The effect of driving force in Gibbs energy on the fraction of martensite

Similar documents
Combining bainite and martensite in steel microstructures for light weight applications

Influence of Vanadium and Tungsten on the Bainite start temperature

Phase Transformations in Metals Tuesday, December 24, 2013 Dr. Mohammad Suliman Abuhaiba, PE 1

Institutional repository of Jönköping University

Heat treatment and effects of Cr and Ni in low alloy steel

Chapter 10: Phase Transformations

EXPERIMENT 6 HEAT TREATMENT OF STEEL

Lecture 31-36: Questions:

Heat transfer coefficient and latent heat of martensite in a medium-carbon steel

Precipitation Hardening. Outline. Precipitation Hardening. Precipitation Hardening

Designing martensitic steels: structure & properties Enrique Galindo-Nava and Pedro Rivera

Martensite Formation in Partially and Fully Austenitic Plain Carbon Steels

Chapter 10, Phase Transformations

Experiment E: Martensitic Transformations

A DISLOCATION MODEL FOR THE PLASTIC DEFORMATION OF FCC METALS AN ANALYSIS OF PURE COPPER AND AUSTENITIC STEEL

J = D C A C B x A x B + D C A C. = x A kg /m 2

MTLS 4L04 Steel Section. Lecture 6

The Iron Iron Carbide (Fe Fe 3 C) Phase Diagram

Features of Carbide Precipitation During Tempering of 15H2NMFA and 26HN3M2FA Steels

Titanium and titanium alloys. Josef Stráský

Correlation of the Effects of Alloying Elements on the Hardenability of Steels to the Diffusion Coefficients of Elements in Fe

Master examination. Metallic Materials

Introduction: Ferrous alloys - Review. Outline. Introduction: Ferrous alloys

SIMULATION OF DIFFUSIONAL PROCESSES DURING SOLIDIFICATION IN AUSTENITIC STEELS

PHASE TRANSFORMATIONS ON STEEL DURING INDUCTION HARDENING

Effect of the austenitising heat treatment on the microstructure and hardness of martensitic stainless steel AISI 420

The Cementite Spheroidization Process in High-Carbon Steels with Different Chromium Contents

AISI 304 steel: anomalous evolution of martensitic phase following heat treatments at 400 C

Abstract INTRODUCTION

Phase transformation kinetics and microstructure of NiTi shape memory alloy: effect of hydrostatic pressure

The heat treatment of austenitic stainless steel wire with different work hardening degree

This article appeared in a journal published by Elsevier. The attached copy is furnished to the author for internal non-commercial research and

Heat Treating Distortion and Residual Stresses

THE EFFECT OF ALLOYING ELEMENTS ON THE TEMPERATURE RANGE OF PEARLITE TO AUSTENITE TRANSFORMATION IN LOW ALLOY HYPOEUTECTOID STEELS

Cu/Ag Eutectic System

MATERIALS SCIENCE-44 Which point on the stress-strain curve shown gives the ultimate stress?

CHAPTER 4 MODELING OF PHASE TRANSFORMATION IN INDUCTION HARDENING

Keywords: Dual phase steel, martensite, ferrite, intercritical annealing, quenching, K, n

EFFECT OF POST SINTERING THERMAL TREATMENTS ON DIMENSIONAL PRECISION AND MECHANICAL PROPERTIES IN SINTER-HARDENING PM STEELS

CHAPTER 9 PHASE DIAGRAMS

New functionalities of TC-DICTRA

E-BRITE E-BRITE. Technical Data Sheet. Stainless Steel: Superferritic GENERAL PROPERTIES PLANAR SOLID OXIDE FUEL CELLS CHEMICAL COMPOSITION

Modeling of Ferrite-Austenite Phase Transformation Using a Cellular Automaton Model

The Concept Of Weldability Of Metals

Lecture 11: Metallic Alloys

Phase Diagrams of Pure Substances Predicts the stable phase as a function of P total and T. Example: water can exist in solid, liquid and vapor

Hot Deformation Behavior of High Strength Low Alloy Steel by Thermo Mechanical Simulator and Finite Element Method

Lab Materials Science

Phase Investigation of Austempered Ductile Iron

Heat Treatment of Steel Lab Report. Justin Lance 11/16/2011 Engineering 45 Lab Section 3 Troy Topping

Solid solution strengthening effect on creep strength of austenitic stainless steel

1. Use the Ellingham Diagram (reproduced here as Figure 0.1) to answer the following.

PRELIMINARY INVESTIGATIONS OF LOW-NICKEL STAINLESS STEELS FOR STRUCTURAL APPLICATIONS

CHAPTER 9 PHASE DIAGRAMS PROBLEM SOLUTIONS

Ferrous Alloys. Metal Alloys. Ferrous. Non ferrous. Grey iron. Carbon Low Alloy High Alloy. Nodular iron White iron Malleable iron Alloy cast irons

THE MICROSTRUCTURES AND HARDNESS ANALYSIS OF A NEW HYPOEUTECTOID STEELS WITH 0.3 % MO. Edyta ROŻNIATA

An Investigation of Microstructural Change of Low Alloy Steel AISI 4150 by Seebeck Coefficient

MICROMECHANISMS OF CLEAVAGE FRACTURE IN THE HAZ OF C-MN COMMERCIAL STEEL WELD

Sensitization & Corrosion Behaviour of Austenitic Stainless Steel 304 & 316

Creep and High Temperature Failure. Creep and High Temperature Failure. Creep Curve. Outline

Nanoscale austenite reversion through partitioning, segregation, and kinetic freezing: Example of a ductile 2 GPa Fe-Cr-C steel

Microstructural Evolution of Secondary Phases in the Cast Duplex Stainless Steels CD3MN and CD3MWCuN

Effect of Heat Treatment on Microstructure and Mechanical Properties of Medium Carbon Steel

Part IV : Solid-Solid Phase Transformations I Module 3. Eutectoid transformations

XRD and TEM analysis of microstructure in the welding zone of 9Cr 1Mo V Nb heat-resisting steel

CHAPTER 9 PHASE DIAGRAMS PROBLEM SOLUTIONS

Retained Austenite: Transformation-Induced Plasticity

Mat E 272 Lecture 19: Cast Irons

Module #25. Martensitic Transformations and Strengthening

Development of TBF Steels with 980 MPa Tensile Strength for Automotive Applications: Microstructure and Mechanical Properties

Torsional Fatigue Performance of Induction Hardened 1045 and 10V45 Steels

THE CARBON CONTENT OF LOW CARBON MARTENSITE

Available online at ScienceDirect. Procedia CIRP 14 (2014 )

APPLICATIONS OF Fe-C PHASE DIAGRAM

Numerical Analysis of Heat Conduction and Phase. Transformation in Laser Transformation. Hardening: Influences of Heating Duration and

NEW OBSERVATIONS ON FLASH BAINITE. Petch Janbanjong. Bachelor of Engineering in Mechanical Engineering, KMITL, Thailand, 2011

Copyright 1999 Society of Manufacturing Engineers. FUNDAMENTAL MANUFACTURING PROCESSES Heat Treating NARRATION (VO): RESISTANCE OF METALS AND ALLOYS.

Postprint.

Finite element analysis of residual stress in the welded zone of a high strength steel

Phase Diagram Applications

Seam Welded Air-Hardenable Corrosion Resistant Steel Tubing: Automotive Applications Overview

Carbon Concentration of Austenite in Austempered Ductile Iron (ADI)

Technical Reference on Hydrogen Compatibility of Materials

CHAPTER 5 IMPERFECTIONS IN SOLIDS PROBLEM SOLUTIONS ev /atom = exp. kt ( =

Types of stainless steel

Uniaxial Ratcheting Behaviors of Metals with Different Crystal Structures or Values of Fault Energy: Macroscopic Experiments

DLPP DANIELI LONG PRODUCTS PROPERTIES PREDICTOR. ITA Ltd. Ostrava, Czech republic. Danieli Morgardshammar Buttrio, Italy

Diode laser beam absorption in laser transformation hardening of low alloy steel

Hot Forging Behaviour of Sintered AISI 8720 P/M Steels

Properties of Carbon Steel Shot

Non-Magnetic Stainless Steels Reinvestigated a Small Effective Field Component in External Magnetic Fields

Transactions on Engineering Sciences vol 17, 1997 WIT Press, ISSN

Microstructures of Mild Steel Spring after Heat Treatment.

Title: Modeling of microstructure in the HAZ for microalloyed steel S700 MC

Deformation Induced Martensitic Transformation In 304 Stainless Steels

Advances in steel and Al alloy materials High strength (HSS) & Advanced high strength steel (AHSS)

Secondary Graphite Formation in Tempered Nodular Cast Iron Weldments

STAINLESS STEELS. Chromium and nickel content in the various groups of stainless steels

Benefits of thermodynamic and microstructure simulations towards reliable continuous casting of steels

Predicting Distortion and Residual Stress in a Vacuum Carburized and Gas Quenched Steel Coupon

Transcription:

The effect of driving force in Gibbs energy on the fraction of martensite Erik Andersson Andreas Johansson Supervisor: Associate Prof. Annika Borgenstam 2013 Dept. of Material Science and Engineering Royal Institute of Technology, KTH Stockholm, Sweden

Abstract The background to this bachelor thesis is an on-going project within the VINN Excellence Center Hero-m. The task in this thesis is to perform a literature survey about the martensite transformation and investigate how the resulting fraction depends on cooling below the M s -temperature. Instead of calculating the undercooling for each of the known fractions of martensite the driving force will be evaluated. Several efforts have been made through the years to describe the relationships between fraction transformed austenite and temperature. The approaches to the first models were empirical and derived from collections of data regarding the amount of retained austenite at different quenching temperatures. Lately, studies have been made to derive a thermodynamical relationship using how the Gibbs energy is affected by increments in volume transformed austenite. Two equations are derived by calculating the resulting driving force at different known quenching temperatures and the respective percentage transformed martensite found in previous works. The data for the steels used show a characteristic slope when linearised. A trend for the steels which have a high characteristic slope is that they also have a high M s temperature, and the steels which have a low characteristic slope tend to have a low M s. Previous relationships which describe the martensitic transformation have considered the importance of the M s temperature only in it being a starting temperature for the transformation. To further incorporate the M s temperature in the equations presented, further research of the martensitic transformation is required. The approach in this thesis of using thermodynamically calculated data is a base for further investigation of the range of the martensite transformation. Keywords: martensite, model, driving force, Gibbs energy, fraction, M s -temperature

Table of Contents Introduction... 1 Previous works... 2 Empirical models... 2 Thermodynamical approach... 4 A new approach to a different model... 4 Calculations and Results... 5 Discussion and conclusions... 12 Acknowledgements... 13 References... 14

Introduction The austenite to martensite transformation is utilized in hardening of steels. The martensitic structure is hard and brittle but can be tempered to become more ductile without losing too much of its hardness. The transformation is diffusionless and occurs very rapidly, the growth rate can be up to 1000 m/s. The utilized phenomenon in the hardening process is the change in solubility of carbon in the different phases, as austenite has a much higher solubility than ferrite. As the transformation from austenite to ferrite is initialized, the change in carbon content is hindered and the result is a BCT-structure, which is an elongated BCC-ferritic structure where the carbon atoms distort the otherwise cubic lattice. The transformation begins at a specific temperature which is denoted the martensitic start temperature, M s. For most alloying elements M s is shifted to lower temperatures as the alloying content increases and the austenite is stabilized. Due to the nature of the martensitic transformation, the inner shear stresses associated with the transformation are increased with higher carbon content and this increases the driving force needed for the transformation. At a specific temperature T 0, the driving force in Gibbs energy for martensite and austenite is the same. Below this temperature the transformation is thermodynamically possible but for nucleation of martensite an increase in driving force is needed to overcome the non-chemical energy associated with the shear stresses. To achieve this increase in driving force the temperature is lowered by further cooling down to M s. At M s the degree of driving force needed for nucleation of martensite is reached. By further cooling the transformation increases until it reaches total transformation at a temperature denoted M f where all austenite is transformed. The transformation process is schematically shown in Fig. 1. As the martensite is thermodynamically unstable it is not possible to calculate any driving force thermodynamically. The phase is approximated to have the same Gibbs energy as α-ferrite. The background to this bachelor thesis is an on-going project within the VINN Excellence Center Hero-m, where a new steel is designed. The new steel will have a martensitic based structure; therefore it is important to know how the fraction of martensite varies when the steel is cooled below the M s -temperature at different temperature intervals. It is of interest to use a more theoretical approach with thermodynamically based calculations, such as the driving force in Gibbs energy. The relationship between the fraction of martensite and Gibbs energy will be implemented in a software model to design new steels. The task in this thesis is to perform a literature survey about the martensite transformation and investigate how the resulting fraction depends on cooling below the M s -temperature. Instead of calculating the undercooling the driving force will be evaluated for each of the known fractions of martensite. This is done by thermodynamical calculations using the TCFE6 database in 1

Thermo-Calc [1]. This data is a base for evaluating an equation describing the fraction of martensite formed from austenite on cooling below M s. Fig. 1 A schematic figure from Thermo-Calc [1] which shows the Gibbs energy for α-ferrite and austenite. The whole transformation range is presented from thermodynamical equilibrium at T 0, initial martensite transformation at M s and full transformation at M f. The driving force for the martensitic transformation is the difference in Gibbs energy between the two phases. Previous works Empirical models Several efforts have been made through the years to describe the relationships between the fraction transformed austenite and temperature. The first models were empirical and derived from collections of data regarding the amount of retained austenite at different temperatures below M s. Lately, studies have been made to derive thermodynamical relationships relating the Gibbs energy to volume transformed austenite. The first empirical model was described by Harris and Cohen [2]. They investigated four steels with the same carbon content but with variations in chromium and nickel content using lineal analysis to 2

determine the fraction of martensite formed. Lineal analysis uses a randomly placed line on the metallographic sample and the proportion of the phases along the line is measured. The line fraction is used as an approximation of the volume fraction. From the measured data, they determined the following equation [ ] 5,32 P= M T [Eq. 1] 15 100 6,95 10 455 ( S ) to describe the amount of transformed austenite. They also assumed that the relationship is not affected by varying variables such as chemical composition, austenitizing treatments, presence of undissolved carbides and variations in grain size. Harris and Cohen assumed that the variations in the aforementioned variables only affect the resulting M s and that the transformation follows [Eq. 1]. The M s -values used in this study were extrapolated from the martensite transformation curves. Another model was presented by Koistinen and Marburger [3]. They investigated the martensitic transformation in pure iron-carbon alloys and plain carbon steels. The amount of retained austenite was measured by x-ray diffraction. In this model, the only variable is quenching temperature and martensitic starting temperature. The martensitic starting temperature used by Koistinen and Marburger was derived from previous works. The volume fraction austenite was plotted logarithmically against the undercooling below the M s. The following equation was derived by the given fractions of martensite: 2 Vγ = exp 1,10 10 ( MS Tq) M 80 S > Tq > C [Eq. 2] Where V γ is the volume fraction retained austenite, Tq is the quenching temperature. Koistinen and Marburger assumed that the effect of chemical composition, austenitizing temperature and other variables is only visible in the change of M s. Due to the nature of this equation the amount of retained austenite can never reach zero, and as a result there is no temperature in which the martensitic transformation reaches completion. 3

Thermodynamical approach H.Y Yu [4] describes a theoretical model based on thermodynamical theories where the following assumptions are made; the amount of transformation is independent of time, a fraction of martensite is formed very rapidly at a constant temperature. The amount of transformation is characteristic of temperature, provided that other variables such as grain size are held constant. Martensite is formed at cooling below the martensitic starting temperature at a rate that is independent of temperature. The Gibbs energy of the whole volume of the system is considered and any change in energy followed by transformed volume austenite is discussed. This change in energy is divided into two parts, the chemical which is related to Gibbs energy and non-chemical free energy which is related to the shear stresses associated with the transformation. The final equation is derived from these assumptions: ε q( Ms T) ε = M T + (1 βε ) ( T T) s q q q [Eq. 3] This equation uses existing data where a fraction of martensite is known, ε q, at a temperaturet q. The variable, β, is the ratio between the slopes of martensite and austenite Gibbs energy plotted as a function of temperature. A new approach to a different model The previous models use the undercooling, M s -T, as the variable and these models try to describe the relationship between the fraction of martensite and undercooling. In the model presented in this work where the prediction of the martensite fraction for a new steel is needed, a more theoretical approach would be of interest as there exist no previous empirical data on the fraction of martensite formed upon quenching of the steel. A suggestion based on the relationship between fraction martensite and the driving force in Gibbs energy is more relevant and this new relationship will only use Gibbs energy as a variable. All data for the calculations is available with the use of a thermodynamically based software such as Thermo-Calc using the TCFE6 database [1]. The input data for the calculations is derived from a number of existing studies. The driving force in Gibbs energy of a known temperature below M s and corresponding fraction martensite is calculated using the Gibbs energy for austenite. 4

Calculations and Results The gathered data from previous studies is used to calculate the Gibbs energy for each separate phase. As the martensite is thermodynamically unstable it is not possible to calculate any driving force thermodynamically in Thermo-Calc. The phase is approximated to have the same Gibbs energy as α-ferrite. The Gibbs energy for austenite is also calculated in Thermo-Calc. A large number of steels with a varying composition are crucial to generate a general expression for the fraction of martensite formed below M s. Data for a number of different steels is therefore gathered and presented in Table A. A wide range of carbon contents is especially important as carbon has a major influence on the transformation processes. There is also a wide range in composition in other alloying elements such as chromium and nickel. The austenitizing temperatures for all the steels are presented; the holding time for austenitizing is unknown. In Fig. 2 the Gibbs energy for the two phases is shown for a range of temperatures. Steel C Cr Ni Si Mn Cu M s ( C) Austenitizing temperature ( C) SAE- 0,35 - - 0,19 1,8-338 843 T1335 [7] SAE-2340 0,37-3,41 0,21 0,68-330 788 [7] SAE-3140 0,38 0,49 1,32 0,21 0,72-332 843 [7] Fe-0,46C 0,46 0,2 0,1 0,26 0,71 0,23 315 900 [8] SAE-1065 0,63 - - 0,22 0,87-274 816 [7] 0,9C [7] 0,89 - - 0,15 0,29-216 885 10C2N 0,92-2,66 - - - 161 1100 [6] 1,5-pct- 1,09 1,52-0,29 0,32-125 1038 Cr [2] 2,8-pct- 1,08 2,83-0,29 0,27-115 1038 Cr [2] 5,4-pct- 1,12-5,36 0,24 0,51-80 1038 Ni [2] 12C2N 1,26-2 - - - 110 1100 [6] 13C5N [6] 1,32-4,72 - - - 50 1100 Table A The gathered data from previous articles is here presented with increasing carbon content (wt%). 5

Fig. 2 Gibbs energy for phases α-ferrite, BCC_A2, and γ-austenite, FCC_A1, calculated in Thermo-Calc. The Gibbs energy is calculated for each phase and each temperature and the resulting driving force for martensite is evaluated. The driving force for the austenite to martensite transformation is taken to be the Gibbs energy for α-ferrite minus the Gibbs energy for γ-austenite. This subtraction is given the notation ΔG γ α and is plotted against the percentage of martensite in Fig. 3. % Martensite 100 90 80 70 60 50 40 30 20 10 0-4000 -3500-3000 -2500-2000 -1500-1000 -500 0 ΔG γ α (J/mole) 0,9C SAE-1065 SAE-T1335 SAE-2340 SAE-3140 1,5-pct-Cr 2,8-pct-Cr 5,4-pct-Ni 10C2N 12C2N 13C5N Fe-0,46C Fig. 3 Estimated percentage of martensite plotted against the Gibbs energy for the austenite to martensite transformation. 6

The Gibbs energy associated with reaching initial transformation does not affect the additional needed Gibbs energy to reach full completion. In Fig. 4 the data in Fig. 3 has been normalised according to Gibbs energy associated with the initial martensitic transformation. 100 90 80 70 60 50 40 30 20 10 0-2500,00-2000,00-1500,00-1000,00-500,00 0,00 % Martensite ΔG γ α - ΔG Ms (J/mole) 0,9C SAE-1065 SAE-T1335 SAE-2340 SAE-3140 1,5-pct-Cr 2,8-pct-Cr 5,4-pct-Ni 10C2N 12C2N 13C5N Fe-0,46C Fig. 4 Estimated percentage of martensite plotted against the driving force in Gibbs energy for the austenite to martensite transformation minus the Gibbs energy needed to initiate transformation. As many processes can be described as an exponential function a first approach to find the relationship is to investigate the amount retained austenite on a logarithmic scale, like previously suggested by Koistinen and Marburger [3], where a best fitting straight line is used to get the related equation. However, as seen in Fig. 5, the straight line does not fit the data in a satisfactory way. This clearly shows that the best fitting function cannot be described by a simple exponential function as the spread of the data is large. Another approach is the Gompertz function, [Eq. 4], which is used to describe processes where the rate increase is slow at the beginning and at the end of the process as clearly shown in Fig. 4: Gmprtz( x) = a exp( b exp( cx)) [Eq. 4] where the constant a is the upper asymptote, b describes the variable displacement and c gives the steepness of the slope [5]. Due to the nature of the function it is impossible to determine the constants algebraically and hence a numerical software is required. Using numerical software and the given data for all the steels, [Eq. 5] gives the best fit to all of the existing data: 3 ( ) V = 94.85exp -3.84 exp(( Ψ ) 5.13 10 ) [Eq. 5] α γ where Ψ= G - G M s 7

100 % Austenite 10 1-2500,00-2000,00-1500,00-1000,00-500,00 0,00 ΔG γ α - ΔG Ms (J/mole) 0,9C SAE-1065 SAE-T1335 SAE-2340 SAE-3140 1,5-pct-Cr 2,8-pct-Cr 5,4-pct-Ni 10C2N 12C2N 13C5N Fe-0,46C Fitting straight line Fig. 5 The best fitting straight line to the values for the estimated amount of austenite on a logarithmic scale. As can be seen, this accuracy is not satisfactory. As seen in Fig. 6 the influence of Fe-0,46 is greater than any other series as it consists of a larger number of values. It is therefore interesting to see how the equation changes when removing the data for Fe-0,46. [Eq. 6] gives the best fit to the remaining data: 3 ( ) V = 94.93exp -4.50 exp(( Ψ ) 5.19 10 ) [Eq. 6] [Eq. 5] and [Eq. 6] is plotted with the data in Fig. 6. To more clearly show the initial transformation the first thirty per cent martensite is highlighted in Fig. 7 with [Eq. 5] and [Eq. 6]. As the spread of the data in Fig. 6 is significant and the slopes of the steels vary it is of interest to find a relationship between the slope and a material constant such as M s. It is hard to find a slope in Fig. 6 because of the behaviour of the transformation. The retained austenite is used to linearize the data from Fig. 6 in Fig. 8. The data used to calculate the slopes is taken from the intervals in Fig. 6 where the behaviour of the transformation is the most linear. The different slopes are presented in Table B with the corresponding M s, where the corresponding c constant is also listed. 8

For comparison the most used equations, [Eq.1] and [Eq. 2], are plotted with the data used in the present study, but instead regarding undercooling, T-M s, in Fig. 9. 100 90 80 70 60 50 40 30 20 10 0-2250,00-1750,00-1250,00-750,00-250,00 % Martensite ΔG γ α - ΔG Ms (J/mole) 0,9C SAE-1065 SAE-T1335 SAE-2340 SAE-3140 1,5-pct-Cr 2,8-pct-Cr 5,4-pct-Ni 10C2N 12C2N 13C5N Fe-0,46C [Eq. 5] [Eq. 6] Fig. 6 The best fitting equations compared with existing data for the martensite transformation with and without Fe- 0,46C. 9

30 25 20 15 10 5 0-300,00-250,00-200,00-150,00-100,00-50,00 0,00 % Martensite ΔG γ α - ΔG Ms (J/mole) 0,9C SAE-1065 SAE-T1335 SAE-2340 SAE-3140 1,5-pct-Cr 2,8-pct-Cr 5,4-pct-Ni 10C2N 12C2N 13C5N Fe-0,46C [Eq. 5] [Eq. 6] Fig. 7 [Eq. 5] and [Eq. 6] compared with existing data for the initial martensite transformation according to the driving force. 100 % Austenite 10-700,00-600,00-500,00-400,00-300,00-200,00-100,00 0,00 ΔG γ α - ΔG Ms (J/mole) 0,9C SAE-1065 SAE-1335 SAE-2340 SAE-3140 1,5-pct-Cr 2,8-pct-Cr 5,4-pct-Ni 10C2N 12C2N 13C5N Fe-0,46C Fig. 8 The remaining austenite is logarithmically plotted against the driving force. The black lines show the best linear approximation to the data. 10

Steel Characteristic slope in Fig. 8 M s ( C) C constant SAE-T1335 0,007510 338 1.0301E-02 SAE-3140 0,006589 332 1.0644E-02 SAE-2340 0,005991 330 8.8065E-03 Fe-0,46C 0,002835 315 4.3789E-03 SAE-1065 0,003888 274 6.5227E-03 0,9C 0,005018 216 6.0858E-03 10C2N 0,002648 161 4.3792E-03 1,5-pct-Cr 0,001933 125 4.2333E-03 2,8-pct-Cr 0,001850 115 5.2420E-03 12C2N 0,002570 110 4.8278E-03 5,4-pct-Ni 0,001620 80 6.4646E-03 13C5N 0,001933 50 3.8951E-03 Table B The calculated slopes and values for the c-constant from Fig. 8 with the corresponding M s in descending order. 100 90 80 70 60 50 40 30 20 10 0-300,00-250,00-200,00-150,00-100,00-50,00 0,00 % Martensite T-Ms (K) 0,9C SAE-1065 SAE-T1335 SAE-2340 SAE-3140 1,5-pct-Cr 2,8-pct-Cr 5,4-pct-Ni 10C2N 12C2N 13C5N Fe-0,46C H.C [Eq. 1] K.M [Eq. 2] Fig. 9 [Eq. 1] and [Eq. 2] compared with existing data for the martensite transformation according to undercooling. 11

Discussion and conclusions When evaluating existing relationships previously presented by Harris and Cohen [Eq. 1] and by Koistinen and Marburger [Eq. 2] it is evident that they do not give a satisfactory approximation for the fraction of the transformed austenite as a function of temperature below M s. At small levels of undercooling the calculated amount of martensite is too high and for higher undercooling the expected amount is too low. Also, both [Eq. 1] and [Eq. 2] do not consider the exponentially increasing amounts at low levels of undercooling. [Eq. 3] has a number of constants that varies for each steel and it can only be used to evaluate one steel at a time and therefore it is not possible to compare all the steels using this equation. [Eq. 3] has therefore not been compared with the other equations. Ultimately it is preferable to have a model that considers the behaviour of the transformation for the whole range. The two relations [Eq. 5] and [Eq. 6] presented in this thesis are two equations that gives the best fit to previous data on the relationship between the martensite and the resulting driving force at different quenching temperatures. As seen in Fig. 6 the [Eq. 5] and [Eq. 6] follows the characteristic behaviour of the transformation. This is also shown in Fig. 7 where the exponential behaviour for the initial martensite transformation is captured by [Eq. 5] and [Eq. 6]. However the spread of the data makes it impossible to find an equation that describes all the values without considering other effects. In Fig. 8 it is obvious that there is a spread of the data as the different steels have different characteristic slopes. In Table B the characteristic slope is presented with the corresponding M s and the c-constant. A trend for the steels which have a high characteristic slope is that they also have a high M s temperature, and the steels which have a low characteristic slope tend to have a low M s temperature. The same characteristics are shown for the c-constant, as the constant gives the steepness of the slope. It is possible that a high M s temperature indicates a transformation that requires a lower driving force. This is also apparent by the fact that the range of driving force for the transformation is narrower than for a steel which have a lower M s temperature. Previous relationships which describe the martensitic transformation have considered the importance of the M s temperature only in which way it affects the resulting undercooling. This is also used in the presented equations [Eq. 5] and [Eq. 6]. However, the effect of the M s temperature is, as shown in Fig. 8, more extended to the behaviour of the whole transformation process; a steel with a higher M s temperature generally have a high characteristic slope and also a narrower range of transformation in driving force. To further incorporate the M s temperature in the [Eq. 5] and [Eq. 6] it is necessary to determine a relationship between the constant relating to the slope of the function and the M s temperature. The constant increases with increasing M s temperature and therefore it is reasonable that there is a proportional relationship between the slope and the M s temperature. To further investigate the martensitic transformation it is necessary to create series of samples of varying composition with a complete range of the transformation, as a few of the existing sources do not cover the whole range of the transformation. It is also crucial to evaluate the correct amount of martensite in each sample; otherwise the data will be misleading and therefore harmful for the approximation. This is also an aspect when determining the M s temperature. 12

The approach in this thesis of using thermodynamically calculated data is a good base for further investigation of the range of the martensite transformation. A first rough approximation in the here presented equations show a satisfactory behaviour, especially at lower fractions of martensite. Acknowledgements Our supervisor Annika Borgenstam is gratefully acknowledged for helpful discussions and comments. 13

References [1] J.-O. Andersson, T. Helander, L. Höglund, P. Shi, and B. Sundman, Thermo-Calc & DICTRA, computational tools for material science, Calphad, vol. 26, no. 22 [2] W. J. Harris, M. Cohen, Stabilization of the Austenite-martensite Transformation, Metals Trans. AIME (1949), vol. 180, pp. 447-470 [3] D.P. Koistinen, R.E. Marburger, A general equation prescribing the extent of the austenitemartensite transformation in pure iron-carbon alloys and plain carbon steels, ACTA Metallurgica (1959), vol. 7, pp. 59-60 [4] H.Y Yu, A New Model for the Volume Fraction of Martensitic Transformations, Metallurgical and Materials Transactions A (1997), vol. 28A, pp. 2499-2505 [5] K. Deep, Proceedings of the International Conference on Soft Computing for Problem Solving (2011), vol. 1, pp. 171 [6] E. R. Morgan, T. Ko, Thermal Stabilization of Austenite in Iron-Carbon-Nickel Alloys, ACTA Metallurgica (1953), vol. 1, pp. 36-48 [7] R. A. Grange, H. M. Stewart, The Temperature Range of Martensite Formation, Metals Technology AIME (1946) Technical Publication No. 1996, pp. 1-24 [8] S.M.C. Van Bohemen, J. Sietsma, Martensite Formation in Partially and Fully Austenitic Plain Carbon Steels Metallurgical and Materials Transactions A (2009), vol. 40A, pp. 1059-1068 14