DYMAT 2009 - 9th Int. Conf. Mech. Phys. Behav. Mater. Dyn.

Load., EDP Sciences, Brussels, Belgium, 2009, pp. 1031–1037.

[30] L.J. Swartzendruber, V.P. Itkin, C.B. Alcock, J. Phase Equilibria 13 (1992) 585–585.

[31] A.D. Romig, J.I. Goldstein, Metall. Trans. A 9 (1978) 1599–1609.

[32] (n.d.).

[33] C.W. Marschall, R.F. Hehemann, A.R. Troiano, Trans. Am. Soc. Met. 55 (1962) 135–148.

[34] D.H. Sherman, S.M. Cross, S. Kim, F. Grandjean, G.J. Long, M.K. Miller, Metall. Mater.

Trans. A 38 (2007) 1698–1711.

Page 138: Microstructural Evolution during Heat Treatment and High ...

122

[35] S. Morito, K. Oh-ishi, K. Hono, T. Ohba, ISIJ Int. 51 (2011) 1200–1202.

[36] L. Morsdorf, C.C. Tasan, D. Ponge, D. Raabe, Acta Mater. 95 (2015) 366–377.

[37] J. Speer, D.K. Matlock, B.C. De Cooman, J.G. Schroth, Acta Mater. 51 (2003) 2611–2622.

[38] S. Takaki, K. Fukunaga, J. Syarif, T. Tsuchiyama, Mater. Trans. 45 (2004) 2245–2251.

[39] M.-M. Wang, C.C. Tasan, D. Ponge, A. Kostka, D. Raabe, Acta Mater. 79 (2014) 268–281.

[40] M.-M. Wang, C.C. Tasan, D. Ponge, D. Raabe, Acta Mater. 111 (2016) 262–272.

[41] L. Yuan, D. Ponge, J. Wittig, P. Choi, J.A. Jiménez, D. Raabe, Acta Mater. 60 (2012)

2790–2804.

[42] O. Dmitrieva, D. Ponge, G. Inden, J. Millán, P. Choi, J. Sietsma, D. Raabe, Acta Mater. 59

(2011) 364–374.

[43] S. Floreen, ed., in: Met. Handb. 9th Ed. Vol. 4 Heat Treat., 9th ed., American Society for

Metals, Metals Park, Ohio, 1991, p. 131.

[44] A. Marchand, J. Duffy, J. Mech. Phys. Solids 36 (1988) 251–283.

[45] J.H. Giovanola, Mech. Mater. 7 (1988) 59–71.

[46] M. Zhou, A.J. Rosakis, G. Ravichandran, J. Mech. Phys. Solids 44 (1996) 981–1006.

[47] M. Zhou, G. Ravichandran, A.J. Rosakis, J. Mech. Phys. Solids 44 (1996) 1007–1032.

[48] P.R. Guduru, A.J. Rosakis, G. Ravichandran, Mech. Mater. 33 (2001) 371–402.

[49] C. Zener, J.H. Hollomon, J. Appl. Phys. 15 (1944) 22–32.

[50] B. Hwang, S. Lee, Y.C. Kim, N.J. Kim, D.H. Shin, Mater. Sci. Eng. A 441 (2006) 308–320.

[51] M.A. Meyers, Y.B. Xu, Q. Xue, M.T. Pérez-Prado, T.R. McNelley, Acta Mater. 51 (2003)

1307–1325.

[52] S.M. Walley, Metall. Mater. Trans. A 38 (2007) 2629–2654.

[53] K. Cho, Y.C. Chi, J. Duffy, Metall. Trans. A 21A (1990) 1161–1175.

[54] H.C. Rogers, Annu. Rev. Mater. Sci. 9 (1979) 283–311.

[55] R.W. Armstrong, S.M. Walley, Int. Mater. Rev. 53 (2008) 105–128.

[56] Y.B. Xu, Y.L. Bai, Q. Xue, L.T. Shen, Acta Mater. 44 (1996) 1917–1926.

[57] Y. Xu, J. Zhang, Y. Bai, M.A. Meyers, Metall. Mater. Trans. A 39 (2008) 811–843.

[58] Q. Xue, G.T. Gray, Metall. Mater. Trans. A 37 (2006) 2435–2446.

[59] A. Belyakov, K. Tsuzaki, H. Miura, T. Sakai, Acta Mater. 51 (2003) 847–861.

[60] M.A. Meyers, U.R. Andrade, A.H. Chokshi, Metall. Mater. Trans. A 26 (1995) 2881–2893.

[61] A.G. Odeshi, S. Al-ameeri, M.N. Bassim, J. Mater. Process. Technol. 162–163 (2005) 385–

391.

[62] Q. Wei, L. Kecskes, T. Jiao, K.T. Hartwig, K.T. Ramesh, E. Ma, Acta Mater. 52 (2004)

1859–1869.

[63] M.E. Backman, S.A. Finnegan, in: R.W. Rohde, B.M. Butcher, J.R. Holland, C.H. Karnes

(Eds.), Metall. Eff. High Strain Rates, Springer US, Boston, MA, 1973, pp. 531–543.

[64] K.T. Ramesh, in: Springer Handb. Exp. Solid Mech., Springer, 2008, pp. 929–960.

[65] U.S. Lindholm, J. Mech. Phys. Solids 12 (1964) 317–335.

[66] F.J. Zerilli, R.W. Armstrong, J. Appl. Phys. 61 (1987) 1816–1825.

[67] B.K. Kad, J.-M. Gebert, M.T. Perez-Prado, M.E. Kassner, M.A. Meyers, Acta Mater. 54

(2006) 4111–4127.

[68] C.O. Mgbokwere, S.R. Nutt, J. Duffy, Mech. Mater. 17 (1994) 97–110.

[69] Q. Xue, G.T. Gray, Metall. Mater. Trans. A 37 (2006) 2447–2458.

[70] S.P. Timothy, Acta Metall. 35 (1987) 301–306.

[71] M.A. Meyers, V.F. Nesterenko, J.C. LaSalvia, Q. Xue, Mater. Sci. Eng. A 317 (2001) 204–

225.

Page 139: Microstructural Evolution during Heat Treatment and High ...

123

[72] M.A. Meyers, G. Subhash, B.K. Kad, L. Prasad, Mech. Mater. 17 (1994) 175–193.

[73] H.-R. Pak, C.L. Wittman, M.A. Meyers, in: L.E. Murr, K.P. Staudhammer, M.A. Meyers

(Eds.), Metall. Appl. Shock-Wave High-Strain-Rate Phenom., M. Dekker, New York, 1986,

pp. 749–760.

[74] X. Teng, T. Wierzbicki, H. Couque, Mech. Mater. 39 (2007) 107–125.

[75] K. Thompson, D. Lawrence, D.J. Larson, J.D. Olson, T.F. Kelly, B. Gorman,

Ultramicroscopy 107 (2007) 131–139.

[76] B. Gault, A.J. Breen, Y. Chang, J. He, E.A. Jägle, P. Kontis, P. Kürnsteiner, A.

Kwiatkowski da Silva, S.K. Makineni, I. Mouton, Z. Peng, D. Ponge, T. Schwarz, L.T.

Stephenson, A. Szczepaniak, H. Zhao, D. Raabe, J. Mater. Res. 33 (2018) 4018–4030.

[77] H. Zhao, F. De Geuser, A. Kwiatkowski da Silva, A. Szczepaniak, B. Gault, D. Ponge, D.

Raabe, Acta Mater. 156 (2018) 318–329.

[78] M. Gouné, F. Danoix, S. Allain, O. Bouaziz, Scr. Mater. 68 (2013) 1004–1007.

[79] C. Zhu, X.Y. Xiong, A. Cerezo, R. Hardwicke, G. Krauss, G.D.W. Smith, Ultramicroscopy

107 (2007) 808–812.

[80] Y. Toji, G. Miyamoto, D. Raabe, Acta Mater. 86 (2015) 137–147.

[81] Y. Toji, H. Matsuda, M. Herbig, P.-P. Choi, D. Raabe, Acta Mater. 65 (2014) 215–228.

[82] A.J. Clarke, J.G. Speer, M.K. Miller, R.E. Hackenberg, D.V. Edmonds, D.K. Matlock, F.C.

Rizzo, K.D. Clarke, E. De Moor, Acta Mater. 56 (2008) 16–22.

[83] J.G. Speer, D.V. Edmonds, F.C. Rizzo, D.K. Matlock, Curr. Opin. Solid State Mater. Sci. 8

(2004) 219–237.

[84] M. Belde, H. Springer, G. Inden, D. Raabe, Acta Mater. 86 (2015) 1–14.

[85] E.J. Seo, L. Cho, B.C. De Cooman, Acta Mater. 107 (2016) 354–365.

[86] A.J. Clarke, M.K. Miller, R.D. Field, D.R. Coughlin, P.J. Gibbs, K.D. Clarke, D.J.

Alexander, K.A. Powers, P.A. Papin, G. Krauss, Acta Mater. 77 (2014) 17–27.

[87] L. Morsdorf, O. Jeannin, D. Barbier, M. Mitsuhara, D. Raabe, C.C. Tasan, Acta Mater. 121

(2016) 202–214.

[88] W. Sha, L. Chang, G.D.W. Smith, Liu Cheng, E.J. Mittemeijer, Surf. Sci. 266 (1992) 416–

423.

[89] M. Thuvander, J. Weidow, J. Angseryd, L.K.L. Falk, F. Liu, M. Sonestedt, K. Stiller, H.-O.

Andrén, Ultramicroscopy 111 (2011) 604–608.

[90] Z. Peng, D. Zanuttini, B. Gervais, E. Jacquet, I. Blum, P.-P. Choi, D. Raabe, F. Vurpillot,

B. Gault, J. Phys. Chem. Lett. 10 (2019) 581–588.

[91] Z. Peng, F. Vurpillot, P.-P. Choi, Y. Li, D. Raabe, B. Gault, Ultramicroscopy 189 (2018)

54–60.

[92] J. Talonen, H. Hänninen, P. Nenonen, G. Pape, Metall. Mater. Trans. A 36 (2005) 421–432.


Recommended