effect of thermo-mechanical treatment on texture evolution

115
Florida State University Libraries Electronic Theses, Treatises and Dissertations The Graduate School 2005 Effect of Thermo-Mechanical Treatment on Texture Evolution of Polycrystalline Alpha Titanium Gilberto Alexandre Castello Branco Follow this and additional works at the FSU Digital Library. For more information, please contact [email protected]

Upload: others

Post on 13-Mar-2022

6 views

Category:

Documents


0 download

TRANSCRIPT

Florida State University Libraries

Electronic Theses, Treatises and Dissertations The Graduate School

2005

Effect of Thermo-Mechanical Treatment onTexture Evolution of Polycrystalline AlphaTitaniumGilberto Alexandre Castello Branco

Follow this and additional works at the FSU Digital Library. For more information, please contact [email protected]

THE FLORIDA STATE UNIVERSITY

COLLEGE OF ENGINEERING

EFFECT OF THERMO-MECHANICAL TREATMENT ON

TEXTURE EVOLUTION OF POLYCRYSTALLINE ALPHA

TITANIUM

By

GILBERTO ALEXANDRE CASTELLO BRANCO

A Dissertation submitted to the Department of Mechanical Engineering

In partial fulfillment of the requirements for the degree of

Doctor of Philosophy

Degree Awarded Summer Semester, 2005

Copyright © 2005 Gilberto Alexandre Castello-Branco

All Rights Reserved

The members of the Committee approve the dissertation of GILBERTO ALEXANDRE

CASTELLO BRANCO defended on May 16, 2005.

Hamid Garmestani

Professor Directing Dissertation

Chuk Zhang Outside Committee Member

Justin Schwartz

Committee Member

Chiang Shih

Committee Member

Approved: Chiang Shih, Chairman, Department of Mechanical Engineering Ching-Jen Chen, Dean, College of Engineering The Office of Graduate Studies has verified and approved the above named committee members.

Dedicated to my parents Gilberto and Léa, my sister Leila, my wife Cristiane

and my dear relatives Beatriz and Jorge Alberto.

iii

ACKNOWLEDGEMENTS

To God for giving me the strength to overcome all the obstacles that I have found in my

way.

I am very grateful and indebted to my advisor, Dr. Hamid Garmestani for his endless

support, encouragement and optimism during the course of this study. I would like to thank the

members of my committee.

I would like to thank my professors, Dr. Luiz Brandão, Dr Said Ahzi and Dr Anthony

Rollett for their support in several occasions during my research program. I also would like to

tank Dr. Ayman Salem, Dr. Mike Glavicic for their help and suggestions, Dr. Scott Schoenfeld

and Dr. Lee Semiatin for providing funds and the material used in this research. This study was

partially funded under the AFOSR grant # F49620-03-1-0011 and Army Research Lab contract #

DAAD17-02-P-0398, DAAD17-02-P-0928.

I am grateful to the National High Magnetic Field Laboratory (NHMFL) and

MARTECH, Tallahassee, Florida for the facilities, the Department of Material Science and

Engineering of the Georgia Institute of Technology for allowing me to use the rolling facility,

and also to several members of the NHMFL, who in one way or another contributed to the

success of my work. Especial thanks go to: Mr. Robert Goddard for his guidance and assistance

in running the ESEM/OIM facility, the FSU staff personnel, especially Mr. George Green, my

friends in Tallahassee, especially Mr. Donald Hollett and family for their kindness, friendship

and support and my research colleagues at FSU and Georgia Tech.

I would like to thank my friends in Brazil, who were always giving me support even

though the distance. A special thanks goes to my dear friend Bernardino.

Many thanks are due to my all colleagues at CEFET-RJ, for their support and

encouragement.

iv

I would like to express my profound gratitude to my parents, my sister, to Geracinda and

all my family, who have always given me their love, encouragement and endless support

throughout these years.

Finally I wish to express my heartful appreciation to my beloved wife, Cristiane, who has

always been walking by my side, sharing the good and bad moments, tirelessly helping and

encouraging me.

Gilberto Alexandre Castello Branco Florida State University, Tallahassee May, 2005

v

TABLE OF CONTENTS

LIST OF TABLES……………………………………………………………………........ ix LIST OF FIGURES……………………………………………………………………….. x ABSTRACT………………………………………………………………………………... xiii 1. INTRODUCTION……………………………………………………………………… 1 2. BACKGROUND……………………………………………………………………….. 4 2.1- Titanium and its Alloys……………………………………………………………. 4 2.1.1 - Physical metallurgy of Titanium and Titanium Alloys…………………... 6

2.1.2 - Classification of Titanium Alloys ……………………………………….… 7 2.1.2.1 - Alpha-Titanium Alloy ………………………………………………….... 7 2.1.2.2 - Near-Alpha Titanium Alloys ...…………………………………………... 8 2.1.2.3 – Alpha/Beta ( α + β ) Alloys…………………………………………….…. 8

2.1.2.4 - Beta, Near-Beta and Metastable-Beta alloys.……………………………. 9 2.2 - Mechanical Behavior of Titanium and its Alloys………………………………... 11

2.2.1 – Slip Modes in HCP Metals ……………………………………………..….. 11 2.3- Texture …………………………………………………………………………...… 17 2.3.1- Cold Rolling Texture………………………………………………………… 27

2.3.2- Hot and Warm Rolling Texture…………………………………………….. 28

2.4 – X-ray Peak Profile Analysis……………………………………………………… 28

vi

2.4.1 - X-ray Peak Profile Analysis from MWP and Methodology for

Determining the Burgers Vector Populations…………………………………………… 34

2.5 – Self-Consistent Modeling of Deformation Texture…….…………………….…. 38 2.5.1 – The Single-Crystal Constitutive Law……………………………………… 40 2.5.2 – Polycrystal Constitutive Law………………………………………………. 43 2.5.3 – The Self-Consistent Approach……………………………………………... 47 3. EXPERIMENTAL PROCEDURE…………………………………………………… 50 3.1 - Material …………………………………………………………..……………….. 50 3.2 - Thermo-Mechanical Processing…………………………………………..…….... 51 3.2.1- Cold Rolling……………………………………………………………….….. 54 3.2.2 - Hot Rolling…………………...…………………………………………........ 54 3.3 - Metallographic Sample Preparation……………………………………………... 55 3.3.1 - Mechanical Polishing………………………………………………………... 56 3.4 - Characterization Techniques……………………………………………………... 56 3.4.1- Texture Measurement……………………………………………………..… 57 3.4.2 - Peak Profile Measurements………………………………………………… 58 4. RESULTS…………………………………………………………………………......... 60 4.1 - Texture Evolution…….…………………………………………………………… 60 4.1.1 - As Received Sample …………………………………………….................... 60 4.1.2 – Cold Rolled Sample......................................................................................... 61 4.1.3 - Warm Rolled Samples..................................................................................... 67 4.2 - X-ray Peak Profile Analysis………………………………………………………. 72 4.3- Texture Simulation………………………………………………………………… 78 5. DISCUSSION…………………………………………………………………………... 83

vii

5.1 - Deformation Texture…………………………………………………………….... 83 5.2 - X-Ray Peak Profile Analysis……………………………………………..…......... 86 5.3 - Self Consistent Simulation of the Deformation Texture………………...……… 87 6 - SUMMARY AND FUTURE WORK…………………………………………………. 90 6.1 – Summary………………………………………………………………………...… 90 6.2 – Future Work………………………………………………………………………. 91 REFFERENCES…………………………………………………………………………… 93 BIBLIOGRAFICAL SKETCH………………………………………………………….... 100

viii

LIST OF TABLES Table 2.1 – Some properties of titanium and it’s alloys…………………………..………... 5 Table 2.2 – Summary of commercial and semi-commercial grades and alloys of titanium... 10 Table 2.3- Number of grains showing a specific glide system for different samples………. 14 Table 2.4 - The most important deformation systems in hcp metals and their influence on the texture evolution ………………………………………………………………………..

15

Table 2.5 – The most typical correlations between diffraction peak aberrations and the different elements of microstructure ………………………………………………………..

30

Table 2.6 - The most common slip systems in hexagonal crystals: (a) Edge dislocations and (b) Screw dislocations ……………..

33

Table 3.1 - Chemical composition (weight %) …………………………………………….. 50 Table 3.2 - Typical mechanical properties of the CP Ti Gr2……………………………….. 50 Table 3.3 Physical properties of the CP Ti Gr2 ……………………………………………. 51 Table 3.4 - Nomenclature of the samples………………………..………………………….. 54 Table 3.5 - Metallographic preparation procedure …………………………………………. 56 Table 4.1 - Dislocation densities and arrangement parameter, M, obtained from MWP evaluation for Ti samples deformed at different reduction levels …………………………..

73

ix

LIST OF FIGURES

Figure 2 1 - Commercial production of Titanium ……………………................................... 6 Figure 2.2 – The hexagonal unit cell (a) and the first order slip and twinning planes for hcp metals (b)………………………………………….................................................................

12

Figure 2.3 – Glide systems in alpha titanium ……………………………...……………….. 13 Figure 2.4 - Schematics of all investigations carried out and definition of sample short names. The starting texture of the different materials is given in the form of (0001) and /1010/ X-ray pole figures. Sample short names are composed as follows: (1) chemical composition; (2) sheet thickness in mm; (3) deformation mode; (4) angle between RD and tension direction (0°, 45°, 90°) or deformation degree (2%,4%)……………………………

16

Figure 2.5 – Sheet textures in hcp materials as a function of c/a ratios (schematically)……. 18 Figure 2.6 – Ideal cold rolling texture component for flat-cold rolled titanium: 2115 <1010>………………………………………………………………………………………..

19

Figure 2.7 – Typical textures………………………………………………………………… 20 Figure 2.8 - Positioning and movement of the sample on the texture goniometer inside the X-ray machine (a). The relation between crystallite coordinates (Xc, Yc, Zc) and sample coordinates (Xs, Ys, Zs), (b), (c) and (d)…………………………………………………….

21

Figure 2.9 – As received material: a) Pole figures and b) Inverse pole figures……………... 22 Figure 2.10 – Pole figure representation of the cold rolling and the recrystalization texture components…………………………………………………………………………………...

23

Figure 2.11 - Three consecutives Euler rotations defining an orientation ………………….. 24 Figure 2.12 – Relationship between sample and crystal axis directions…………………….. 25 Figure 2.13 – Constant φ sections through the Eulerian space: a) 0°, b) 20°, c) 30°, d) 40 and e) 60°……………………………………………………………………………………..

26

Figure 2.14 – Location of the cold rolling and recrystalization components on the constant

x

phi sections of the Euler space using Roe’s definition………………………………………. 27 Figure 2.15 – The parabolas describing the average contrast factors for the eleven slip systems, in the case of Titanium, as a function of x = (2/3)(l/ga)2 …………………………..

35

Figure 2.16 - Slip systems in hexagonal crystal systems ……………..…………………….. 36 Figure 3.1 – As received material: OIM/SEM micrograph…………………………………. 52 Figure 3.2 – Rolling mill machine.………………………………………………………….. 53 Figure 3.3 – Schematic setup of the thermo-mechanical processing.……………………….. 53 Figure 3.4 - X-ray machine Philips X’Pert MRD equipped with texture goniometer………. 57 Figure 3.5 - Surfaces examined by X-ray diffraction: normal direction (ND); rolling direction (RD); transverse direction (TD)……………………………………………………

58

Figure 3.6 - Example of the instrumental broadening of the Alpha-1 Panalytical Diffractometer measured using LaB6 660a NIST standard compared with the peak broadening measured for deformed α-Ti. The dashed line is the 220 reflection of LaB6 and the continuous line is the 11.0 reflection of α-Ti deformed at the 60% reduction rate……...

59

Figure 4.1 - (0002) and (1010) pole figure for the as received sample……………………... 61 Figure 4.2 - ODF sections of φ =0° and φ =30°, Roe notation, for the as received sample…. 61 Figure 4.3 – (0002) and (1010) pole figures of the cold rolled samples…………………….. 63 Figure 4.4- ODF sections of φ =0° and φ = 30°, Roe notation, for the samples cold rolled at: a) 20%, b) 40%, c) 60%, d) 80% and e) 95%..................................................................…….

64

Figure 4.5 - Skeleton lines of the orientation distribution functions for the samples 20%, 40%, 60%, 80% and 95% cold rolled………………………………………………………..

65

Figure 4.6 - Development of the 0002//ND fiber texture for the as received (AR) and 20%, 40%, 60%, 80% and 95% cold rolled (CR) samples…………………………………..

66

Figure 4.7 - Variation in volume fraction of the 0002//ND fiber texture with degree of cold rolling reduction. The as received material corresponds to the 0% cold rolling reduction..................................................................................................................................

67

Figure 4.8 – (0002) and (1010) pole figures of the warm rolled samples…………………... 68

Figure 4.9 - ODF sections of φ =0° and φ = 30°, Roe notation, for the samples warm rolled at: a) 20%, b) 40%, c) 60%, d) 80% and e) 95%…………………………………………….

69

xi

Figure 4.10 - Skeleton lines of the orientation distribution functions for the samples 20%, 40%, 60%, 80% and 95% warm rolled ……………………………………………………...

70

Figure 4.11 - Development of the 0002//ND fiber texture for the as received (AR) and 20%, 40%, 60%, 80% and 95% warm rolled (WR) samples.………………………………..

71

Figure 4.12 - Variation in volume fraction of the (0002)//ND fiber texture with degree of warm rolling reduction. The as received material corresponds to the 0% cold rolling reduction……………………………………………………………………………………..

72

Figure 4.13 - The hi fractions of the three fundamental Burgers vector types, <a>, <c> and <c+a>, as a function of rolling reduction. Note that in the figure the solutions to equations. (2.8, 2.9 and 2.10), the hi fractions, were transformed in percentages…………...

74

Figure 4.14 - The line profiles of (0002) Bragg reflections for different deformations levels. On the x-axes K is given by K=2sinθ/λ, where θ is the Bragg angle and λ is the wave length of the used radiation……………………………………………………………

75

Figure 4.15 - The line profiles of (1120) Bragg reflections for different deformations levels. On the x-axes K is given by K=2sinθ/λ, where θ is the Bragg angle and λ is the wavelength of the used radiation………………………………………………………....…..

76

Figure 4.16 - (2110), (0001) and (2113) pole figures of alpha titanium at a rolling reduction of (a) 0%, (b) 40% (c) 60%, (d) 80%, respectively…………………………………………..

77

Figure 4.17 - Evolution of intensities of components with RD//2110, RD//0001 and RD//2113, respectively, during rolling reduction.

78

Figure 4.18 – (0002) pole figures for the as-received material: (a) experimental and (b) discrete grains file. Axes convention: RD in the vertical direction and TD in the horizontal direction……………………………………………………………………………………...

79

Figure 4.19 – Experimental and simulated results of the (0002) pole figures for the cold rolled samples deformed (a) 20%, (b) 40%, (c) 60%, (d) 80% and (e) 95%...........................

81

Figure 4.20 – Experimental and simulated results of the (0002) pole figures for the warm rolled samples deformed (a) 20%, (b) 40%, (c) 60%, (d) 80% and (e) 95%...........................

82

Figure 5. 1 - Variation of: (a) twin volume fraction; (b) strain accommodated by twinning as a function of rolling temperature………………………………………………………….

88

Figure 5.2 – Optical micrographs of warm rolled 80% and 95% reduction………………… 89

xii

ABSTRACT

The present work attempts to establish a unified path model for characterization as well

as prediction of microstructure evolution, in terms of texture, in commercially pure titanium that

have undergone thermo-mechanical processing. Two deformation temperatures, room

temperature (cold rolling) and 260°C (warm rolling), and five different degrees of deformation,

20%, 40%, 60%, 80% and 95% were used in this investigation. X-ray measurements (texture

measurements and peak profile analysis) have been used to characterize the texture and to

evaluate the relative activity of the various slips systems activated during the process.

Simulations of the resultant textures after each mode of deformation were performed using a

crystal plasticity self-consistent scheme, and comparisons, in the form of pole figures, between

the experimental results and the predicted deformation textures were performed in order to

validate the results obtained from peak profile analysis.

The experimental texture results show that except for the samples 95% deformed, the

warm rolling has shown to develop a deformed texture different from the cold rolling.

The results of peak profile analysis carried out for the 40%, 60% and 80% warm rolled

samples show that the <a> type of dislocation was prevalent in all samples while the <c> type of

dislocation was only marginal. The X-ray peak profile analysis, based on the dislocation model

of anisotropic peak broadening, show the dislocation densities, distributions and type during the

rolling process in good agreement with the texture evolution.

Even though twining was not taken into account during simulation of the cold rolled

samples, there was a reasonable agreement between the experimental and the predicted pole

figures with a small divergence on the distribution of in the TD-RD plane for the higher

deformed samples.

The results of simulated deformation texture of warm rolled CP-Ti are in good agreement

with the experimental results and with the peak profile analysis findings.

xiii

CHAPTER 1

INTRODUTION

Despite being discovered as early as 1790, it was not until late 1940’s that interest in

titanium and its alloys, as structural materials, began to accelerate, as their potential as high-

temperature, high-strength/density ratio and corrosion resistant materials with aeronautical

applications became apparent [Boyer et al., 1994; Froes, 1990] and in a relatively short time,

titanium has come to be used for many different and important purposes. Its greatest

disadvantage is the high cost compared to competing materials which frequently offset’s

titanium’s engineering advantages and restrings the market for titanium applications. Aiming to

change this perspective, just as other metals, such as aluminum, have had cost breakthroughs that

have dramatically expanded their use, a great deal of money and time has been put in basic

research to lower production costs improving both extraction and processing technologies.

Titanium and other metals with hexagonal crystal structure develop sharp deformation

textures that lead to a pronounced plastic anisotropy of the polycrystalline sample [Phillipe,

1995; Zaefferer, 2003]. Various factors can cause anisotropy in metals, among them are: grain

morphology [Kocks and Chandra, 1982], second phase precipitates [Mizera et al., 1996; Crosby

et al., 2000] and substitutional alloying elements [Phillipe, 1988]. As a consequence, the

deformation texture may vary with slight changes of the material composition [Zaefferer, 2003].

Researchers [Crosby et al., 2000; Fjeldly and Roven, 1996] agree that crystallographic textures

resulting from thermomecanical processing such as hot or cold rolling are most directly

responsible for anisotropy in metal alloys. Anisotropy of mechanical properties is a concern in

the forming of metals into shapes and parts; and the control of texture throughout the process can

provide beneficial use of the variety of available textures in α, near-α and other titanium alloys

[Zhu, 1997].

1

In this scenario it becomes evident that an understanding on how the thermo-mechanical

processing affects the final properties of a semi-finished or finished material is of major

significance. Moreover, considering that the cost associated with the testing and development of

a product is somehow enormous and time consuming, availability of experimental

characterization techniques and computational tools capable of providing reliable data leading to

the prediction of “optimal” processing paths linking the commercially available “raw material”

to its semi-finished or finished forms, is of strategic importance.

In order to model deformation processes, it is fundamental the knowledge on the

evolution of parameters such as dislocation density and the relative activity of the various slips

systems activated during the process. The measurement of such parameters is normally executed

employing established techniques as transmission electron microscopy (TEM), electron back

scattering (EBSD) and trace analysis. However, the measurement of these parameters in

specimens that have undergone large strains and the consequent large number of dislocations

introduced (which are the actual characteristics of materials in industrial practices) is difficult

and time demanding with a considerable cost associated.

On the other hand, the characterization of material defects using X-ray and neutron-

diffraction techniques has received considerable attention during the past few decades and line

broadening analysis can be an attractive alternative, in substitution or associated with TEM, in

the study and evaluation of the substructure developed during thermo-mechanical processes. The

most attractive feature of these techniques is that they can be used to measure materials heavily

deformed. Other advantages are the easy sample preparation routine, relative short time required

by the “state of the art” equipments to render results and the fact that the data obtained is

statistically averaged over the area/volume irradiated.

The goal defined for this work is to establish a unified path model for characterization

and also prediction of microstructure evolution, in terms of texture, in materials that have

undergone thermo-mechanical processing. Once the behavior of a material is ‘mapped’ by means

of a reliable characterization scheme, other possible ‘routs’ of processing can be simulated

before any actual processing/testing is conducted in order to verify the accuracy of a chosen path

for processing.

In order to achieve this objective, it was decided to apply the following methodology to a

polycrystalline HCP material, which have undergone unidirectional rolling at different degrees of

2

deformation (20% ~ 95% reduction in thickness or equivalent true strain of -0.22 ~ -3.0) and at

different isothermal conditions (25°C and 260°C). The choice of temperatures aimed to isolate

different mechanisms of deformation. The first step consisted in the characterization of

microstructure evolution (texture) by means of X-ray pole figure measurements and ODF

analysis. The second task was to employ X-ray peak profile analysis to study the microstructure

evolution evaluating the densities, distribution and type of dislocations. The third step consisted

in validating the results obtained from Peak Profile analysis by simulation of deformation texture

evolution using a crystal plasticity self-consistent scheme, and comparison of experimental

results with the predicted deformation texture in the form of pole figures.

3

CHAPTER 2

BACKGROUND 2.1 – Titanium and its Alloys

Titanium is the fourth most abundant structural metal and the ninth most abundant

element, making up about 0.6% of the Earth's crust. It occurs in many mineral forms, but only

three present significant economical interest: leucoxene, rutile and ilmenite. Titanium was first

discovered (as rutile) by W. Gregor in 1791 and by M. H. Klaproth in 1795 [Boyer, Welsch and

Collings, 1994] who named the metal after the Greek mythological god Titan [Bomberger, Froes

and Morton, 1985; Ogdom and Gonser, 1956; McQuillan and McQuillan, 1956] and the first cost

effective non-vacuum process for titanium extraction from its ore was developed by W.J. Kroll

[Hunter, 1910]. Interest in titanium and its alloys, as structural materials, began to accelerate in

the late 1940s and early 1950s, as their potential as high-temperature, high-strength/density ratio

and corrosion resistant materials with aeronautical applications became apparent [Froes, 1990;

Boyer, Welsch and Collings, 1994]. Due to its unique set of properties (see table 2.1), nowadays,

titanium and its alloys have been widely used throughout the aerospace industry for most types

of structural components, including airframes and engine components, as well as in many non-

aerospace applications. Just to mention a few, as a metal, cars, sports equipment such as racing

yacht parts, golf clubs, tennis racquets and bike frames, wrist watches, underwater craft, and

general industrial equipment. Its non-toxicity also makes it useful for surgical implants such as

pacemakers, artificial joints and bone pins. Titanium is also used to manufacture chlorine. As

Titanium dioxide, it is used in paints (replacing the use of lead), lacquers, paper, plastics, ink,

rubber, textiles, cosmetics, sunscreens, leather, food coloring, and ceramics. It is also used as a

coating on welding rods. Titanium dioxide is one of the whitest and brightest substances known.

Due to its reflective properties, Titanium dioxide, add richness/brightness to colors and provides

4

UV protection. Finally, as a compound known as titanium tetrachloride, it is used for

smokescreens and skywriting.

Titanium usage is, however, strongly limited by its higher extraction and production cost

relative to competing materials such as aluminum, grades of stainless steel and other steels just

to mention.

Table 2.1 – Some properties of titanium and it’s alloys

High strength-to-weight ratio

Corrosion-resistant

High melting point (1660°C).

Non-toxic

Titanium dioxide is one of the whitest and brightest substances

Provides protection from UV rays

The basic commercial production process of titanium (figure 2.1) starts with the

extraction, which involves treatment of the ore (leucoxene, rutile or ilmenite) with chlorine gas

to produce titanium tetra-chloride, which is purified and reduced to what is known as titanium

sponge. The sponge is then blended with alloying elements and vacuum melted giving origin to

an ingot.

After obtaining a homogeneous ingot, it is processed into suitable shapes and sizes,

typically, by forging followed by rolling. Forging and rolling are not the only forming processes

employed to transform the material from its “raw” form into the final desired product. There are

many other processing routes for the thermo-mechanical processing of titanium products

described elsewhere. The present investigation is focused in the texture and microstructure

evolution resultant from rolling, therefore this discussion will be limited to cold and warm

rolling.

5

Figure 2 1- Commercial production of Titanium

2.1.1 - Physical Metallurgy of Titanium and Titanium Alloys

Titanium is an allotropic element, which means that it exists in more than one

crystallographic form. At room temperature, titanium exhibits a hexagonal close-packed (hcp)

crystal structure, referred to as "alpha" phase. However, as the temperature is raised through

882.5C° (1621 °F) [Collings, 1994], pure titanium undergoes an allotropic transformation from

the α-phase (hcp) to a body-centered cubic (bcc) crystal structure, called "beta" (β) phase. The

temperature at which this transformation occurs is known as beta transus and it is defined as the

lowest equilibrium temperature at which the material is 100% β (bcc) phase. Alloying element

addition to pure titanium can either cause the transformation temperature to increase decrease or

remain unaffected. These elements are generally classified as alpha or beta stabilizers. The group

of alloying elements that favor the α-phase and stabilize it by raising the beta-transus

temperature include aluminum, gallium, germanium, carbon, oxygen and nitrogen. The β phase

is stabilized by the addition of elements, which promote the lowering of the beta-transus

temperature. Such elements are classified in two groups: the β isomorphous and the eutectoid.

The former consists of those elements that are miscible in the β phase: molybdenum, vanadium,

tantalum, and niobium. The second is formed by those whose form eutectoid systems with

titanium, having eutectoid temperatures as low as 550°C, i.e., as much as 333°C (600 °F) below

the beta-transus for pure titanium. This group includes manganese, iron, chromium, cobalt,

6

nickel, copper, and silicon. Besides the cited elements, Zr and Tin, due to their extensive solid

solubility, are also employed as strengthening agents and also to retard the rates of phase

transformation.

2.1.2 - Classification of Titanium Alloys

Titanium alloys are classified, basically, taking in account their chemical composition,

the weight % of the alloying elements, and its effect on the resultant microstructure at room

temperature. As an example, the reason why pure titanium is classified as α-titanium is due to

the fact that at room temperature its microstructure is formed entirely by grains with hexagonal

close-packed (hcp) crystal structure. An example of dual phase alloy is Ti-6Al-4V, which

contains both alpha and beta stabilizers, and as a consequence “alpha + beta” alloys exhibit a

certain volumetric fraction of beta phase stabilized at room temperature. Table 2.2 presents a

number of commercially available alloys arranged accordingly to their classification in alpha,

near-alpha, “alpha + beta” and beta alloys [Reed-Hill, 1992].

2.1.2.1 - Alpha-Titanium Alloy

This group consists of both pure titanium (or unalloyed) and those alloys containing α-

stabilizing elements such as Al, Ga and Sn, either singly or in combination. The commonly used

alloys are the several grades of commercially pure (CP) titanium, which are in effect Ti-O alloys,

and the ternary composition Ti-5AI-2.5Sn. As mentioned previously at ordinary temperatures

these are HCP materials [Collings, 1994]. As alpha alloys are single-phase materials, tensile

strengths are relatively low especially for low oxygen grades, although their high thermal

stability leads to reasonable creep strengths. These alloys are also characterized by good ductility

down to very low temperatures, reasonable strength, toughness and good weldability [Wood,

1972]. However, due to the fact that the alloys in this group are single phased with hexagonal

crystal structure they also exhibit a high rate of strain hardening, being the high content of

oxygen associated with its limited formability [Polmear, 1995].

7

2.1.2.2 - Near-Alpha Titanium Alloys

Developed to meet demands for higher operating temperatures, this class of alloys,

possess higher room-temperature tensile strength than that exhibit by alpha alloys. They also

show the greatest creep resistance of all titanium alloys at temperatures above 400°C. Usually,

near alpha-alloys are forged and heat treated in the “alpha + beta” field so that primary beta-

grains are always present in the microstructure.

Improved creep performance has been achieved in special compositions by carrying out

these operations at higher temperatures in the upper “alpha + beta” and beta fields resulting in a

change to a more elongated alpha microstructure. The two alloys, which currently show the

highest creep resistance with a maximum operating temperature about 600°C (1112°F) are the

Timetal 1100, and Timetal 834. Timetal 1100 is processed by forging just below the β-transus

and the resultant microstructure exhibits a mixture of equiaxed and elongated alpha grains, which

provides a balance of good creep and low cycle fatigue resistance [Polmear, 1995].

2.1.2.3 – Alpha/Beta (α + β ) Alloys

These alloys have both α and β phases in equilibrium at room temperature. They combine

the strength of α alloys with the ductility of β alloys, and their microstructure and properties can

be varied widely by appropriate heat-treatments and/or thermo-mechanical processing. The most

known and used (α + β) alloy is the Ti-6Al-4V or Ti-6-4. Other commercially available alloys in

this class are the Ti-6-6-2 and Ti-6-2-4-6 whose can exhibit, in certain cases, higher strengths

than the high temperature near-alpha alloys. Other characteristics of these alloys are the good

weldability, which is a function of β-stabilizing contents, good combination of properties having

a wide processing window meaning less stringent processing requirements than those required

for other alloys types and their capability for applications up to 400°C. They also can be

strengthened with a solution treatment to establish the hardenability followed by aging. The

amount of strengthening that can be achieved is a function of section thickness and chemical

composition of the alloy: as the β-stabilizing content increases the hardenability increases.

8

2.1.2.4 - Beta, Near-Beta and Metastable-Beta alloys

There is no clear-cut definition for beta titanium alloys. Conventional terminology

usually refers to near-beta alloys and metastable-beta alloys as classes of beta titanium alloys. A

near-beta alloy is generally one that has appreciably higher beta stabilizer content than a

conventional alpha-beta alloy such as Ti-6Al-4V, but is not quite sufficiently stabilized to readily

retain an all-beta structure with an air cool of thin sections. For such alloys, a water quench even

of thin sections is required. Due to the marginal stability of the beta phase in these alloys, they

are primarily solution treated below the β-transus to produce primary alpha phase which in turn

results in an enriched, more stable beta phase. The Ti-10V-2Fe-3Al alloy is an example of a

near-beta alloy. On the other hand, the metastable-beta alloys are even more heavily alloyed with

beta stabilizers than near-beta alloys and, as such, readily retain an all-beta structure upon air-

cooling of thin sections. Due to the added stability of these alloys, it is not necessary to heat treat

below the β-transus to enrich the beta phase. Therefore, these alloys do not normally contain

primary alpha since they are usually solution treated above the β-transus. These alloys are termed

“metastable” because the resultant beta phase is not truly stable, it can be aged to precipitate

alpha for strengthening purposes. Alloys such as Ti-15-3, B120VCA, Beta C, and Beta III are

considered metastable-beta alloys.

Unfortunately, the classification of an alloy as either near-beta or metastable beta is not

always obvious. In fact, the “metastable” terminology is not precise since a near-beta alloy is

also metastable, i.e., it also decomposes to alpha plus beta upon aging. There is one obvious

additional category of beta alloys: the stable beta alloys. These alloys are so heavily alloyed with

beta stabilizers that the beta phase will not decompose to alpha plus beta upon subsequent aging.

There are no such alloys currently being produced commercially. An example of such an alloy is

Ti-30Mo.

The interest in beta alloys stems from the fact that they contain a high volume fraction of

beta phase, which can be subsequently hardened by alpha precipitation. Thus, these alloys can

generate quite high strength levels (in excess of 200 ksi) with good ductility. Also, such alloys

are much more deep hardenable than alpha-beta alloys such as Ti-6Al-4V. Finally, many of the

more heavily alloyed beta alloys exhibit excellent cold formability and as such offer attractive

sheet metal forming characteristics.

9

Table 2.2 – Summary of commercial and semi-commercial grades and alloys of titanium [Reed-Hill, 1992].

10

2.2 - Mechanical Behavior of Titanium and its Alloys

In hexagonal close-paked (hcp) metals the low number of easy slip systems, their

asymmetrical distribution, and the strict crystallographic orientation relationships for twinning

results in the formation of a strong deformation texture. The deformation mechanism together

with the texture is responsible for the strong anisotropy of the mechanical properties [Hosford

and Backofen, 1964]

In hcp alpha-titanium, slip occurs most commonly on the basal 0001, prismatic 1010,

and pyramidal 1011 slip planes (figure 2.2). The actual dominant slip planes depend on the c/a

ratio, as well as alloy composition, temperature, grain size, and crystal orientation. In general,

slip will tend to occur on the plane having the largest inter-planar distance. For hexagonal

materials exhibiting c/a ratio less than 1.663 (considered the ideal ratio), the prismatic plane is on

average the most densely packed plane. For alpha-titanium (c/a = 1.587), the prismatic plane is

the most densely packed. As a consequence, the smallest resolved shear stress occurs at the

prismatic slip plane. This is the case of high purity alpha-titanium. If high interstitial levels of

oxygen and/or nitrogen are present, as it is the case in low purity alloys (i.e., CP titanium), all

three slip planes are activated, but the prismatic plane is still the one with lower resolved shear

stress required to initiate slip. Hexagonal materials, due to its 6-fold rotation symmetry do not

exhibit a complete set of slip systems. As a consequence of this limited number of slip system

capable of being activated, further deformation is accommodated either by <c+a> pyramidal

glide or traction/compression twinning.

Twinning results when a portion of the crystal takes up an orientation that is related to the

orientation of the rest of the untwined lattice in a definite symmetrical way. The plane of

symmetry between the two portions is known as twinning plane. In titanium, the most common

twinning plane is (1012) and twinning direction is [1011] [Dieter, 1986].

2.2.1 – Slip Modes in HCP Metals

The primary slip systems operative in HCP metals with c/a ratio less than the ideal

1.633are the prismatic 1010 planes in the basal <1210> directions. The other first order

possible slip systems are the basal (0001) and pyramidal 1011 planes with basal directions

11

<1210>. These systems will provide combinations of 4 independent slip systems, since they all

occurs on the basal direction.

Figure 2.2 – The hexagonal unit cell (a) and the first order slip and twinning planes for hcp metals (b) [Dieter, 1986].

Differently from materials with cubic crystalline structure that posses 5 or more glide

systems, in hexagonal close packed metals the most common basal and prismatic glide modes

have only 2 or 3 independent glide systems respectively [Groves, 1963]. As a consequence, since

at least four or five independent slip systems are necessary to accommodate arbitrary plastic

strains, secondary systems like pyramidal glide with <c+a> Burger’s vector, or twinning systems

can be activated contributing to accommodate the imposed strain [Yoo, 1981 and Partridge,

1967]. Figure 2.3 shows the primary and secondary glide systems for Titanium.

Regarding hexagonal metals, the activation of slip and twinning systems is normally

affected by parameters like c/a ratio, interstitial constituent (i.e. Oxygen content principally in

the case of CP Ti), strain hardening, strain rate and temperature.

At room temperature, as a consequence of cold rolling, Ti deforms by prismatic glide

1010<1210>, pyramidal glide 1011<1120> with <a> Burger’s vector and secondary

1011<1123> with <c+a> Burger’s vector, 1012<1011> (and in some cases 1121 twinning

12

in tension and 1122<1123> twinning in compression [Rosi et al. 1956; Conrad, 1981; Chin,

1975].

Figure 2.3 – Glide systems in alpha titanium.

13

In the case of high purity titanium deformed in uniaxial compression at 20°C, it was reported

(via EBSD analysis), the activation of three types of twins: 1122<1123>, 1012<1011> and

1121<1126>, in the proportions of 40%-30%-30% respectively [Salem 2002, Kalidindi et al.

2004].

Zaefferer investigated the relation between the formation of cold rolling textures and the

activated glide and twinning systems during deformation of polycrystalline Titanium [Zaefferer,

2003]. Samples of three different titanium alloys (Ti-6Al-4V and two commercially pure

Titanium grades designated in the work as T40 (1000ppm O) and T60 (2000ppm O)) were

deformed up to 5% by uniaxial or biaxial. Zaefferer observed a considerable activity of <c+a>

and twinning in the case of the T40 alloy with a pronounced TD-type texture and for the T60

Alloy, the higher oxygen content completely suppressed twinning and strongly reduced <c+a>

activity resulting in a less developed TD-type texture which was a result of a combination of

<c+a> and basal slip. The results reported by Zaefferer are summarized below in the tables 2.3

and 2.4 and the main textures observed are presented in the figure 2.4.

Table 2.3- Number of grains showing a specific glide system for different samples Slip system TA3Z0 (%) TA3Z45 (%) TA3Z90 (%) TA1Z (%) TA1B (%) T401B (%) T601B (%)

<a>-Basal 1 (6) 3 (28) 4 (16) 2 (5) 11 (35) 5 (14) 9 (37)

<a>-Prismatic 4 (27) -- 1 (4) 9 (23) 3 (10) 2 (6) 1 (4)

<a>-Pyramidal 3 (20) 2 (18) 3 (12) 3 (7) 3 (10) 8 (21) 3 (13)

<a>-Screw 3 (20) 4 (36) 5 (20) 26 (65) 9 (30) 6 (16) 4 (16)

<c+a>-Pyramidal 4 (27) 2 (18) 12 (48) -- 5 (15) 13 (34) 5 (21)

Others T401B - <c+a>-Prismatic 4 (9) and T601B- <c+a>-Prismatic 2 (8)

Phillipe and Fundenberger [Phillipe et all, 1995, Fundenberger et all, 1997], working with

cp-Titanium grade 1(T35) and grade 2(T60) respectively, studied the activation of glide and twin

systems during cold rolling and observed the occurrence of 1010 <1120> prismatic slip and a

very low activity of Basal and pyramidal <a> slip. They also observed activation of two twinning

14

systems: 1012 tension twins and 1122 compression twins. In the case of second order

pyramidal slip <c+a> it was observed a low activity of this type of gliding up to 50% of

deformation but from this point up to 80% reduction in thickness, twinning is suppressed and to

accommodate further deformation in <c> direction, the <c+a> pyramidal gliding was activated

instead of 1122 compression twinning.

Table 2.4 - The most important deformation systems in hcp metals and their influence on the texture evolution [Zaefferer, 2003]. Burgers vector or

shear direction

Glide or

shear plane*

Name

Related cold-rolling

texture type

1/3<1120> (<a>) 0001

1010

1011

Basal glide

Prismatic glide

<a> Pyramidal glide

r-type [Sakai and

fine,1974],c-type

[Conrad, 1981]

r-type [Philippe et al.

1988]

1/3<1123> (<c+a>) 1011

1122

<c+a> Pyramidal (I) glide

<c+a> Pyramidal (II) glide

t-type

t-type

<1011> 1012 1012 Twin Under tension – c-type

(Ti); compression – r-

type (Zn)

<1126> 1121 1121 Twin Under tension – c-type

<1123> 1122 1122 Twin Under compression –

t-type

* Glide plane in case of dislocations, shear plane in case of twinning.

15

Figure 2.4 - Schematics of all investigations carried out and definition of sample short names. The starting texture of the different materials is given in the form of (0001) and /1010/ X-ray pole figures. Sample short names are composed as follows: (1) chemical composition; (2) sheet thickness in mm; (3) deformation mode; (4) angle between RD and tension direction (0°, 45°, 90°) or deformation degree (2%, 4%) [Zaefferer, 2003].

In titanium, Rosi et al. observed no twins of any type at 800°C, while McHarque and

Hamond reported a small amount of 1122 and 1121 twinning at 815°C. At room

temperature and below that titanium slips along the <1120> direction on the 1010, (0001) and

1011 planes. Changes in length along the c axis are not possible with <1120> slip alone,

requiring a slip direction lying out of the basal plane (0001). Such slip has been reported in

16

commercially pure titanium as a result from the motion of the <c+a> dislocations along the

<1123>. A length change along the c axis can also be accomplished by twinning. In titanium,

1012, 1121 and 1123 twins allow an extension along the c axis, while 1122, 1124

and 1010 twins allow a reduction in the c axis; whish generally becomes less important as the

deformation temperature increases. Paton and Backofen [Paton and Backofen, 1970]

investigating iodide titanium single crystals under compression at temperatures from 25°C to

800°C, have found that reduction of up to a few percent strain along the c axis is accommodated

almost entirely by 1122 twinning from 25°C to 300°C. According to their results, although

<c+a> slip is not responsible for a significant amount of strain below 300°C, it is important as a

means of accommodating the shear ahead of a propagating 1122 twin.

2.3 - Texture

The most commonly and important used materials for industrial practice, such as metals,

ceramics and some polymers are polycrystalline materials and their component units are referred

to as crystals or grains. Grain orientations in polycrystals are rarely random due to the processing

history that the polycrystalline materials are normally submitted to, such as solidification from

melting, hot rolling, cold rolling and annealing among other thermo-mechanical processes.

Therefore, in most materials there is a pattern in the orientations, which are present and a

tendency for the occurrence of certain orientations. This tendency is known as preferred

orientation of crystals or texture. The relevance of texture to materials lies in the fact that many

materials properties are texture-dependent. According to Bunge-1987, the influence of texture on

material’s properties is, in many cases, 20-50% of the properties values. Some examples of

properties which depend on the average texture of a material are: Young’s Modulus, Poisson’s

ratio, strength, ductility, toughness, magnetic permeability, electrical conductivity and thermal

expansion (in non-cubic materials) [Randle and Engler, 2000].

Texture, in hexagonal materials, is represented by the Miller indices hkil<uvtw> where

hkil corresponds to the family of crystallographic planes parallel to the surface of the sample

and <uvtw> corresponds to the family of crystallographic directions parallel to the rolling

17

direction (RD) of the sample. The resulting rolling textures, in the form of pole figures, as a

function of c/a ratio is shown in figure 2.5.

Figure 2.5 – Sheet textures in hcp materials as a function of c/a ratios (schematically).

18

The ideal cold rolling texture component is represented in figure 2.6 and other typical

textures in hexagonal materials are shown in figure 2.7.

Figure 2.6 – Ideal cold rolling texture component for flat-cold rolled titanium: 2115 <1010>.

Texture can be determined by means of X-ray diffraction, neutron diffraction and

electron diffraction using Transmission Electron Microscope (TEM) or Scanning Electron

Microscope (SEM). X-ray diffraction is the most commonly applied technique but the neutron

and electron diffraction techniques are gaining interest because it permits one to correlate

microstructures, neighbor relations and texture [Kocks, 1998].

Among the ways to describe texture, pole figure (PF), inverse pole figure (IPF) and

orientation distribution function (ODF) are the most usual methods. Pole figure is a projection

[Cullity; 2001], more often represented as a stereographic projection, which shows the variation

of pole density with pole orientation for a selected set of crystal planes having the rolling

direction (RD), the transversal direction (TD) and the normal direction of the sample as reference

axis.

19

Figure 2.7 – Typical textures [Wang, 2003].

20

Pole figures are measured using x-ray diffraction and in order to have a specific (hkil)

reflection, the following condition, known as Bragg’s law (equation (2.1)), must be satisfied.

nλ = 2 dhkl . sin θ (2.1)

During the pole figure measurement, to determine a pole density, the x-ray detector

remains stationary at the proper 2θ angle, to receive the (hkil) reflections, while the specimen

rotates in two particular ways. These rotations permit a complete scanning of the specimen’s

surface and the positioning of the sample on the texture goniometer is shown in figure 2.8.

Figure 2.8 - Positioning and movement of the sample on the texture goniometer inside the X-ray machine (a). The relation between crystallite coordinates (Xc, Yc, Zc) and sample coordinates (Xs, Ys, Zs), (b), (c) and (d).

21

The α and β angles, which are respectively the polar and the azimuthal angles, define the

movements of the sample during the pole figure measurement.

The inverse pole figure (IPF) is a pole density projection of the (hkil) planes referred to

the stereographic triangle. Inverse pole figure presents an advantage over the pole figure because

an IPF shows the density distribution of all planes within the stereographic triangle instead of

showing only the density of a specific crystallographic plane (see figure 2.9).

RD

(0002) (1010) (2110)

TD

a)

ND TD RD (1010)

(0002) (2110)

b)

Figure 2.9 – As received material: a) Pole figures and b) Inverse pole figures.

The pole figure and the inverse pole figure are very helpful tools however principal

orientations of the texture cannot be precisely determined from them because they do not provide

information regarding the crystallographic directions in the plane of the sample. The figure 2.10

22

exemplifies a situation where two different texture components, the cold rolling and the

recrystalization components, exhibit the same (0002) pole figure, which can be misleading if the

analysis is based only on basal pole figures.

Figure 2.10 – Pole figure representation of the cold rolling and the recrystalization texture components.

It has been well established that the orientation distribution in textured materials can be

qualitatively as well as quantitatively evaluated by the crystallite orientation distribution function

analysis (ODF) developed by Bunge and by Roe [Bunge, 1982; Roe, 1965]. The ODF describes

the frequency of occurrence of particular orientations in a three-dimensional orientation space.

This space is defined by three Euler angles (ψ, θ, φ) which are related to the macroscopic axis of

the sample, defined as rolling direction (RD) axis, transversal direction (TD) axis and normal

direction (ND) axis through a set of three consecutive rotations that must be given to each

crystallite in order to bring its crystallographic axes into coincidence with the specimen axes.

23

Figure 2.11 shows the rotations where ψ represents a rotation around the ND axis, θ represents a

rotation around the TD axis and φ represents a second rotation around the ND axis.

Figure 2.11 - Three consecutives Euler rotations defining an orientation.

ODF is a three dimensional description of texture but direct measurement of ODF is not

possible since conventional texture goniometry is only capable of determining the distribution of

crystal poles of diffracting planes normal, i.e., pole figures. Mathematical models have been

developed which allow the ODF to be calculated from the numerical data obtained from several

pole figures. Therefore, in order to compute the orientation distribution function for a

polycrystalline sample, pole figures measurements are required. The number of pole figures

needed for ODF calculation depends upon the crystal symmetry of the sample that is being

measured. For HCP materials, as it is the case of titanium, a minimum of five pole figures are

needed. The most widely adopted methods for calculating ODFs are those proposed

independently by Roe (1965) and by Bunge (1982), who used generalized spherical harmonic

functions to represent the crystallite distributions. The three Euler angles employed by Bunge to

describe the crystal rotations are φ1, Φ and φ2, whereas the set of angles proposed by Roe are

referred to as ψ, θ and φ respectively. The relationships between the Bunge and the Roe angles

are the following:

φ1 = π/2 - ψ; Φ = θ; φ2 = π/2 - φ

24

According to Roe, 1965, an ODF may be expressed as a series of generalized spherical

harmonics in the form of equation (2.2):

∞ l l

(ψ, θ, φ) = Σ Σ Σ Wlm Zlmn (cos θ). exp (-imψ). exp(inφ) (2.2)

l=0 m=-1 n=-1

Where Wlmn are the series coefficients and Zlmn (cosθ) is a generalization of the associated

Legendre functions, the so-called augmented Jacobian polynomials.

For hexagonal/orthotropic crystal/specimen symmetry, a three-dimensional orientation

volume may be defined by using three orthogonal axes for ψ, θ and φ with each of the Euler

angles ranging from 0 to 90°. The value of the orientation density at each point in this volume is

simply the intensity of that orientation in multiples of random units. Regions of higher and lower

orientation density are separated by three-dimensional contour surfaces and it is usual to take a

series of parallel sections through this space for ready visualization of the data contained in the

three-dimensional plot. In the case of hcp materials, due to their crystal symmetry, the

fundamental space can be reduced to the space spanned by the Euler angles ψ (from 0 to 90°),

θ(from 0 to 90°) and φ(from 0 to 60°) with sections every 5 or 10 degrees. Davies [Davies et al.,

1971] published a set of charts for hexagonal materials designed to aid on the task of indexing

the texture components of rolled materials with hexagonal symmetry. In this development

Davies has used a definition of Euler angles by Roe and has taken crystal directions <0002>

parallel to ND and <1010> parallel to RD (see figure 2.12).

Figure 2.12 – Relationship between sample and crystal axis directions.

25

The charts published by Davies are shown in the figure 2.13 and in figure 2.14 an

example of the advantage of using the ODF in texture analysis.

Figure 2.13 – Constant φ sections through the Eulerian space: a)0°, b)20°, c)30°, d)40 and e)60°

26

Figure 2.14 – Location of the cold rolling and recrystalization components on the constant phi sections of the Euler space using Roe’s definition [Roe, 1965]. 2.3.1- Cold Rolling Texture

Hexagonal materials, such as, titanium and zirconium, have a limited number of slip

systems and generally develop a strong texture after cold rolling. Knight, 1978; investigated the

texture evolution of commercially pure titanium sheets after cold rolling at 21.4% and 89.4% of

reduction and observed that the most intense texture component, for both degrees of reduction

was the (2115) [0110]. Guillaume et al., 1981, when working with cold rolled titanium sheets,

found the same result. The (2115) [0110] orientation is 35° around the (0002) pole in the

transversal plane, which involves a rotation of the (0002) pole around the rolling direction, in the

plane defined by the transversal and the normal directions. Philippe et al. [Philippe, 1984], have

also found the same texture components after cold rolling of titanium and zirconium alloys.

27

Inagaki [Inagaki, 1991] working with hot rolled and annealed pure titanium presenting a very

strong texture, found that after cold rolling reductions below 30% the textures were weakened by

twinning and slip rotations. At cold rolling reductions between 30 and 50% twinning occurred

less frequently and at rolling reductions above 50%, crystal rotation about <0110>//RD axis is

induced by slip deformation. Orientations located near the 0001 <0110> were rotated toward

the 2115 <0110> orientation, becoming stable at this orientation at rolling reductions above

80%. Inagaki also found that the [0001]//ND fiber texture increased remarkably at rolling

reduction between 30 and 50% and that it decreased rapidly at rolling reductions above 50%. The

[0110]//RD fiber, on the other hand, developed at rolling reductions above 50%.

2.3.2- Hot and Warm Rolling Texture

In the past, hot rolling textures in titanium have been studied by only few investigators

and warm rolling textures in titanium have called even less attention from the researchers.

Inagaki [Inagaki, 1990] investigated the effect of hot rolling temperature (750, 800, 850, 900 and

950°C) on the development of hot rolling textures on commercially pure titanium plates.

According to Inagaki, the textures observed in the specimens hot rolled at temperatures below

800°C are essentially the same as the cold rolling texture and their main orientation is 2115

<0110>. Hot rolling at temperatures between 800 and 850°C enhances the development of the

2110 <0110> and 2118 <8443> main orientations, which seem to be formed by the

recrystallization that occurs during and after hot rolling. Hot rolling at temperatures above 880°C

results in the formation of a strong transformation texture where the 2110 <0110> texture

component, derived from the BCC β phase rolling texture, is the main orientation.

2.4 – X-ray Peak Profile Analysis

In order to improve and to control the mechanical proprieties of any material it is

important to understand and to explain how variables such as dislocation density, dislocation

type and slip system activation affect the formation and evolution of certain microstructures

28

during the deformation process. The study and determination of the dislocations slip systems

type is usually carried out with conventional techniques such as TEM. However, when the

material is highly deformed and the dislocation density reaches values as high as 1010/cm2, TEM

analysis is rather difficult. Also, throughout the sample preparation process required for TEM

experiments the original microstructure may change. Other alternatives on investigating the

microstructure are X-ray and neutron diffraction techniques. In recent decades, new applications

for the X-ray diffraction method (traditionally used for phase identification, quantitative analysis

and the determination of structure imperfections), have extended its usage to new areas, such as

the determination of crystal structures and the extraction of microstructural properties of

materials. Recent works have shown that X-ray diffraction peak profile analysis (XDPPA) is a

powerful alternative to transmission electron microscopy for describing the microstructure of

crystalline materials and providing information about the dislocation densities and dislocation

type extracted from the X-ray pattern [Ungár, 1999; Ribárik, 2001; Dragomir, 2002; Scardi,

2002; Glavicic, 2004; Scardi 2004; Ungár, 2004; Dragomir, 2005a and 2005b]. Besides that,

since the parameters provided by the two different methods are never identical, XDPPA is also

complementary to TEM enabling a more detailed understanding of microstructures.

X-ray diffraction peaks broaden when the crystal lattice becomes imperfect. The

microstructure means the extent and the quality of lattice imperfectness. According to the theory

of kinematical scattering, X-ray diffraction peaks broaden either due to crystallites smallness

(≈1μ ), lattice defects are present in large enough abundance ( in terms of dislocations this means

a dislocation density larger than about 5x1012m-2), stress gradients and/or chemical

heterogeneities.

Peak broadening is caused by crystallite smallness, lattice defects, stress gradients and/or

chemical heterogeneities. As a consequence of these deviations from perfect crystalline lattice

the shape of the X-ray diffraction lines no longer consists of narrow, symmetrical, delta-function

like peaks, such as the diffraction lines given by an ideal powder diffraction pattern. The

aberrations from the ideal powder pattern can be conceived as: (i) peak shift, (ii) peak

broadening, (iii) peak asymmetries, (iv) anisotropic peak broadening and (v) peak shape. The

main correlation between these peak aberrations and the different elements of microstructure are

summarized in table 2.5.

29

Table 2.5 – The most typical correlations between diffraction peak aberrations and the different elements of microstructure (Ungár, 2004). Sources of strain Peak aberrations

shift broadening asymmetry Anisotropic

broadening

shape

Dislocations √ √ √ √

Stacking faults √ √ √ √ √

Twinning √ √ √ √ √

Microstresses √

Long-range internal stresses √ √

Grain boundaries √ √

Sub-boundaries √ √

Internal stresses √

Coherency strains √ √ √

Chemical heterogeneities √ √ √

Point defects √

Precipitates and inclusions √ √

Crystallite smallness √ √ √

The effect of these defects can be divided into two main types of broadening: size- and

strain broadening. The first depends on the size of coherent domains and may include effects of

stacking and twin faults and sub-grain structures (small-angle grain boundaries) whereas the

latter is caused by different lattice imperfection, especially dislocations. The two different effects

interplay with each other and very often are not easy to separate. Krivoglaz [Krivoglaz, 1969]

has shown that strain broadening can be described, in general, in terms of broadening caused by

dislocations. In the case of single crystals or coarse-grained polycrystalline materials, strain

broadening caused by dislocations can be well described by a special logarithmic series

expansion of the Fourier coefficients [Krivoglaz, 1969; Wilkens, 1970; Groma et al., 1988,

Ungár et al., 1989]. When grain size plays a role, the two effects (i.e. size and strain broadening)

overlap. In such cases the grain size or the properties of the dislocation structure can only be

30

determined by the correct separation of the two effects. Two classical procedures are employed

in order to separate the strain and domain-size components of the broadening: Williamson-Hall

method and Warren-Averbach method. The first procedure [Williamson and Hall, 1953] is based

on the full width at half maximum (FWHM) and the integral breadths while the second is based

on the Fourier coefficients of the profiles [Warren and Averbach, 1950; Warren, 1959]. The

particle-size and dislocation microstrains are convoluted but can be separated, because the

particle-size broadening is independent of the order of the diffraction line, whereas the strain

broadening is not. In the Warren-Averbach method, the diffraction line profile is transformed

into its Fourier components and processed in order to separate the two broadening effects (after

correction for instrumental broadening). Evaluations carried out with both methods provide

apparent size parameters of crystallites or coherently diffracting domains and values of the mean

square strain but grain shape anisotropy and also strain anisotropy can turn difficult and

complicate the evaluation process [Louër et al., 1983; Caglioti et al., 1958]. In practical terms,

strain anisotropy means that neither the full width of half maximum (FWHM) in the Williamson-

Hall plot [Williamson and Hall, 1953] nor the Fourier coefficients in the Warren-Averbach

analysis [Warren and Averbach, 1952; Warren, 1959] are smooth functions of the diffraction

vector g. Ungár proposed that a way to interpret strain anisotropy is to assume that dislocations

are one of the major sources for lattice distortions [Ungár and Borbély, 1996]. Two different

approaches can well account for the phenomenon, especially in the case of random

polycrystalline or powder specimen. One is a phenomenological approach assuming that the

random displacements of atoms are weighted by the anisotropic elastic constants of the crystal

[Stephens, 1999] and the FWHM is scaled by the fourth order invariants of the hkl indices, given

for different crystal classes e.g. by Nye, (1957) or Popa, (1998). The other approach operates

with the anisotropic diffraction contrasts of dislocations [Stokes and Wilson, 1944; Ungár and

Borbély, 1996]. In the case of randomly oriented polycrystalline or powder specimen the

dislocation model has been shown to be formally equivalent to the phenomenological approach

[Ungár and Tichy, 1999] and the model is able to provide quantitative results, which have

physical relevance to the microstructure of the crystal [Cheary et al., 2000]. An advantage of this

model is that it also works in the case of a heavily deformed polycrystalline material or a single

crystal [Mohamed et al., 1997; Cheary et al., 2000; Borbély et al., 2000], situations in which a

strong preferred orientation is present.

31

In polycrystalline material populated with dislocations the anisotropic line broadening

can be taken into account by using that the dislocation model of the mean square strain, <εg,L2>,

(where L is the Fourier length and εg is the distortion tensor component in the direction of the

diffraction vector, g) [Wilkens, 1970a and 1970b]. In this model the dislocations are assumed to

have a restrictedly random distribution within a region defined by Re as the effective outer cut-

off radius [Wilkens, 1970a]. Here the anisotropic effect can be summarized in the average

contrast factors, C, which depends on the relative orientations of the line and Burgers vectors of

dislocations and the diffraction vector [Ungár and Borbély, 1996; Wilkens, 1970b; Klimanek and

Kuzel, 1988; Kuzel and Klimanek, 1988 and 1989; Ungár and Tichy, 1999; Dragomir and

Ungár, 2002]. The contrast factor of dislocations is a measure of the “visibility” of dislocations

in the X-ray diffraction experiments. Since, the contrast effect is mainly a characteristic of

dislocations, the theoretical values of the contrast factors and those obtained from the profile

evaluation enable the determination of the active dislocation slip system(s) in the studied sample

[Klimanek and Kuzel, 1988; Kuzel and Klimanek, 1988 and 1989; Ungár and Tichy, 1999;

Dragomir and Ungár, 2002].

Because of the complexity of the mechanical properties of hexagonal crystals [Chung &

Buessem, 1968; Gubicza et al., 2000; Solas et al., 2001; Tomé et al., 2001] for a better

understanding of the bulk dislocation structure and the Burgers vector populations it is desirable

to complement TEM studies by X-ray diffraction profile analysis. When comparing the

hexagonal crystal to the cubic crystal it becomes evident the higher level of complexity involved

when dealing with the hexagonal systems. Instead of three elastic constants hexagonal crystals

exhibit six elastic constants and two lattice constants (c and a) while cubic systems have only one

(a). Moreover instead of one, hexagonal crystal present two different types of anisotropy: shear

and compression [Chung and Buessem, 1968]; and finally, while in cubic systems there is one

major slip system, in hexagonal there are three different major slip systems related to the three

glide planes: basal, prismatic and pyramidal. If it is taken in account the different glide directions

and dislocation character (i. e., edge and screw) it is possible to group the slip systems into

eleven sub-slip-systems as shown in Table 2.6 [Yadav and Ramesh, 1977; Jones and Hutchinson,

1981; Honeycombe, 1984; Castelnau et al., 2001]. Dragomir and Ungár (2002) have recently

published a modified methodology to obtain the contrast factors for both cubic and hexagonal

crystals. They concluded that, at the actual state of the art regarding line broadening analysis, it

32

is impracticable to compile the dislocation contrast factors for hexagonal systems in a similar

manner as it was done for cubic crystals and instead, they proposed to compile the average

contrast factors of the sub-slip-systems. The average contrast factor of a specific sub-slip-system

in a hexagonal crystal can be given by three parameters versus the fourth-order invariant of the

hkil Miller indices [Ungár and Tichy, 1999]: C hk.0, q1 and q2 and once these parameters are

determined all average contrast factor corresponding to the sub-slip-system in question can be

obtained [Dragomir and Ungár, 2002].

Table 2.6 - The most common slip systems in hexagonal crystals: (a) Edge dislocations and (b) Screw dislocations. (a) Edge dislocations:

Major slip systems

Slip-systems Burgers vector Slip plane Burgers vector types

Basal

BE >< 0112 0001 a

PrE >< 1102 0101 a

PrE2 >< 0001 0101 c

Prismatic

PrE3 >< 1132 0101 c + a

PyE >< 0121 1110 a

Py2E >< 1132 2112 c + a

PyE3 >< 1132 1211 c + a

Pyramidal

PyE4 >< 1132 1110 c + a

(b) Screw dislocations:

Slip-systems

Burgers vector Burgers vector types

S1

>< 0112 a

S2

>< 1132 c + a

S3

>< 0001 c

33

2.4.1 - X-ray Peak Profile Analysis from MWP and Methodology for Determining the

Burgers Vector Populations

It is well known that the Fourier coefficients of the physical profiles can be written as a

multiplication of the Fourier coefficients corresponding to the size and distortion effect [Ungár

and Tichy, 1999]:

AL = ALS AL

D = ALS exp [- 2π2L2g2 <εg,L

2> ] (2.3)

where S and D indicate size and distortion, g is the absolute value of the diffraction vector,

<εg,L2> is the mean square strain and L is the Fourier variable. As shown in [Wilkens, 1970a and

1970b; Krivoglaz, 1996] in a dislocated crystal the mean square strain can be written in terms of

dislocation density and the strain anisotropy, which can be taken in account by introducing the

dislocation contrast factors, <εg,L2> ≅ (ρ 2Cb /4π) f( ), where and b are the density and the

modulus of the Burgers vectors of dislocations and 2Cb is the average contrast factor of the

dislocations present in the sample multiplied by the square of the dislocations burgers vector.

f( )-function is the Wilkens’s function, where =L/Re, Re is the effective outer cut-off radius of

dislocations, L is the Fourier length defined as nλ/2(sinθ2-sinθ1) with n being an integer starting

from zero, λ the x-ray wave length and (θ2-θ1) the angular range of the measured profile

[Wilkens, 1970a].

Contrast effect of dislocations depends not only on the material, but also on the relative

orientation of the diffraction vector, g, line vector, l, and Burgers vector, b [Klimanek and Kuzel,

1988; Kuzel and Klimanek, 1988 and 1989; Dragomir and Ungár, 2002]. Due to this in the case

of hexagonal crystals the three major slip systems (basal, prismatic and pyramidal) have to be

divided in 11 sub-slip systems by taking into account the different slip system types and the

dislocation character (i.e. edge or screw). These eleven sub-slip-systems are illustrated in figure

2.16 and listed in table 2.6. It has been shown in earlier studies that in the case of hexagonal

crystals for a given sub-slip system the average contrast factor of dislocation can be written as

[Dragomir and Ungár, 2002]:

34

C hk.l = C hk.0 [1 + q1x + q2 x2 ] (2.4)

where x = (2/3)(l/ga)2, q1 and q2 are parameters which depend on the elastic properties of the

material, C hk.0 is the average contrast factor corresponding to the hk.0 type reflections, a is the

lattice constant in the basal plane, g is the diffraction vector and l is the last index of the hk.l

reflection for which the C hk.l is evaluated. The equation (2.4) is valid only when it can be

assumed that within a sub-slip system the dislocation can slip with the same probability in all

directions permitted by the hexagonal crystal symmetry. This equation also means that the

average contrast factors corresponding to a specific sub-slip-system and material constants

(elastic constants C11/C12, C13/C12, C33/C12, C44/C12 and the lattice constant c/a) have to follow a

parabola as a function of x having the parameters C hk.0, q1 and q2 as parameterization

parameters. In the case of Titanium, the parabolas corresponding to the eleven sub-systems

described in table 2.6 and shown in figure 2.16 are presented in the figure 2.15.

Figure 2.15 – The parabolas describing the average contrast factors for the eleven slip systems, in the case of Titanium, as a function of x = (2/3)(l/ga)2 [Dragomir et al., 2002].

35

BASAL

<2 1 1 0> 0001

PRISMATIC

< 2 110>01 1 0 < 2 113>01 1 0 <0001>1 1 00

PYRAMIDAL

< 1 2 1 0 >10 1 1 < 2 113 >10 1 1 < 2 113 >11 2 1 < 2 113 >2 1 1 2

Figure 2.16 - Slip systems in hexagonal crystal systems [Honeycombe, 1984; Klimanek and Kuzel,1988]. As it has been shown by Dragomir and Ungár, in the case of hexagonal crystal the measured

average )(

2m

Cb characteristic to the examined sample can be written as follows [Dragomir and

Ungár, 2002]:

)(2

m

Cb =∑=

N

i

i

i

i bCf1

2)( (2.5)

36

where N is the number of the different activated sub-slip systems, )(i

C is the theoretical value of

the average contrast factor corresponding to the ith sub-slip system and fi are the fractions of the

particular sub-slip systems by which they contribute to the broadening of a specific reflection.

On the left hand side of the equation (2.5) the m superscript refers to the measurable strain

anisotropy parameter, 2Cb . For the hexagonal crystal structure, equation (2.5) can be written for

the three fundamental Burgers vectors types defined in the hexagonal systems: b1=1/3<2110>,

b2=<0001>, and b3=1/3<2113>:

)(

.2

m

lhkCb = 21b ∑

><

=

aN

i

i

i Cf1

)( + 22b ∑

><

=

cN

j

j

j Cf1

)( + 23b ∑

>+<

=

acN

n

n

n Cf1

)( (2.6)

where N<a>, N<c> and N<c+a> are the numbers of sub-slip systems with the Burgers vector

types <a>, <c> or <c+a>, respectively. The possibility of measuring q1 and q2 offers three

independent equations and eleven unknowns. It means that equation (2.6) can give an exact

solution only by making certain assumptions about the activated dislocation slip systems. In the

present work it is assumed that a particular Burgers vector type has random (or uniform)

distribution in the different slip systems. In this case equation (2.6) can be written:

)(

.2

m

lhkCb =∑=

3

1

2)(

i

i

i

i bCh (2.7)

where hi is the fraction of the dislocations population in the sample with the same Burgers

vector, bi. )(iC is the averaged contrast factor over the sub-slip systems, for the same Burgers

vector type. Inserting equations (2.4) into (2.7) the following three equations are obtained:

)(1

mq = ∑=

3

1

)(1

2)(0.

1

i

i

i

i

hki qbChP

, , (2.8)

37

)(2mq = ∑

=

3

1

)(2

2)(0.

1

i

i

i

i

hki qbChP

(2.9)

∑=

3

1i

ih =1 (2.10)

where P=∑=

3

1

2)(0.

i

i

i

hki bCh =)(

0.2

m

hkCb and 0≤ hi ≤1. To solve equations (2.8), (2.9) and (2.10) the

numerically calculated values of C hk.0, q1 and q2 for all sub-slip systems are required. The

theoretical values of C hk.0, q1 and q2 for the most common sub-slip systems were published

previously elsewhere [Dragomir and Ungár, 2002]. The contrast factor of dislocation for

hexagonal crystals in elastic anisotropic and isotropic media have been treated by Kuzel and

Klimanek [Klimanek and Kuzel, 1988; Kuzel and Klimanek, 1988 and 1989].

The measured values of q1 and q2 parameters are obtained from the Multiple Whole

Profile (MWP) fitting procedure [Ribárik et al., 2001]. In this procedure the Fourier-transformed

of multiple hkl reflections are fitted simultaneously by equation (2.3). Here, throughout the q1

and q2 parameters in equation (2.4), the dislocation contrast factor becomes a fitting parameter.

Finally, the q1(m) and q2(m) parameters obtained for the sub-slip-systems families are used in the

analysis of Burgers vector populations [Dragomir and Ungár, 2002] providing a prediction of

slip systems activity during the deformation process.

2.5 – Self-Consistent Modeling of Deformation Texture

The modeling of deformation texture evolution based on the formulation of the plasticity

of polycrystalline materials has received considerable attention and has been the object of many

studies and different approaches [Voigt, 1889; Reuss, 1929; Morris, 1970; Kroner, 1972;

Hutchinson, 1976; Ahzi, 1987; Canova et al., 1989].

As early as 1938, Taylor proposed a uniform strain model, which assumes that the

imposed plastic strain in each grain is identical to the macroscopic plastic strain [Taylor, 1938].

38

Models based upon Taylor’s assumption [Bishop and Hill 1951; Bunge 1970] have often

demonstrated first-order agreement with the measurement of mechanical anisotropy of

polycrystalline materials. However, the hypothesis of plastic strain uniformity is somewhat crude

and it has been shown to fail when plastic strain heterogeneities are evident, as it is the case of

uniaxial compression of fcc polycrystalline metals, uniaxial tension of bcc polycrystalline metals

and in general for polyphase materials.

For the case of small deformations, a self-consistent approach was proposed and

developed by different authors [Hill, 1965 and 1967; Hutchinson, 1970and 1976; Kröner, 1961;

Berveiller-Zaoui, 1980]. This approach takes into account the interaction of each grain with the

surrounding environment. Here, the medium (surrounding) is replaced by an equivalent medium

assumed to be a homogeneous equivalent medium (H.E.M). An extension to large elastoplastic

deformations was proposed by Iwakuma and Nemat-Nasser [Iwakuma and Nemat-Nasser, 1984]

and one application to large elastoviscoplastic deformations by Nemat-Nasser and Obata

[Nemat-Nasser and Obata, 1987]. An extension to large deformation plasticity has been

developed by Iwakuma and Nemat-Nasser [Iwakuma and Nemat-Nasser, 1984] and used for

two-dimensional problems. Hutchinson [Hutchinson, 1976] used an upper bound theorem and

self-consistent model proposed by Hill [Hill, 1965 and 1967] to derive macroscopic relations

between stress and strain-rates for a creep power law polycrystalline material in small

deformations.

The so-called Relaxed Constraints theory [Honneff and Mecking, 1978; Kocks and

Canova, 1981; Houtte, 1981] takes into account the strain heterogeneities produced by

anisotropic grain shapes and the predicted results have a better agreement with experiments. This

is true particularly in some specific cases, like large strain rolling and torsion of fcc metals.

Asaro and Needleman have proposed an extension of Taylor theory for large deformations

including elastic deformations [Asaro and Needleman, 1985]. It should be noticed that all

theories based on Taylor’s strain uniformity only fulfill compatibility conditions but not

equilibrium conditions at grain boundaries.

In 1987, using a scheme developed by Zeller and Dederichs (1973) in heterogeneous

elasticity, Molinari, Canova and Ahzi [Molinari et al., 1987] proposed a self-consistent approach

for the large deformation polycrystal viscoplasticity which they applied to the prediction of

deformation textures in the cases of tension, compression, rolling and torsion [Molinari et al.,

39

1987]. Almost at the same time, Nemat-Nasser and Obata [Nemat-Nasser and Obata, 1987] also

proposed a model to approach the macroscopic behavior of an elastic viscoplastic polycrystal by

using Hill’s self-consistent model [Hill, 1989]. They applied their theory to a plane problem with

double slip modelization of the single crystal behavior.

In this work it was used the approach proposed by Molinari, Canova and Ahzi [Molinari

et al., 1987], which consists of a self-consistent formulation for the large deformation polycrystal

viscoplasticity. The properties of the polycrystal are developed from the single crystal behavior

[Zeller and Dederichs, 1973; Berveiller and Zaoui, 1980]. In this approach, equilibrium and

incompressibility equations are used to arrive at an integral equation for local velocity gradient.

This integral equation can be solved via different approximate schemes. In the self-consistent

model of Molinari [Molinari et al., 1987], a grain is considered as an inclusion embedded in a

homogeneous equivalent medium. The interaction law derived from the integral equation results

in a nonlinear relation between stress and strain rate, which is solved by a straightforward

Newton method.

2.5.1 – The Single-Crystal Constitutive Law

As in certain models [Hutchinson, 1976; Pan and Rice, 1983; Asaro and Needleman,

1985] of single-crystal behavior, here it is assumed that slip is slightly sensitive to velocity. We

also assume that the plastic-shear rate on slip system s depends on the shear stress τsγ& s resolved

on the system, according to the following law:

( )msss

00 γγττ &&= (2.11)

The parameter m is the strain-rate sensitivity coefficient. τs

is associated with the deviatoric

Cauchy stress tensor by the relation

SmSm s

ij

s

ij

s :==τ (2.12)

where

40

s

j

s

i

s

ij nbm = (2.13)

n

s and b

s are the normal to the slip plane and the slip direction, respectively, for slip system s.

and represent a reference stress and the corresponding reference strain velocity,

correspondingly.

s

0τ s

0γ&

The strain-velocity tensor D is defined by

Dij = ( υi,j + υj,i)/2 (2.14)

where υi are the components of the displacement velocity. The comma designates the derivative

with respect to the reference coordinates associated with the laboratory. The strain-velocity

tensor is associated with the microscopic shear velocity by the relation:

∑=s

ss

ijij rD γ& (2.15)

where,

( ) 2/s

ji

s

ij

s

ij mmr += (2.16)

By using the microscopic stress-strain relation (2.11) and the definition of resolved shear stress

(equation (2.12)), the single-crystal constitutive law, where elasticity is neglected, finally results

in a non-linear relation between the strain velocity and the deviatoric Cauchy stresses Skl:

n

s

kl

s

kls

ij

ij Srr

D

⎟⎟⎠

⎞⎜⎜⎝

⎛=

00 τγ& (2.17)

where,

n=1/m (2.18)

The variation of as a function of time characterizing the microscopic intracrystalline

work hardening and the variation of are associated with the orientations changes of the crystal

s

s

ijr

41

lattice. The development of the microscopic work hardening is represented by the following

relation:

∑=r

rsrs H γτ &&0 (2.19)

where the terms Hsr represent the elements of the microscopic strain-hardening matrix.

The crystal-lattice rotation-velocity tensor Ω* is obtained from the difference between the

total rotation-velocity tensor ΩP and the plastic rotation velocity.

Ω = Ω* + ΩP (2.20)

where, Ω is the antisymmetric part of the displacement-velocity gradient L

Ω = ( L – LT )/2 (2.21)

and ΩP is given bu the following relation

( ) 2/s

S

STTSP mm γ&∑ −=Ω (2.22)

Knowing the total-rotation velocity Ω and the microscopic shear velocity , it can be

determined the crystal-lattice rotation velocity from equation (2.20).

sγ&

The nonlinear relation (2.17) may be written in the form of:

Dij = Gij (S) (2.23)

By inverting this relation, one can express the stresses S as a function of the strain-

velocity tensor D

Skl = Fkl (D) (2.24)

The solution of equation (2.24) is unique since a convex viscoplastic potential exists. The

tangent behavior is obtained by the Taylor development to first order from the law (2.24)

concerning the applied strain velocity D’:

42

( ) 0' klmnklmnkl SDDAS += (2.25)

with

( ) ( )1

''−

⎥⎦⎤

⎢⎣⎡

∂∂

=∂∂

=klmnmn

klklmn

S

GD

D

FDA (2.26)

where (∂G/∂S) is the Jacobian matrix of the function G(S).

( )'0 DSkl is defined by:

( ) ( ) klmnklmnkl SDDADS +−= ''0 (2.27)

The tensor A has the following symmetry properties:

Aklmn = Alkmn = Aklnm = Amnlk

The Cauchy stress tensor is obtained by addition of the hydrostatic term to the deviator

term S.

= A (D):D + S0 (D) – pI (2.28)

where p is the hydrostatic pressure and I the identity tensor of second order.

2.5.2 – Polycrystal Constitutive Law

Locally, the law defining the behavior of the medium (HEM) is given by the equation

(2.28), where both A and S0 depend on the position r. In each grain A and S0 depend on the

microscopic work hardening and grain orientation.

Imposing to the polycrystal a macroscopically-homogeneous displacement-velocity

gradient L

43

jiij VL ,= (2.29)

which satisfies the incompressibility condition iiL = 0, and a macroscopic pressure p , the aim is

to obtain the local pressure p and the displacement-velocity L defined by:

Lij = Vi,j (2.30)

The equilibrium equation must be satisfied throughout the polycrystal and in the absence of body

forces, this may be expressed by:

σij,j = 0 (2.31)

According to equation (28.2), the equilibrium equation is written:

( ) 0,0,,, =−+ ijijjlkijkl pSVA (2.32)

The macroscopic constitutive law is represented by the tangent behavior

000 SDAS += (2.33)

where D is the applied strain velocity, and A0 and S00 are dependent of D .

Decomposition of the tensors A (and S0) into the sum of a uniform part A0 (and S00 ) and

a part dependent of the spatial position r leads to

( ) ( )rr AAA~0 +=

( ) ( )0000 ~rr SSS +=

(2.34)

By using this decomposition in equation (2.32) it is obtained the following Navier’s

equations:

0,,

0 =+− ijljkijkl fpVA (2.35)

where fi may be considered as a pseudo force:

44

( ) 0,,,

~~jijjlkijkli SVAf += (2.36)

Adding the incompressibility equation to the set represented by equations (2.35),

Vk,k = 0 (2.37)

the result is a system with four equations and the unknowns υi and p, which may be solved by

the method of Green’s functions.

The Green’s functions Gij and Hk are solutions to the following system of equations:

( ) ( ) 0')'(' ,,0 =−+−−− rrrrHrrGA imimljkmijkl δδ

0, =kkmG (2.38)

Considering an infinite medium such that at infinity Gij= phi and Hm= phi, the function δ(r –r’)

is the Dirac delta-function in r’. For fixed m, the term δimδ(r –r’) represents the ith component

of a unit force concentrated at point r’ and parallel to the direction m. For an infinite space, the

Green’s functions satisfy the following properties:

Gjm(r,r’) = Gjm(r – r’)

Hk(r,r’) = Hk(r – r’) (2.39)

The solutions υn and p of the system of equations (2.35) and (2.37) can be written in the

following form:

ininn fGVV *+=

ii fHpp *+= (2.40)

where the symbol * represents the convolution product:

( ) ( ) 3'''* drrfrrGfG iniini ∫ −= (2.41)

45

V is a linear function obtained by integration of the equation Vn,k equal to constant throughout the

volume. By using expression (2.36) for fi, and after differentiation of equation (2.40) it is

obtained:

]~

)~

[(* 0,,,,,, jijjlkijklmnimnmn SVAGVV ++= (2.42)

integrating by parts:

( ) 0,,,,,

~*

~* ijmjnilkijklmnimnmn SGVAGVV ++= (2.43)

If Ã0, Ã and 0~S are known, the equations (2.43) constitute a linear system of integral

equations whose unknowns are the Vk,l. Taking the symmetric part of equation (2.43) one

obtains:

( ) 0~*

~* ijnmijklijklmnijmnnm SDADD Γ+Γ+= (2.44)

with

Γmnij = (Gni,mj + Gmi,nj + Gnj,mi + Gmj,ni )/4 (2.45)

We shall discuss later how Ã0, Ã, 0~S and D may be calculated by an iterative process. For a

given local strain rate D, S is obtained by numerical inversion of equation (2.17) and one may

thus calculate the shear rate and the plastic rotation Ωsγ& p.

The total rotation rate is obtained by considering the asymmetric part of equation (2.43):

( ) 0~*

~* ijnmijklijklnmijmnnm SBDAB ++Ω=Ω (2.46)

With

BBnmij = (Gni,mj + Gmi,nj + Gnj,mi + Gmj,ni )/4 (2.47)

Knowing the total rotation Ω and plastic rotation Ωp rates, the lattice rotation rate Ω* can be

determined from equation (2.20).

46

2.5.3 – The Self-Consistent Approach

In order to solve the integral equation (34.2), the strain rate tensor D is assumed to be

uniform in each grain and thus disregarding the intracystalline heterogeneities that may appear in

the grains. Therefore, considering the uniformity of the strain rate in each grain, a first

approximation is made, so that

( ) ( )rgDDg

g

r Δ= ∑ (2.48)

where Δg(r) is the characteristic function of the grain g defined by equation (2.49) and Dg is the

uniform strain rate in Vg..

The characteristic function is given by

( )Vrif

Vrifrg

g

∉∈

=Δ0

1 (2.49)

where Vg is the volume of the grain and V is the complementary of Vg in the euclidian space.

Now, recalling that A and S0 depend on D, it follows that

( ) ( )rgAA g

r Δ= ∑

( ) ( )rgSSg

r Δ= ∑ 00 (2.50)

with Ag and S0g uniform in the grain and dependent on Dg.

The substitution of equations (2.48) and (2.50) into equation (2.44) results in the

expression (2.51):

( )∑ ∫ −Γ++='

', ')'(

~~)( 3'0''

gVg

nmij

g

ij

g

kl

g

ijklmnmn rdrrSDADrD (2.51)

47

At this point, it is important to note that Dnm is not uniform and in order to fulfill the hypothesis

of uniformity of the strain rate, the uniform strain rate Dg can be taken as the volume average of

D in the grain g and after averaging the relation (2.51) it results in

( )∑ +Γ+='

' '0'' ~~

g

g

ij

g

kl

g

ijkl

gg

nmijmn

g

mn SDADD (2.52)

where

[ ]∫ ∫ −Γ=ΓVg Vg

gg

nmij

g

gg

nmij rdrdrrV

'

''

.)(1 3'3' (2.53)

If now it is assumed that Ãg and S0g are known for an infinitely extended polycrystal, the system

of equations (2.51) becomes a linear set with a infinite number of unknowns . Approximate

solutions of this system can be obtained by considering the interaction of the grain with its

nearest neighbors and replacing the remaining grains by the HEM whose behavior is assumed

identical to the macroscopic behavior of the polycrystal and is formulated in terms of the tangent

modulus A

g

mnD

0 as follows:

( ) ( )DSDDAS 000 += 43 (2.54)

Now, assuming the approximation of a grain with an ellipsoidal shape embedded in the HEM

and assuming the strain rate uniform in the grain and A0 and S00 uniform in the HEM, the system

of equations (2.52) can be taken as

[ ])(~

)(~ 0 gg

ij

g

kl

gg

ijkl

gg

mnijmn

g

mn DSDDADD +Γ+= (2.55)

where

rdrdrrV Vg Vg

nmij

g

gg

nmij

3'3' )(1 '

∫ ∫ ⎟⎠⎞⎜

⎝⎛ −Γ=Γ (2.56)

48

The Strain rate Dg in the grain is calculated using the equation (2.55) and to derive the reaction

stresses in the inclusions due to the difference between the strain rate in the inclusion and that at

infinite, taking into account the relationship (2.34), equation (2.55) is rewritten as

( ) )(::: 0 DDASSDD ggggggg −Γ−−Γ=− (2.57)

or,

[ ]( ) ( )DDASS gggg −+Γ=−−

:0

1 (2.58)

In a similar way, a relationship for the rotation rates Ωg is given by

[ ]( )( )DDB gggggg −Γ=Ω−Ω−1

: (2.59)

with

rdrdrrBV

BVg Vg

nmij

g

gg

nmij

3'3' )(1 '

∫ ∫ ⎟⎠⎞⎜

⎝⎛ −= (2.60)

A more detailed explanation can be found in the works of Ahzi (1987) and Molinari et

Al. (1987).

49

CHAPTER 3

EXPERIMENTAL PROCEDURE 3.1 - Material

Commercially pure Titanium grade 2 (Timetal 50A) specimens were used in this

investigation. The as-received material was provided by TIMETTM in the form of plate

specimens (5 pieces) in as-rolled and Mill annealed conditions (ASTM-B-265), with dimensions

of 3.5” in length, 3” wide and thickness of 5/8”. The chemical composition, typical mechanical

and physical properties are shown in tables 3.1, 3.2 and 3.3, respectively.

Table 3.1 - Chemical composition (weight %)

Material O C N Fe H Residual Elements Titanium

CP Ti Gr2 0.25 0.008 0.03 0.30 0.015 0.40 Remainder

Table 3.2 - Typical mechanical properties of the CP Ti Gr2

Yield 0.2%

Ksi (MPA)

UTS

Ksi (MPA)

Elongation

(%)

Reduction in Area

(%)

Bend Radius

T (thickness)

Room Temp. 50 (345) 70 (485) 28

572°F

(300°C) 18 (127) 33 (229) 43

28 2.5 T

50

Table 3.3 - Physical properties of the CP Ti Gr2 Property English units S I units

Density 0.163 Lb in-3 4.51 g.cm-3

Elastic Modulus 15.2 ~ 17.4 Msi 105 ~ 120 Gpa

Thermal Conductivity 12.60 Btu hr-1ft-1°F-1 21.79 Wm-1K-1

Beta Transus 1680 °F 915 °C

Electrical Resistivity 21 μΩ•in 0.53 μΩ•m

Magnetic Permeability Nonmagnetic

The specimens were cut in an Electric Discharge Machining (EDM) equipment, resulting

in working samples, with dimensions 1”x3”x5/8”, which were used on the thermo-mechanical

processing. EDM procedures can alter properties such as fatigue through surface contamination

and residual surface stresses, hence the cutting surfaces which also happen to be the future

rolling planes, in this case, were ground to remove the thermally affected layer and to reduce the

possibility of any induced external effect as a consequence of the necessary machining of the as-

received material.

In the figure 3.1 it is presented an OIM/SEM picture representative of the microstructure

of the as-received material showing the grains and the presence of twinned grains as a result of

previous cold and hot working. The average grain size for the departing material is

approximately 30 microns.

3.2 - Thermo-Mechanical Processing

The thermo-mechanical processing of the samples was conducted at the rolling facility of

the Materials Science Department at the Georgia Institute of Technology, Atlanta - Georgia. The

equipment used was a conventional two-high rolling mill by Fenn, figure 3.2.

51

CI= 0.883 IQ= 94

Electropolishing solution ( 60 ml Per. Acid, 590 ml Methanol). Parameters: 20 V at –20oC for 3-4 min. Without etching

FEG/SEM 20V and 10 nA Probe current

Rolling plane

Rolling direction

Figure 3.1 – As received material: OIM/SEM micrograph (Courtesy of Dr. Ayman Salem – AFRL/MLLM)

The schematic setup for both processing sequences is shown in figure 3.3. Samples of the

as received material, hot rolled and annealed at TIMET, were rolled at room temperature, 25°C

(77°F), and warm rolled at 260°C (500°F). The same degrees of reduction were chosen for the

two different rolling temperatures, 20%, 40%, 60%, 80% and 95%. After rolling, coupon

specimens were taken out from each of the samples for characterization. Table 3.4 shows the

nomenclature of the samples according to their history.

52

Figure 3.2 – Rolling mill machine.

As Received Sample

Hot Rolling at 260C 20%, 40%

Cold Rolling 20%, 40%, 60%,

80%

(*) (*)

(*), 60% ,

80% an 95% reduction d , 95% reduction

Figure 3.3 – Schematic setup of the thermo-mechanical processing.

Texture Measurement Peak Profile Analysis

Deformation Texture Simulation

Texture Measurement Deformation Texture Simulation

(*)

53

Table 3.4 - Nomenclature of the samples. Sample History

AR As Received sample (hot rolled and annealed)

CR20 20% cold rolled sample

CR40 40% cold rolled sample

CR60 60% cold rolled sample

CR80 80% cold rolled sample

CR95 95% cold rolled sample

WR20 20% warm rolled sample

WR40 40% warm rolled sample

WR60 60% warm rolled sample

WR80 80% warm rolled sample

WR95 95% warm rolled sample

3.2.1 - Cold Rolling

In order to obtain an as-much-as-possible homogeneous deformed microstructure

evolution throughout the thickness, the desired final reductions were achieved by means of

unidirectional rolling with intermediary steps of 5% of reduction, i.e., to achieve 20% of

reduction 4 passes were employed, for 40% 8 passes were needed and so on.

3.2.2 - Hot Rolling

The hot working scheme in this investigation consisted of rolling at 260°C in

combination with five degrees of reduction ranging from 20% up to 95%. To heat the plate

samples it was used a Lindberg/Blue model BF51800 electrical furnace with nominal capacity of

1100°C. When heated, any material will require a certain time to have its temperature raised to

an intended level and also exhibits a homogeneous temperature distribution. This is commonly

known, as homogenization temperature, and as practice, in the case of steel as an example, it is

accepted to leave the sample inside the furnace, at the targeted temperature for at least 15

54

minutes per inch in thickness. However, intending to assure that a sample inside the furnace have

reached homogenization in terms of temperature, spare samples instrumented with an Omega K

type thermocouple, with capacity for 1100°C were used. Homogenization times of 10, 15, 20 and

30 minutes, already taken into account the amount of time needed after every load/unload

operation were tested and it was decided to work with a homogenization time of 30 minutes for

samples with nominal thickness from 5/8” down to 3/16” and from this point down to the last

reduction step, around 0.7mm (90~97%), it was used 15 minutes as homogenization time.

Following the same systematic procedure used for the cold rolling, the samples were

rolled up to nominal reductions of 20%, 40%, 60%, 80% and 95% in thickness with steps of 5%.

With the furnace at the corresponding soaking temperature, and only after temperature

stabilization, the sample would be placed inside the furnace for heating up to the rolling

temperature and then removed for one rolling pass. After each pass of 5%, the sample was

immediately returned to the furnace for reheating and execution of the next pass. After achieving

the desired final reduction the sample was allowed to cool down. This cycle was repeated for

each one of the samples from the first till the last pass, one sample at a time.

3.3 - Metallographic Sample Preparation

Before microstructural characterization the samples require metallographic preparation

such as grinding, polishing and etching. The characterization technique, normally dictates the

proper sample preparation procedure.

Another aspect of sample preparation was the need of cutting and sectioning both the as

received material as well as the rolled samples. These operations were carried out using an EDM

machine for coarse cutting and precision cutting/sectioning with a Struers Accutom 5 equipped

with the carbide disk saw recommended by Struers for titanium. The samples were then ground

to remove any deformed and/or heat affected layer as it was the case of the samples machined by

EDM.

55

3.3.1 - Mechanical Polishing

For the mechanical polishing the specimens were hot mounted in phenolic resin at 160°C

and 4200 PSI using Buehler Mounting Press, model Simplimet 3, machine. The

grinding/polishing procedure used was a modified scheme (table 3.5) derived from a technical

note from Struers. The grinding was performed using silicon carbide paper 320, 500, 800, 1000,

1200 grit at 150~300rpm until flat with water as lubricant fluid. In order to eliminate any

deformed or stressed layer formed during the grinding and polishing operations, an oxide

polishing formed by a solution of OP-S (colloidal silica with grain size of approximately

0.04mm and a pH of about 9.8) and H2O2-30% (70~90% of OP-S and 10~30% of H2O2-30%)

at 150 rpm for 10 minutes to 20 minutes was used.

3.4 - Characterization Techniques

In order to measure texture of the as received and deformed samples, X-ray diffraction

measurements were performed at the National High Magnetic Field Laboratory at Tallahassee-

FL. The X-ray measurements for line broadening and peak profile analysis were performed at the

X-Ray facility kindly provided by the School of Material Science and Engineering of the

Georgia Institute of Technology, Atlanta-Ga.

Table 3.5 - Metallographic preparation procedure.

Agent Cloth Lubricant Speed (rpm) Pressure Time

sandpaper P320 SiC Tap water 300 hand As required

sandpaper P500 SiC Tap water 150 hand As required

sandpaper P800 SiC Tap water 150 hand As required

sandpaper P1000 SiC Tap water 150 hand As required

sandpaper P1200 SiC Tap water 150 hand As required

Colloidal Silica 0.04

μm MD-Chem

OP-S (70%~90% )

+ H2O2 30%

(30%~10% )

150 10 N 10 min or

more

56

3.4.1 - Texture Measurement

The texture measurements were performed using a Philips X’Pert PW 3040 MRD

equipped with texture goniometer mounted with the axis vertical (figure 3.4). The machine

settings used during the measurements (voltage and current) was 40kV and 45mA. Scan

measurements from 20° to 90° (2θ angle) were run in all samples in order to determine the exact

position of the peaks. Five incomplete pole figures: (0002), (1011), (1012), (1120) and (1013)

were scanned in a 5° by 5° grid with Cu Kα (λ = 1.54 angstroms) radiation. The rw1 data files

generated by the Philips software were converted into raw archives, recognizable by PopLA

[Kallend et al., 1991], using PC-Texture 3.0. The PopLA software was used to recalculate the

five incomplete pole figures measured and to generate the ODFs.

In order to investigate the texture gradients, measurements were carried out through-

thickness at depths of 5% (19/20t), 20% (8/10t), 35% (15/20t) and at the mid-thickness (t/2) of

each sample with a separate set of samples cold rolled and cold rolled and duplex annealed. The

results for this are being shown in the Appendix.

Figure 3.4 - X-ray machine Philips X’Pert MRD equipped with texture goniometer.

57

3.4.2 - Peak Profile Measurements

The diffraction profiles necessary for the Peak Profile Analysis were measured using an

Alpha-1 Panalytical Diffractometer set up in Bragg-Brentano geometry. With the use of

symmetrical incident beam Johansson monochromator only the Kα1 component of Cu radiation

was used. In order to reduce the instrumental broadening effect 1/4o

divergent slit, 1/2o

anti-

scattering slit and 0.02 rad soller slits was used on the incident beam path. On diffracted beam

side a 5.0 mm anti-scattering slit and a 0.02 rad. soller slit was used. A mask of 5mm was used to

adjust the size of the probing X-ray spot. The profile data acquisition was done using a solid-

state position-sensitive ultra-fast detector (X’Celerator, Panalytical).

A set of three samples for each degree of deformation investigated here was cut (and

prepared for measurement) from the rolled specimens as shown schematically in figure 3.5. The

reason for this is the need for emulation of a “randomly oriented” material since the method is

originally suited for this and the specimens in study here are highly anisotropic after being

heavily rolled. For each set of sample the reflections (0002), (1011), (1012), (1120), (1010),

(1013) and (0004) were measured.

Figure 3.5 - Surfaces examined by X-ray diffraction: normal direction (ND); rolling direction (RD); transverse direction (TD).

58

The instrumental broadening was measured using LaB6660aNIST standard. Figure 3.6

shows the (220) reflection of LaB6 and the (11.0) reflection of α-Ti deformed at 60% reduction

rate. As the measured profile is a convolution of the physical with the instrumental profile the

Stokes-correction [Stokes and Wilson, 1944] based on the Fourier transforms of the profiles was

used to determine the physical line profiles. Background and instrumental profile correction were

done with the MKDAT program described elsewhere [Ribarik et al., 2001].

-0.1 0.0 0.1

0.0

0.3

0.6

0.9220 reflection of

LaB6, NIST 660a standard11.0 reflection of

α-Ti 60% reduction

No

rma

lise

d I

nte

nsity

Δk [1/nm]

Figure 3.6 - Example of the instrumental broadening of the Alpha-1 Panalytical Diffractometer measured using LaB6 660a NIST standard compared with the peak broadening measured for deformed α-Ti. The dashed line is the 220 reflection of LaB6 and the continuous line is the 11.0 reflection of α-Ti deformed at the 60% reduction rate.

The X-ray diffraction profile analysis is carried out using the Multiple Whole Profile

fitting procedure (MWP) and is described in detail elsewhere [Ribárik et al., 2001, Dragomir et

a., 2002].

59

CHAPTER 4

RESULTS

This chapter will present the results of microstructure evolution, line broadening, and

texture simulation by means of self-consistent approach for cold and warm rolled alpha-titanium

samples. This chapter will be divided in three distinct sections. Section 4.1 will show the texture

results, obtained through X-Ray measurements, of the resultant microstructures after cold and

warm deformation. The texture results will be mainly presented in the form of pole figures and

ODFs; and as the most important texture components for titanium are presented at the φ = 0° and

φ = 30° constant ODF sections only results of these two sections will be shown. Skeleton lines

and the development of the most interesting fiber texture for titanium, the (0002)//ND, as well as

the variation in its volume fraction with degree of reduction for both cold and warm rolling will

also be shown in section 4.1. Section 4.2 will show the results of X-ray peak profile analysis

extracted from X-ray diffraction pattern for the 40%, 60% and 80% warm rolled samples; and

section 4.3 will present the results of deformation texture simulation.

4.1 - Texture Evolution

4.1.1 - As Received Sample

The as received sample, about 16mm thick, was hot rolled and annealed at TIMET; and

characterized at the FSU/NHMFL facilities. The pole figures and ODFs for the as received

sample are shown in figures 4.1 and 4.2, respectively. According to figure 4.1 the texture of the

as received sample, nevertheless not strong, was not random showing that the <0002> directions

of most of the crystals were distributed on the plane formed by the normal and the rolling

60

directions. The orientation distribution function results shown in figure 4.2 confirms that the as

received material is formed by a initial texture of medium intensities, 3 times random at most,

with no presence of fibers or any important texture component.

(0002) Pole figure (1010) Pole figure

Contours: 0.5 1.0 1.5 2.0 2.5

Contours: 0.5 1.0 1.5 2.0

Figure 4.1 - (0002) and (1010) pole figure for the as received sample. θ = 0° θ = 90°

Contours: 1 2 3

ψ = 0°

ψ= 90°

Phi= 0.0

Phi=30.0

Figure 4.2 - ODF sections of φ =0° and φ =30°, Roe notation, for the as received sample. 4.1.2 – Cold Rolled Sample

Figure 4.3 shows the (0002) and (1010) pole figures for the cold rolled samples. In figure

4.3 (a), the 20% cold rolled sample, shows a low intensity, 3 times random, of (0002) poles

located at the center of the figure and at four other places, two towards the rolling direction and

two towards the transversal direction. As the cold reduction increases to 40%, fig. 4.3 (b), one

61

can see that the two poles towards the RD remain with the same intensity as before while the

poles located at the center and towards the TD have their intensities increased to 4 and 7 times

random, respectively. Moving forward in the degree of reduction it is possible to see that the two

poles towards the RD vanish while the poles at the center and towards the TD remain with their

intensities about the same value. It is interesting to note that as the cold rolling reduction

increases the shape of the circles around the center of the pole figure change from concentric to

oblong and the circles at the center decrease in radius becoming almost a point (fig. 4.3 (e)). The

intensity of the (0002) pole becomes a maximum at 95% reduction when it reaches 7 times

higher than a random material. The intensity of the (1010) poles does not seem to be affected by

the cold rolling, increasing from 2 times random in the AR sample to 2.5 times random after

20% of deformation and to 3 times random after 40% remaining at this value until 95% of cold

reduction. Although the intensity of the (1010) poles has not been affected by the degree of cold

reduction, the distribution of the (1010) poles has been changed and this changing can be

observed since the lowest degree of deformation, 20%, where the distribution of these poles start

assuming a more organized distribution towards the TD and the ND of the sample. After 40% of

deformation the distribution of the (1010) poles towards the TD vanishes and from this point

until 95% of cold deformation a concentration of these poles around the RD is observed.

Figure 4.4 shows the ODFs for the cold rolled samples and according to this figure the

general aspect of the crystallographic orientation distribution changes dramatically after cold

rolling. After 20% of reduction, the main texture components (around 4 times random) were the

(1013) [3032], the (1015) [1210], and the (3031) [1013], all located in the constant section of

φ=0°. As the degree of reduction increased, the (1013) [3032] component disappeared and the

(1015) [1210] component was intensified becoming the most intense texture component. A fiber

along the (3031) [uvtw] component, in the section of φ=0°, was formed since the lowest degree

of deformation and remained present with the same average intensity at all 5 degrees of

deformation. A second fiber along the (2115) [uvtw], in the section of φ =30°, started being

formed after 40% of deformation no significant change in intensity was observed with increase

in cold rolling reduction.

62

(0002) Pole figure (1010) Pole figure

(a) CR20

Contours: 1 2 3

Contours: 0.5 1 1.5 2 2.5

(b) CR40

Contours: 1 2 3 4 5 6 7

Contours: 1 2 3

(c) CR60

Contours: 1 2 3 4 5 6

Contours: 1 2 3

(d) CR80

Contours: 1 2 3 4 5 6

Contours: 1 2 3

(e) CR95

Contours: 1 2 3 4 5 6 7

Contours: 1 2 3

Figure 4.3 – (0002) and (1010) pole figures of the cold rolled samples.

63

θ = 0° θ = 90°

(a) CR20 Contours: 1 2 3 4

ψ = 0°

ψ= 90°

(b) CR40 Contours: 1 2 4 6 8

(c) CR60 Contours: 1 2 4 6 8

(d) CR80 Contours: 1 2 4 6 8

(e) CR95 Contours: 1 2 4 6 8 10

Phi= 0.0

Phi=30.0

Figure 4.4 - ODF sections of φ =0° and φ = 30°, Roe notation, for the samples cold rolled at: a) 20%, b) 40%, c) 60%, d) 80% and e) 95%.

64

The most important features of grain orientation distributions through the entire Euler space can

be clearly illustrated by plotting skeleton lines [Inagaki, 1992]. Skeleton lines are plotted by

connecting points of the maximum orientation density in each constant φ section of the

orientation distribution functions. The skeleton lines obtained on samples cold rolled 20, 40, 60,

80 and 95% are shown in figure 4.5. The shape of the five curves is very similar indicating that

the maximum orientation components were present around the same phi sections, or in other

words, the components have in common the same value of phi (third Euler angle). With

exception for the 20% deformed sample (CR20), which showed a maximum intensity at the

(1013) [3032] texture component, all the other four samples (CR40, 60, 80 and 95) showed the

(1015) [1210] component as the most intense texture component. According to figure 4.5, from

40% to 80% of deformation the intensity of this component tended to have a small decrease from

9.6 to 8.1 times random, and then increasing to 10.6 times random after 95% of cold

deformation.

0

2

4

6

8

10

12

14

0 10 20 30 40 50 60

Phi angle (degrees)

Inte

nsity (

tim

es r

an

do

m)

CR20

CR40

CR60

CR80

CR95

Figure 4.5 - Skeleton lines of the orientation distribution functions for the samples 20%, 40%, 60%, 80% and 95% cold rolled. A significant normal plastic anisotropy can be achieved with the increase of the (0002) fiber

texture, also known as basal type texture. Therefore, the development of this fiber was

65

investigated after each degree of cold and warm rolling reduction and the results are shown in

figures 4.6 and 4.9 for the cold and warm rolled samples, in that order. According to figure 4.6

after 20% of cold deformation the overall intensity of the (0002) fiber texture decreased and it

continued decreasing as the degree of deformation increased to 40 and 60%. At 80% of

deformation the intensity of this fiber started increasing and after 95% of deformation its

intensity was a little higher but still much lower than the intensity found in the material in its

initial condition (before cold rolling). Figure 4.7, which shows the variation in volume fraction

of the (0002) fiber texture as a function of the degree of cold rolling reduction, corroborates what

was just mentioned. The volume fraction was calculated within 10° of the ideal orientation using

the following equation.

(4.1)( ) ⎟⎟

⎞⎜⎜⎝

⎛⎟⎠⎞

⎜⎝⎛ +−⎟

⎠⎞

⎜⎝⎛ −Δ= ∑∑∑

2

Δθθcos2

ΔθθcosψΔφφ,θψ,8100

11 ii

ψ θ φif

0002 Fiber

0

0,5

1

1,5

2

2,5

3

0 20 40 60 80

Psi angle (degrees)

Inte

nsity (

tim

es r

andom

)

AR

CR20

CR40

CR60

CR80

CR95

Figure 4.6 - Development of the 0002//ND fiber texture for the as received (AR) and 20%, 40%, 60%, 80% and 95% cold rolled (CR) samples.

66

0002 Fiber

0

0,005

0,01

0,015

0,02

0,025

0 20 40 60 80

Cold Rolling Reduction (%)

Volu

me F

raction

Figure 4.7 - Variation in volume fraction of the 0002//ND fiber texture with degree of cold rolling reduction. The as received material corresponds to the 0% cold rolling reduction. 4.1.3 - Warm Rolled Samples

The pole figures presented in figures 4.8 (a) to (e) show the evolution of the (0002) and

(1010) pole figures after five different rolling reductions at 260°C. At this temperature of

deformation, two (0002) poles are formed towards the RD and these two poles are intensified as

the degree of deformation increases varying from 4 times random after 20% of deformation to 8

times random after 80% of deformation. From the observation of figures 4.8 (a) to (d) one can

see a spreading in the distribution of the (0002) poles towards the transversal direction. After

95% of deformation, the spreading increases and the two poles towards RD that were present in

the previous pole figures no long exist. A couple of poles, this time, towards the TD were formed

after 95% of deformation changing completely the aspect of the (0002) pole figure when

comparing to the pole figure results of the other four warm rolled samples.

The orientation distribution function for the samples rolled at 260°C, illustrated in figure

4.9, shows that the warm rolling has a great effect on the texture formation of titanium sheets. A

formation of a fiber type of texture in the section of φ =0° at θ=20°, can be observed in the WR20

sample where the maximum intensity is at the (1015) [5052] texture component.

67

(0002) Pole figure (1010) Pole figure

(a) WR20

Contours: 1 2 3 4

Contours: 0.5 1 1.5 2

(b) WR40

Contours: 1 2 3 4 5 6

Contours: 0.5 1 1.5 2

(c) WR60

Contours: 1 2 3 4 5 6

Contours: 0.5 1 1.5 2

(d) WR80

Contours: 1 2 4 6 8

Contours: 1 2 3 4

(e) WR95

Contours: 1 2 4 6 8

Contours: 1 2 3 4 5

Figure 4.8 – (0002) and (1010) pole figures of the warm rolled samples

68

θ = 0° θ = 90°

(a) WR20 Contours: 1 2 3 4

ψ = 0°

ψ= 90°

(b) WR40 Contours: 1 2 4 6 8

(c) WR60 Contours: 1 2 3 4 5 6

(d) WR80 Contours: 1 2 4 6 8

(e) WR95 Contours: 1 2 4 6 8 10

Phi= 0.0 Phi=30.0 Figure 4.9 - ODF sections of φ =0° and φ = 30°, Roe notation, for the samples warm rolled at: a) 20%, b) 40%, c) 60%, d) 80% and e) 95%.

69

As the degree of deformation increases this fiber tends to shift towards the (0002) plane

with maximum components 5° away from the (1015)[5052]. After 80% of deformation the

aspect of the fiber starts changing and now the maximum texture component is the (1015)

[1210], with intensity of around 9.0 and 12.0 times random for the samples WR80 and WR95,

respectively. Still in ODF sections of φ =0°, a weak fiber texture around the (2021) plane with

average value of less than 2 times random was formed and it was not influenced by the degree of

deformation. After 40% of deformation a fiber along the (2115) plane, in the section of φ = 30°,

is formed and it is destroyed as the degree of warm rolling increases.

0

2

4

6

8

10

12

14

0 10 20 30 40 50 60

Phi angle (degrees)

Inte

nsity (

tim

es r

an

do

m)

WR20

WR40

WR60

WR80

WR95

Figure 4.10 - Skeleton lines of the orientation distribution functions for the samples 20%, 40%, 60%, 80% and 95% warm rolled.

Figure 4.10 shows the skeleton lines for the 20%, 40%, 60%, 80% and 95% warm rolled

samples. According to this figure, it is possible to observe that for the samples deformed from 20

to 80% the texture components that presented maximum intensity, at each of the constant

sections of the ODFs, had their intensities around the same value, making the skeleton lines of

these referred samples to be rather constant throughout the seven phi sections (φ = 0° to φ = 60°).

After 95% of warm deformation in contrast, the skeleton line of the sample assumed a behavior

70

similar to the skeleton lines of the cold rolled samples (see figure 4.5) where a maximum

intensity at φ = 0° and a minimum intensity at φ = 35° is clearly seen.

The behavior of the (0002) fiber present in the as received material and resultant after

rolling at 260°C is presented in figure 4.11, which shows that the warm rolling tends, in general,

to intensify the fiber as the percentage of deformation increases from 20% to 80%. Figure 4.12

shows the volume fraction of the (0002) fiber calculated within 10° of the ideal orientation. As

also observed from figure 4.12, after 60% of deformation (sample WR60) there is a small drop in

the intensity of the fiber, which increases again after 80% and then decreases after 95% of

deformation (WR95).

0002 Fiber

0

1

2

3

4

5

6

7

8

9

0 20 40 60 80

Psi angle (degrees)

Inte

nsity (

tim

es r

andom

)

AR

WR20

WR40

WR60

WR80

WR95

Figure 4.11 - Development of the 0002//ND fiber texture for the as received (AR) and 20%, 40%, 60%, 80% and 95% warm rolled (WR) samples.

71

0002 Fiber

0

0,02

0,04

0,06

0,08

0 20 40 60 80

Warm Rolling Reduction (%)

Volu

me F

raction

Figure 4.12 - Variation in volume fraction of the (0002)//ND fiber texture with degree of warm rolling reduction. The as received material corresponds to the 0% cold rolling reduction. 4.2 - X-ray Peak Profile Analysis

In this section, it is presented the evolution of the dislocation densities and types obtained

from the X-ray peak profile analysis in commercially pure alpha-titanium deformed by rolling at

260o C for the following rolling reduction levels: 40%, 60% and 80%. The dislocation densities

and the average contrast factor of dislocation for each sample were determined by using Multiple

Whole Profile fitting procedure (MWP) [Ribárik, Ungár and Gubicza, 2001]. In this evaluation,

it is assumed that the peak broadening is caused by the smallness of the coherently scattering

domains and by strain effect arising from the presence of dislocations. As it has been shown in

[Ribárik, Ungár and Gubicza, 2001] for case of hexagonal crystals, the MWP fitting procedure

enables the evaluation of six microstructure parameters:

(a) The dislocation density and arrangement parameter, ρ and M. By definition [Wilkens, 1970],

M is the dislocation arrangement parameter in the Wilkens function, and the value of M gives the

strength of the dipole character of dislocations, i.e., the higher the value of M, the weaker the

dipole character and the screening of the displacement fields of dislocations [Wilkens, 1970],

(b) The median and the variance, m and σ, in the size part of the profiles,

(c) q1 and q2 the parameters of the dislocation contrast factors, as shown in equation (2.4).

Details of the MWP method can be found in [Ribárik, Ungár and Gubicza, 2001]. In the present

study the effect of crystallite size was found to tend to infinity. This is due to the fact that the

72

coherent domain is greater than few μm and in this case the effect of size broadening is

insignificant compared with the distortion effect.

As shown in Chapter 2 the presence of crystallographic texture complicates the

interpretation of peak broadening observations. In an attempt to emulate a random

polycrystalline specimen the strongest diffraction peaks from the three different faces of the

orthogonal sample were mixed to form a full diffraction pattern and used in the MWP evaluation

[Gubicza et al., 2003; Glavicic et al., 2004]. These mixed patterns were then used in the MWP

evaluation. The results for the dislocation densities and the arrangement parameter, M, for the

titanium specimens deformed at 40%, 60% and 80% reduction levels are listed in table 4.1. The

errors of the dislocations densities listed in this table are calculated from the covariance matrix in

the refinement.

Table 4.1 - Dislocation densities and arrangement parameter, M, obtained from MWP evaluation for Ti samples deformed at different reduction levels. Rolling Reduction [%] 40% reduction 60% reduction 80% reduction

M 2.9 1.7 1.1

ρ[1/cm2] 5x1010 ± 5% 8x1010 ± 10% 10x1010 ± 7%

The hi fractions of dislocation density with the three Burgers vector types, <a>, <c> and <c+a>,

are shown as a function of deformation level in figure 4.13. The results show that:

I) in the specimen deformed at 40% the <c+a> and <a> dislocations type are

dominating the dislocation population present in the sample;

II) at higher deformation levels <a> type is the dominant dislocation type and the

presence of <c+a> dislocation decrease;

III) at all deformation levels the fraction of <c> dislocations type is marginal.

73

30 40 50 60 70 80

20

40

60

80 a type

c type

c+a typeB

urg

ers

Ve

cto

r P

op

ula

tio

n [

%]

rolling reduction [%]

Figure 4.13 - The hi fractions of the three fundamental Burgers vector types, <a>, <c> and <c+a>, as a function of rolling reduction. Note that in the figure the solutions to equations. (2.8, 2.9 and 2.10), the hi fractions, were transformed in percentages.

Figures 4.14 and 4.15 illustrate the (0002) and (1120) Bragg reflections profiles for all

deformations levels studied here. The difference in broadening is due to the difference in the

dislocations types accumulated in each sample. From figure 4.14 it can be observed that the

broadening of the (0002) reflections is decreasing as the deformation increases. However, figure

4.15 shows that the breadth of (1120) reflections increases as the deformation increases. The two

figures are an illustration of the different contrast effect due to different types of dislocations and

are an indication that the content of dislocations with <a> Burgers vector type increases as the

deformations process increases to higher levels.

74

-0.08 -0.04 0.00 0.04 0.08

0.0

0.2

0.4

0.6

0.8

1.0 (0002) reflection

80% deformation60% deformation

No

rma

lize

d I

nte

nsity

K [1/nm]

40% deformation

Figure 4.14 - The line profiles of (0002) Bragg reflections for different deformations levels. On the x-axes K is given by K=2sinθ/λ, where θ is the Bragg angle and λ is the wavelength of the used radiation.

In accordance to the dislocations slip directions, <a>, <c> and <c+a>, (0001), (2110)

and (2113) pole figures were measured for each titanium sample studied here and the results are

being illustrated in figure 4.16 (a) to (d). The as received material, which exhibits a weak texture,

is also being shown in this figure in order to compare the resultant pole figures before and after

warm rolling deformation. Figure 4.16 (a) shows that the <0001> directions of most of the

crystals are distributed on the plane of ND (normal direction)-RD (rolling direction). When the

sample is rolled to a strain equivalent to 40% of reduction, <0001> directions of the crystals

move close to ND as can be seen in figure 4.16 (b), a trend also observed in figure 4.16 (c) for

the sample rolled to a strain of 60%. In figure 4.16 (d) the split from the ND direction becomes

smaller at a higher strain of 80%. Comparing with <0001>, the evolution of <2110> is not so

complicated. The initial distribution of <2110> is random. At a strain of 40%, the <2110>

direction becomes oriented along the plane of RD-TD. At higher deformation rates the <2110>

75

direction becomes concentrated close to RD, but 5° away. This corresponds to the split shown in

the (0001) pole figures. The (0001) pole figures and (2110) pole figures verify that the <a> type

of slip is dominant during the rolling process. The (2113) pole figures in figure 4.16 show that

the number of crystals with <2113> parallel to RD decreases with the increase of rolling strain.

This is further demonstrated in figure 4.17, which shows the evolution pattern of the distribution

density of crystals whose <0001>, <2110> and <2113> are parallel to the RD.

-0.08 -0.04 0.00 0.04 0.08

0.0

0.2

0.4

0.6

0.8

1.0(11-20) reflection

40% deformation

60% deformation 80% deformation

K [1/nm]

No

rma

lize

d I

nte

nsity

Figure 4.15 - The line profiles of (1120) Bragg reflections for different deformations levels. On the x-axes K is given by K=2sinθ/λ, where θ is the Bragg angle and λ is the wave length of the used radiation.

76

(0001) (2110) (2113)

TD

ND

RD (a)

(b)

(c)

(d)

Figure 4.16 - (2110), (0001) and (2113) pole figures of alpha titanium at a rolling reduction of (a) 0%, (b) 40% (c) 60%, (d) 80%, respectively.

77

It can be observed from figure 4.17 that the distribution density of the RD//<2110> increases

with the strain, while the RD//<2113> decreases and the RD//<0001> is always close to zero.

40 60 80

0.0

0.5

1.0

1.5

rolling reduction (%)

Inte

nsity (

tim

es r

an

do

m)

RD//0001

RD//-2113

RD//-2110

Figure 4.17 - Evolution of intensities of components with RD//2110, RD//0001 and RD//2113, respectively, during rolling reduction. 4.3- Texture Simulation

The simulation of the deformation texture is the closure of the methodology proposed in

this work. In this step the evolution of texture was predicted by means of a self-consistent

approach for the large deformation polycrystal viscoplasticity proposed by Molinari, Canova and

Ahzi [Molinari et al., 1987]. The initial texture, corresponding to the As-received material,

consisted of a file with a set of grains orientations extracted from the SOD file (sample

orientation distribution) obtained from the X-ray data for pole figure measurement using the

routine PoPla (Preferred orientation package Los Alamos). Figure 4.18 shows the (0002) pole

figures of measured (a) and discrete grains file (b) for the as-received sample. The sample axes

78

convention adopted here is rolling direction (RD) parallel to the vertical direction, transversal

direction (TD) parallel to the horizontal direction and normal direction (ND), or normal to the

rolling plane, perpendicular to the plane of the paper (or parallel to a vector outwards the paper

sheet). The polycrystalline aggregate represented by the discrete set of orientations consists of

166 grains representing the initial texture.

(a) (b)

Figure 4.18 – (0002) pole figures for the as-received material: (a) experimental and (b) discrete grains file. Axes convention: RD in the vertical direction and TD in the horizontal direction.

The slip systems reported to be activated in different hexagonal crystals ([Rosi et al.

1956; Grooves and Kelly, 1963; Chin, 1975; Conrad, 1981; Phillipe et all, 1995, Fundenberger et

all, 1997; Salem 2002; Glavicic et al. 2003; Zaefferer, 2003; Kalidindi et al. 2004], are of two

types of slip systems: slip systems with <a> Burgers vector (basal, prismatic and pyramidal) and

slip systems with <c+a> Burgers vector (pyramidal I and II). Since <a> Burgers vector lie in the

basal plane of the hexagonal crystal, glide systems with <a> vector like basal and prismatic

together, comprise only four independent systems. For the same reason, addition of pyramidal

<a> set of slip systems will not supply the missing degree of freedom which is responsible for

accommodate deformations in c-axis direction. For that, additional mechanisms such as twinning

and glide in <c+a> direction are required [Tomé and Kocks, 1985].

79

Here twinning is not being taken into account and it is assumed that the major

contribution to accommodate deformation in c-axis direction comes from activation of <c+a>

slip systems. The other slip systems considered for this simulation where basal and prismatic

with <a> Burgers vector. The simulation of cold rolling of CP-Titanium (alpha HCP) was

conducted assuming a ratio between the resolved shear stress for <a> prismatic slip and <a>

basal slip τb/τpr=3 with τp=τ0 ,i.e., τpr is normalized to 1. The resistance for <c+a> slip on

pyramidal planes is about one order of magnitude higher than that of prismatic or basal slip. Here

it was taken as τpyr<c+a>/τpr=10. The glide systems considered for the cold rolling simulation

were then: 1010 <1120> prismatic, 0001 <1120> basal and 1011<1123> pyramidal

<c+a> slip. The effect of hardening is taken into account and is assumed to be linear during the

cold rolling. The imposed macro strain corresponds to a maximum deformation of 95% (or a true

strain of -3.0), applied in steps of 5% and the results were recorded at 20%, 40%, 60% 80% and

95% of reduction in thickness for comparison with the experimental results. The strain rate

consistent with flat rolling is of the order of 1 s-1[Tricot, 1992]. The results of cold rolling

simulation are presented in form of the (0002) pole figures for each of the five degrees of

deformation considered in this work and are shown in figure 4.19. The discussion of these results

is presented in the next chapter.

In the simulation of warm rolling, for medium and high homologous temperatures, the

resistances to basal and prismatic slip are approximately equal. The glide systems considered for

the warm rolling simulation are: 1010 <1120> prismatic, 0001 <1120>, 1122<1123> and

1011<1123> pyramidal <c+a> slip. Now the resolved shear stress for prismatic slip is taken

τpr=0.6 and that for basal slip τb=1.0 reflecting the drop in the flow stress due to the thermal

effect [Levine, 1966]. The resistances for 1122<1123> and 1011<1123> <c+a> slip were

taken as τpyr<c+a>=7.0. The results of line broadening analysis suggest that during the actual

warm rolling process, as a consequence of reheating between passes, recovery was a possible

mechanism affecting the deformation texture evolution. For this reason it was decided to

eliminate the hardening effect from the simulation. Other parameters like strain rate and macro

strain were kept the same. The results are displayed in the figure 4.20.

80

Sample (0002) Experimental (0002) Simulated

(a) CR20

(b) CR40

(c) CR60

(d) CR80

(e) CR95

Figure 4.19 – Experimental and simulated results of the (0002) pole figures for the cold rolled samples deformed (a) 20%, (b) 40%, (c) 60%, (d) 80% and (e) 95%.

81

Sample (0002) Experimental (0002) Simulated

(a) WR20

(b) WR40

(c) WR60

(d) WR80

(e) WR95

Figure 4.20 – Experimental and simulated results of the (0002) pole figures for the warm rolled samples deformed (a) 20%, (b) 40%, (c) 60%, (d) 80% and (e) 95%.

82

CHAPTER 5

DISCUSSION

The texture evolution of cold and warm rolled commercially pure titanium was investigated.

Peak profile analysis of the warm rolled samples deformed at 40%, 60% and 80% was conducted

and simulation of the deformation texture for the samples cold and warm rolled at 20%, 40%,

60%, 80% and 95% was carried out in order to demonstrate the feasibility of the methodology

proposed in this work, which is an unified path model for characterization and also prediction of

microstructure evolution, in terms of texture, in materials that have undergone thermo-

mechanical processing. The results will be reviewed and discussed in this section.

5.1 - Deformation Texture

The texture results for the cold rolled samples shown in figures 4.3 to 4.7 indicate that the

intensity of the (0002) poles almost doubles when the cold reduction increases from 20 to 40%

remaining around the same until the maximum applied cold reduction, 95%. Another effect of

the cold deformation in the distribution of (0002) planes is seen when the degree of reduction

increases to 40% and above. The distribution density of the basal poles towards the rolling

direction decreases and a concentration of these poles tilted ± 20°away from the normal direction

and towards the transversal direction increases with increasing in the degree of cold deformation.

The volume fraction of the basal fiber, (0002)//RD, decreases as the cold reduction increases up

to 60% of deformation. At 80% of deformation the (0002) fiber starts increasing and after 95%

of deformation the sample exhibits volume fraction of (0002) planes, higher than the fraction

observed in the sample 20% cold deformed but still lower than the annealed as received sample.

According to the (0002) pole figures and the profiles of the basal fiber (figures 4.3 and 4.6,

83

respectively), for degrees of cold reduction above 40% the main effect of deformation, rather

than intensification of the basal pole, is the change in the distribution of this pole with regarding

to the normal and the transversal direction of the samples. The main concentration of (0002)

poles tilted ± 20°away from the normal direction and towards the transversal direction of the

sample is very characteristic of cold rolled hexagonal materials with c/a ratios less than the ideal

value of 1.633 as it is the case of titanium [Wang and Huang, 2003]. Keeler and Geisler [Keeler

and Geisler, 1956] working with high purity titanium deformed 99.7% found a tilt of the

hexagonal unit cells in such a manner that the basal planes were rotated 50 ± 10° out of the

rolling plane about the rolling direction. This amount of tilt is much greater than that reported in

this current investigation as well as by earlier investigators [McHargue and Hammond, 1953 and

McGeary and Lustman, 1951]. The reason for that, rather than the high degree of deformation

can be the thermo-mechanical processing and chemical composition of the material.

According to the (1010) pole figure results, the degree of cold rolling reduction do not

affect the intensity of these planes but it plays a role in the distribution of them regarding to the

main sample’s directions (ND, RD and TD). At 20% of cold deformation, the main concentration

of (1010) poles was around the transversal direction of the sample but as the degree of

deformation increased to 40% until 95%, an alignment of the (1010) poles with the rolling

direction started to happen. These results are in agreement with previews results from the

literature [Wang and Huang, 2003].

According to orientation distribution function, figure 4.4, the general aspect of the

crystallographic orientation distribution after cold rolling changed dramatically when compared

to the as received sample. A small percentage of cold reduction, 20%, was able to result in

relative intense texture components, the (1013) [3032] and (1015) [1210], about 4 times random.

As the degree of cold rolling reduction increased, the (1013) [3032] component disappeared and

the (1015) [1210] texture component still presented was intensified after 40% of deformation

reaching about 8 times random, twice as much as its intensity in the CR20 sample. This texture

component remained at this strength until 95% of deformation when its intensity was increased

to 10 times random. The (1015) [1210] component was the most intense texture component in all

four samples that were cold rolled above 40% of deformation. Brandao [Brandao, 1993] found a

close final texture result, (1014) [1210] component, after cold rolling of zircaloy-4 at different

degrees of reduction. The resultant texture after cold rolling has shown to be very dependent

84

from the texture present in the starting material. Inagaki [Inagaki, 1991], working with a highly

textured pure titanium ((1013) [1210] component about 8.3 times random) cold rolled from 15%

to 90%, found different maximum texture components depending on the degree of deformation,

being the (2115) [0110] the stable end orientation.

The distribution of (0002) poles after warm rolling was quite different from the

distribution found after cold rolling. After deformation at 260°C, two (0002) poles are formed

towards the rolling direction with a spreading in the distribution towards the transversal

direction. The two poles (towards the RD) are intensified as the degree of deformation increases

varying from 4 times random after 20% of deformation to 8 times random after 80% of

deformation. After 95% of deformation, the spreading increases and the two poles towards RD

that were present in the samples deformed up to 80% no long exist. A couple of poles, this time,

towards the transversal direction were formed after 95% of deformation changing completely the

aspect of the (0002) pole figure when comparing to the pole figure results of the other four warm

rolled samples. At this point, the (0002) pole figure of sample WR95 becomes similar to the pole

figures for the samples cold rolled, excepting by the concentration density around the normal

direction, which is present solely in the cold rolled samples. The same change in the distribution

of the (0002) poles with respect to the rolling and transversal direction was observed in titanium

specimens hot rolled 75% and 94% at 700°C [Inagaki, 1990].

The orientation distribution function, illustrated in figure 4.9, shows that the warm rolling

has a great effect on the texture formation of titanium sheets. A formation of a fiber type of

texture in the section of φ =0° at θ=20°, can be observed in the WR20 sample where the

maximum intensity is at the (1015) [5052] texture component (about 4 times random). As the

degree of deformation increases this fiber tends to shift towards the (0002) plane with maximum

components (about 8 times random) 5° away from the (1015) [5052]. After 80% of deformation

the texture component that shows maximum intensity is the (1015) [1210] component, which is

the same component present after cold rolling but more intense, around 9.0 and 12.0 times

random for the samples WR80 and WR95, respectively.

The differences in the texture formation after warm and cold rolling can also be seen in

the results of the skeleton lines for both modes of deformation and in the results of the volume

fraction of the (0002) planes parallel to the normal direction, which show that the warm rolling is

more effective in developing the basal type of fiber than the cold rolling.

85

5.2 - X-Ray Peak Profile Analysis

The results for the dislocation densities and the arrangement parameter, M, for the

titanium specimens warm rolled at 40%, 60% and 80% of deformation, listed in table 4.1,.shows

that the small increment of ρ (dislocation densities) is in accordance with a dynamic recovery

effect, where dislocation annihilation may occur due to the high deformation temperature. The

values of the arrangement parameter M in table 4.1 indicate that the dislocations in the titanium

samples studied here exhibit a week dipole character.

According to figure 4.13, which shows the hi fractions of dislocation density with the

three Burgers vector types, <c>, <a> and <c+a>, it can be noticed that at all degrees of warm

deformation the fraction of the <c> type of dislocation was not significant. The <a> type of

dislocation, on the other hand, was prevalent in all samples and the <c+a> type, although

dominant in sample WR40, decreased its population at higher deformation levels. The results of

the (0001) and (2110) pole figures, shown in figure 4.16, corroborates that the <a> type of slip

is dominant during the rolling process.

The activity of <c+a> dislocations plays an important role in dynamic recovery, as screw

dislocations of <c+a> type can move to the next slip planes by double cross slip followed by

dislocation annihilation. Dislocations with b=<0001> Burgers vector, <c> type, are sessile, thus

its presence can be seen practically unchanged in the deformation range studied here [Jones and

Hutchinson, 1982; Song and Gray, 1995]. The results presented here are in good agreement with

previous extensive TEM studies [Jones and Hutchinson, 1982; Song and Gray, 1995], which

show that the <a> dislocations are the most frequent type of dislocations observed in deformed

titanium and that the <c+a> and <c> types, also reported, are less numerous. The differences in

the broadening of the (0002) and the (1120) reflections (figures 4.14 and 4.15) caused by the

different contrast effect due to different types of dislocations are another indication that the

content of dislocations with <a> Burgers vector type increases as the percentage of deformation

increases to higher levels. The results of the population density of the <c>, <a> and <c+a> types

of dislocation were confirmed through the evaluation of the evolution pattern of the distribution

density of crystals whose <0001>, <2110> and <2113> are parallel to the RD (figure 4.17). From

this figure it can be observed that the <2110>//RD increases with the strain, while the

<2113>//RD decreases and the <0001>//RD is always close to zero. This is in good correlation

with the Burgers vector population results, which show that the density of <a> type of

86

dislocations increases, while the density of <c+a> type of dislocations decreases and <c> type

remains at very low values during the whole rolling process.

5.3 - Self Consistent Simulation of the Deformation Texture

The application of self-consistent approaches for the prediction of cold rolling

deformation textures have been the subject of several investigations [Tomé et al.,1991, Nemat-

Nasser et al., 1999; Schoenfeld et al.,1994; Kalidindi, 1997; Philippe et al., 1995; Funderberger

et al., 1997; Glavicic, 2003]. The most common modes of deformation considered in the

literature for cold working deformation processes are 1010 <1120> prismatic, 0001 <1120>

basal, 1011<1123> pyramidal <c+a> slip and twinning modes. In this work twinning is not

considered and therefore the accommodation of deformation imposed in the c-axis direction is

assumed to entirely done by pyramidal slip with <c+a> Burgers vector. The results shown in

figure 4.19, exhibit a reasonable agreement between experimental pole figures and the respective

predicted ones with small discrepancies on the shape for the higher degrees of deformation

which may be explained by the absence of twinning in the model employed in this work. The

well defined rolling texture of high deformed cold worked hcp metals with c/a<1.663 which is

characteristic of Titanium, a t-type rolling texture presenting a well formed spread in the RD-TD

plane with (0002) poles rotated 35° to 40° from ND towards TD, is explained by the operation of

a combination of prismatic slip and twinning modes. The operation of prismatic and pyramidal

<c+a> slip systems only will tend to push the (0002) poles to form a fiber texture in the outer

rings of the pole figure (for the high deformation case) and the operation of twinning

concomitant with those will tend to stabilize the poles in the described end-stable position for the

cold rolling component of texture. In the other hand if the content of oxygen is high, basal slip

maybe favored and its activation tends to drag the (0002) poles closer to ND as it is the case of

the (0002) pole figures shown in the figure 4.19 (d) and (e).

The TD with angles 15° to 30°, similar to those observed in the (0002) pole figure for the

CR95 sample (figure 4.19(e)). The discrepancies in shape observed in the simulated pole figure

for 95% with a less defined spread towards the rolling direction may be explained by the absence

of twinning in the model. The choice of ratios for the resolved shear stresses of the different

87

systems was based on the findings of previous authors [Tomé et al.,1991, Nemat-Nasser et al.,

1999; Schoenfeld et al.,1994; Kalidindi, 1997; Philippe et al., 1995; Funderberger et al., 1997;

Glavicic, 2003].

In the case of warm rolling the considered slip systems are 1010 <1120> prismatic,

0001 <1120> basal, 1122 <1123> and 1011<1123> pyramidal <c+a> slip. The principal

difference in the approach analysis between warm and cold rolling is the weak contribution of

twinning for the deformation process. At 260°C and higher temperatures up to 700°C twinning is

far less active if not almost completely suppressed in Ti and Zr. and the deformation in the c-axis

direction is accommodated by pyramidal slip with <c+a> Burgers vector in the direction <1123>.

The drop in twinning activity (figure 5.1 ) was also observed by Glavicic [ Glavicic et al., 2004].

Working with the same material, for the range of 20C° to 300C° and 10% reduction, their

observations indicated that the main slip activity over the temperature range investigated was

prism <a> and pyramidal <c+a> type slip.

Figure 5. 1 - Variation of: (a) twin volume fraction; (b) strain accommodated by twinning as a function of rolling temperature.

88

Another evidence that supports the conclusion of low activity of twinning comes from the

observation of optical micrographs of warm rolled samples (figure 5.2).

a) WR80, 200x b) WR95, 200x

Figure 5.2 – Optical micrographs of warm rolled 80% and 95% reduction.

100 μ 100 μ

The results from line-broadening analysis however, have shown a different trend with the

density of <c+a> decreasing with increasing applied strain (figure 4.13). This can be explained

by the high content of oxygen present in the material, situation in which the activation of <c+a>

slip is affected and the fact that at moderate to high temperatures the drop in the resolved shear

stress for both prismatic and basal slip with the ratio between then almost being equal to 1,

favoring a high activity of basal slip 0001 <1120>. This slip activity in the basal plane is also

claimed to be responsible for the rolling texture components being closer to ND, a t-type rolling

texture with (0002) poles rotated 15° to 20° from ND towards TD as it can be observed in the

figure 4.20. The results of simulated deformation texture of warm rolled CP-Ti are in reasonably

good agreement with the experimental results and with the peak profile analysis findings.

89

CHAPTER 6

SUMMARY AND FUTURE WORK

6.1 - Summary

The present work attempts to establish a unified path model for characterization as well

as a prediction of microstructure evolution, in terms of texture, in commercially pure titanium

that have undergone thermo-mechanical processing. Cold and warm at five different degrees of

deformation, 20%, 40%, 60%, 80% and 95% were used in this investigation

The texture evolution of the cold and warm rolled commercially pure titanium samples

was carefully measured. Peak profile analysis of the warm rolled samples deformed at 40%, 60%

and 80% was conducted and simulation of the deformation texture for the samples cold and

warm rolled at 20%, 40%, 60%, 80% and 95% was carried out in order to demonstrate the

feasibility of the methodology proposed in this work, which is an unified path model for

characterization and also prediction of microstructure evolution, in terms of texture, in materials

that have undergone thermo-mechanical processing.

The experimental texture results show that excepting for the samples 95% deformed, the

warm rolling has shown to develop a deformed texture different from the cold rolling.

The results of peak profile analysis carried out for the 40%, 60% and 80% warm rolled

samples show that the <a> type of dislocation was prevalent in all samples while the <c> type of

dislocation was only marginal. The X-ray peak profile analysis, based on the dislocation model

of anisotropic peak broadening, show the dislocation densities, distributions and type during the

rolling processing in good agreement with the texture evolution.

Even though twining was not taken into account during simulation of the cold rolled

samples, there is a reasonable agreement between the experimental and the predicted pole figures

with a small divergence on the distribution of in the TD-RD plane for the higher deformed

samples.

90

The results of simulated deformation texture of warm rolled CP-Ti are in good agreement

with the experimental results and with the peak profile analysis findings.

The principal problem and possibly weakness in this work was the fact that in proceeding

with warm or hot rolling, the ideal situation is the one where it is possible to achieve the desired

reduction for each step of the investigation at once or at least in a continuous way, i.e., to achieve

the final reduction in sequential passes if necessary but without the need of intermediary

reheating steps which can cause recovery or even recrystallization what is not desired here since

it is the intention to isolate the mechanisms that where active trough the whole process for each

one of the established regimes of thermomechanical process. This is only possible if the facility

available for the experiment has the characteristics of an industrial continuous rolling mill with

automatic control for feeding, step sizes and rolling temperature.

Another problem emerges from the fact that, in its actual state of the art condition, the

line broadening analysis is not suited for extracting data for individual slip systems activity, in

the case of hexagonal close packed materials, and so far the method allows obtaining this sort of

data for families of sub-slip systems.

6.2 – Future Work

The following are some suggestions for future work based on the findings, conclusions

and problems identified on the course of the present work.

1- In the same line of investigation conducted in this work, it is suggested to extend the line

broadening analysis to higher deformation temperatures like 535°C and 815°C to investigate

what mechanisms are likely to operate in order to accommodate the deformation imposed. This

is a difficult task since dynamic as well as static recrystalization can take place at temperatures as

low as 400°C, for the material investigated here, and it will introduce another level of complexity

for the line broadening analysis.2- To extend and complement the actual work with a SEM/OIM

investigation in order to study the mechanisms of twinning that were possibly present but

disregarded in this investigation and also to study the grain boundary character distribution of the

microstructure.

91

3- To incorporate twinning to the actual self-consistent approach. The model for simulation of

deformation texture used in this work was not developed in such way that twinning was taken

into account as a possible mechanism of de formation. In fact, since twinning is not a major

system for the case in study, this self-consistent scheme has proven to render consistent

predictions in both HCP and Cubic materials. However the complete absence of twinning leads

to small discrepancies like those observed in the simulation results for the heavily warm and cold

rolled samples (figure 4.19(e) and 4.20(e)) where the spread observed in the experimental pole

figures was not present in the simulated pole figures.

4- To use better homogenization theories

5- To extend the calculation of the contrast factors to individual slip systems. In the actual stage,

the line broadening approach is capable to predict the activity of individual slip systems in the

case of materials with cubic structure. The extension of the method to predict the activity of

individual slip systems in the case of materials with hexagonal structure would be for sure a great

breakthrough towards better tools and paths for prediction and control of deformation texture of

this important class of materials.

6 - Extension of the self-consistent formulation to the case of bi-modal structure as in the case of

(α+β) titanium alloys.7- To extent the line broadening analysis to (α+β) titanium alloys taking

into account the individual systems activated in the case of both structures and their interaction

as is the case of transformation textures and variant selection problems.

92

REFFERENCES

Ahzi, S. (1987), A Self-Consistent Approach for the Viscoplastic Behavior of Polycrystals Under Large Deformations, PH. D Thesis, Universite de Metz (France). Asaro R.J., Needleman A. (1985), Acta Metallurgica, vol. 33, p. 923. Berveiller M., Zaoui A. (1980), Proceedings of the 15th Colloquium GFR, Paris, p. 175. Bishop J.F.W., Hill R. (1951), Phil. Mag., vol. 42, p. 1298. Bomberger H.B., Froes F.H., Morton P.H. (1985), Titanium technology: present status and future trends, Titanium Development Association, p. 3. Borbely A., Driver J.H, UNGÁR T. (2000), Acta Materialia, vol. 48, p. 2005. Boyer R., Welsch G., Collings E. W. (1994), Materials properties handbook: Titanium alloys, ASM International, Materials Park, OH, p. 311. Brandao L.P.M. (1993), Textura e Propriedades Mecanicas da Liga Zircaloy-4, Doctorate Dissertation, Rio de Janeiro, Brazil. Bunge H. J. (1970), Kristall und Technik, vol. 5, p. 145. Bunge H.J. (1982), Texture Analysis in Materials Science, Butterworth, London. Bunge H.J. (1987), International Metals Review, vol. 32, p.265. Caglioti G., Paoletti A., Ricii F. P. (1958), Nucl. Instrum. Methods, vol. 3, p. 223. Canova G. R., Molinari A., Ahzi S., Adams B. L. (1989), J. Mechanics and Physics of Solids Castelnau O, Francillette H., Bacroix B., Ledensohn R.A. (2001), Journal of Nuclear Materials, vol. 279, p. 14. Cheary R. W., Dooryhee E., Lynch P., Armstrong N., Dligatch S. (2000), Journal of Applied Crystallography, vol. 33, p. 1271.

93

Chin G.Y. (1975), Metallurgical Transactions A, vol. 6, p. 238. Chung and Buessem (1968), Anisotropy in Single Crystal Refractory Compound, edited by F. W. Vahldiek and S. A. Mersol, Plenum, NewYork, vol. 2. Collings E.W. (1994), The Physical Metallurgy of titanium Alloys, ASM, Metals Park, Oh. Conrad H. (1981), Prog. Materials Science, vol. 26, p. 123. Crosby K.E., Mirshams R.A., Pang S.S. (2000), Journal of Materials Science, vol.35, p. 3189. Cullity B.D. (2001), Elements of X-Ray Diffraction, Prentice-Hall, New Jersey. Davies G.S., Goodwill D.J., Kallend J.S. (1971), Journal of Applied Crystallography, vol. 4, p. 193. Dieter G.E. (1986), Mechanical Metallurgy, McGraw-Hill, New York. Dragomir I. C., Li D.S., Castello-Branco G.A., Garmestani H., Snyder R.L., Ribarik G., Ungar T. (2005), Journal of Materials Characterization – In Press.

Dragomir I.C., Ungar T. (2002), Journal of Applied Crystallography, vol. 35B, p. 556. Dragomir I.C., Ungar T., (2002), Powder Diffraction Journal, vol. 17A, p. 104. Dragomir-Cernatescu I., Li D.S., Castello-Branco G.A., Garmestani H., Snyder R.L. (2005), Scripta Materiallia – In Press.

Fjeldly A., Roven H.J. (1996), Acta Materiallia, vol. 44, p. 3497. Froes F. H. (1990), JOM, vol. 42-3, p.16. Fundenberger J.J., Phillippe M.J., Wagner F., Esling C. (1997), Acta Materiallia, vol. 45, No. 10, p. 4041. Glavicic M.G., Salem A.A., Semiatin S.L. (2004), Acta Materialia, vol. 52, p. 647. Groma I., Ungár T., Wilkens S.M. (1988), Journal of Applied Crystallography, vol. 21, p. 47. Groves G.W., Kelly A. (1963), Philosophic Magazine, vol.8, p. 877. Gubicza J., Dragomir I.C., Ribárik G., Baik S.C., Zhu Y.T., Valiev R.Z. and Ungár T., (2003), Zeitschrift für Metallkunde, vol. 94, p. 1185. Gubicza J., Szepvölgyi J., Mohai I., Zsoldos L., Ungar T. (2000), Materials Science and Engineering A, vol. 280, p. 263.

94

Guillaume M.J., Beauvais C., Hocheid B. (1981), Proceedings of ICOTOM 6, Tokyo, p.975. Hill, R. (1965), Journal of Mechanics of Physics and Solids, vol. 13, p. 89. Hill, R. (1967), Journal of Mechanics of Physics and Solids, vol. 15, p. 79. Hirosuke I. (1992), Zeitshchrift fur Metallkunde, vol. 83-1, p. 40. Honeycombe R.W.K. (1984), The Plastic Deformation of Metals”, Edward Arnold Publ. Ltd., London, p. 112. Honneff H., Mecking H. (1978), Textures of Materials, Springer, Berlin. Hosford W.S. Jr., Backofen W.A. (1964), 9th Sagamore Conference, Army material research conference, Syracuse University Press, p.259. Houtte P. (1981), Proc. ICOTOM6, Iron and Steel Inst., Japan, p. 428. Hunter M.A. (1910), J. American Chemical Society, vol. 32, p. 330. Hutchinson J. W. (1970), Proc. R. Soc. Lond. A, vol. 319, p. 247. Hutchinson J. W. (1976), Proc. R. Soc. Lond. A, vol. 348, p. 101. Inagaki H. (1990), Zeitschrift Fur Metallkunde, vol. 4, p. 282. Inagaki H. (1991), Zeitschrift Fur Metallkunde, vol. 10, p. 779. Inagaki H. (1992), Zeitschrift Fur Metallkunde, vol. 83, p. 40. Iwakuma T., Nemat-Nasser S. (1984), Proc. R. Soc. Lond A, vol. 394, p. 87. Jones I.P., Hutchinson W.B. (1981), Acta Metalurgica, vol. 29, p. 951. Kallend J.S, Kocks U.F., Rollett A.D., Wenk H.R. (1991), Materials Science and Engineering A, vol. 132, p. 1. Klimanek P., Kuzel R. Jr. (1988), Journal of Applied Crystallography, vol. 21, p. 59. Kocks U.F. (1998), Texture and Anisotropy: Preferred orientations in polycrystals and their effect on materials properties, Cambridge University Press, U.K. Kocks U.F., Canova G.R. (1981), Deformation of Polycrystals, Riso Natn. Lab., Roskilde, Denmark, p. 35. Kocks U.F., Chandra H. (1982), Acta Metallurgica, vol. 30, p. 695.

95

Krivoglaz M. A. (1996), X-ray and Neutron Diffraction in Non-ideal Crystals, Springer-Verlag, Berlin Heidelberg, New York. Kröner E. (1961), Acta Metallurgica, vol. 9, p. 155. Kröner E. (1972), Statistical Continuum Mechanics, Springer Verlag, Wien. Kuzel R. Jr., Klimanek P. (1989), Journal of Applied Crystallography, vol. 22, p. 299. Kuzel R.Jr., Klimanek P. (1988), Journal of Applied Crystallography, vol. 21, p. 363. Louër D., Auffredic J.P., Langford J.I., Ciosmak D., Niepce J.C. (1983), Journal of Applied Crystallography, vol.16, p. 183. McGeary R.K., Lutsman B. (1952), Journal of Metals, vol. 191, p. 994. McQuillan A.D., McQuillan M.K. (1956), Titanium, Academic Press, New York, Butterworks scientific publications, London. Mizera J., Driver J.H., Jezierska E., Kurzydlowki (1996), Materials Science and Engineering A, vol. 212, p. 94. Mohamed G., Bacroix B. (2000), Acta Materialia, vol. 48, p. 3295. Molinari A., Canova G.R., Ahzi S. (1987), Acta Metallurgica, vol.35, p. 2983. Morris P.R. (1970), Int. J. Engeng. Sci., vol. 8, p.49. Nemat-Nasser S., Guo W. G., Cheng J. Y. (1999), Acta Materialia, vol. 47, No.13, p. 3705. Nemat-Nasser S., Obata M. (1986), Proc. R. Soc. London A, vol.407, p.343. Nye J.F. (1957), Physical Properties of Crystals, their Representation by Tensors and Matrices, Oxford Clarendon Press. Ogdom H.R., Gonser B.W. (1956), Rare Metals Handbook, C. A. Hempel, Reinhold press New York. Pan J., Rice J.R. (1983), Int. J. Solid Struct., vol. 19, No.11, p. 973. Partridge P.G. (1967), Metallurgical Reviews, vol.12, p. 169. Paton N.E., Backofen W.A. (1970), Metallurgical Transactions A, vol. 1, p. 2389. Philippe M.J., Beauvais C., Hocheid B. (1984), Proceedings of ICOTOM 7, p. 519.

96

Philippe M.J., Mellab F.E., Wagner F., Esling C., Hochheid B. (1988), Proceedings of the Sixth World Conference on Titanium, Les Ullis, p. 135. Philippe M.J., Serghat M., Van Houtte, P., Esling, C. (1995), Acta Metallurgica., vol. 43, p. 1619. Phillipe M.J. (1988), La Revue de Metallurgie, p. 1492. Phillipe M.J., Serghat M., van Houtte P. (1995), Acta Metallurgica et Materiallia, vol. 43, p. 1619. Polmear Ii. (1995), Light Alloys-Metallurgy of the Light Metals, John Wiley and Sons, New York. Popa, N.C. (1998), Journal of Applied Crystallogarphy, vol. 31, p. 176. Randle V., Engler O. (2000), Introduction to Texture Analysis: Macrotexture, microtexture and Orientation Mapping, Gordon and Breach Science Publishers. Reed-Hill R.E. (1992), Physical Metallurgy Principals, PWS-Kent Publishing Company, Boston. Reuss A. (1929), Z. Angew. Math. Mech., vol. 9, p. 49. Ribárik G., Ungár T., Gubicza J. (2001), Journal of Applied Crystallography, vol. 34, p.669. Roe, R.J. (1965), Journal of Applied Physics, vol. 36, p. 2024. Rosi F.D., Dube C.A., Alexander B.H. (1953), Transactions of AIME, February, p. 257. Rosi F.D., Perkins F.C., Seigle L.L. (1956), Transactions of AIME, January, p. 115. Sakai T., Fine M.E. (1974), Acta Metallurgica, vol. 22, p. 1359. Salem A., Kalidindi S.R., Doherty R.D. (2002), Scripta Materialia, vol. 46, p. 419. Scardi P., Leoni M. (2004), Diffraction Analysis of the Microstructure of Materials, E.J. Mittemeijer & P. Scardi editors, Berlin, Springer; p. 51. Schoenfeld S.E., Ahzi S., Asaro R.J. (1995), J. Mech. Phys. Solids, vol. 43, No.3, p. 415. Solas D.E., Tomé C.N., Engler O., Wenk H.R. (2001), Acta Materialia, vol. 49, p. 3791. Song S.G., Gray I.I.I.G.T. (1995), Acta Materialia, vol. 43, p. 2339. Stephens P.W. (1999), Journal of Applied Crystallography, vol. 32, p. 281.

97

Stokes A.R., Wilson A.J.C. (1944), I. Proc. Phys. Soc. London, vol. 56, p. 174. Taylor G.I. (1938), J. Institute of Metals, vol. 62, p. 307. Tomé C.N., Kocks U.F. (1985), Acta Metallurgica, vol. 33, No. 4, p. 603. Tomé C.N., Maudin, P.J., Lebensohn, R.A. and Kaschner, G.C. (2001), Acta Materialia, vol. 49, p. 3085. Ungár T, Borbély A. (1996), Applied Physics Letters, vol. 69, p. 3173. Ungár T, Dragomir I, Révész A., Borbély A. (1999), Journal of Applied Crystallography, vol. 32, p. 992. Ungár T. (2004), Scripta Materialia, vol. 51, p. 777. Ungár T., Groma I., Wilkens M. (1989), Journal of Applied Crystallography, vol. 22, p. 26. Ungár T., Tichy G. (1999), Phys. Stat. Sol., vol.171, p. 425 . Voigt W. (1889), Wied. Ann., vol. 38, p. 573. Wang Y.N., Huang J.C. (2003), Materials Chemistry and Physics, vol. 81, p. 11. Warren B.E. (1959), Progress in Metal Physics, vol. 8, p.147. Warren B.E., Averbach B. L. (1950), Journal of Applied Physics, vol. 21, p. 595. Warren, B.E., Averbach B. L. (1952), Journal of Applied Physics, vol. 23, p. 497. Wilkens M. (1970), Fundamental aspects of dislocation theory, J. A. Simmons, R. De Witt and R. Bullough, editors. NBS Spec. Publ., vol. 317, No.2, p. 1195. Wilkens, M. (1970), Physics of state solidi (a), vol. 2, p. 359. Williamson G.K., Hall, W.H. (1953), Acta Metallurgica, vol. 1, p. 22. Wood R.A. (1972), Titanium Alloys Handbook, Metals and Ceramics, Information Center, Battelle, Publication No. MCIC-HB-02, December. Yadav S., Ramesh K. T. (1997), Material Science and Engineering A, vol. 246, p. 265 . Yoo M.H. (1981), Metallurgical Transactions A, vol.12, p. 409. Zaefferer S. (2003), Materials Science and Engineering, vol. A344, p. 20.

98

Zeller R., Dederichs P.H. (1973), Physica Status Solid (b), vol. 55, p.831. Zhu Z.S., Liu R.Y., Yan M.G., Cao C.X. (1997), Journal of Materials Science, vol. 32, p. 5163.

99

BIOGRAPHICAL SKETCH Name: Gilberto Alexandre Castello Branco

Birth: March 19, 1963; Rio de Janeiro, RJ, Brazil

Citzenship: Brazilian

Occupation: Mechanical Engineer

Education: Graduated on the Fall of 1987 with B.Sc. in Industrial Mechanical Engineering at the

Centro Federal de Educação Tecnológica – Celso Suckow da Fonseca, Rio de Janeiro, Brazil;

graduated on the Summer of 1990 with a M.Sc. in Mechanical Engineering at the Military

Institute of Technology, Rio de Janeiro, Brazil. Thesis title: Sistema de posicionamento

automático de um canhão anti-aéreo. Graduated in 1995 with a specialization degree in

Economic Engineering at the Rio de Janeiro State University, Rio de Janeiro, Brazil. Joined the

Florida State University, Tallahassee, Florida, USA, on the Fall of 1999 for Doctorate studies,

and graduated on the Summer of 2005 with a Ph.D. in Mechanical Engineering. Dissertation

title: Effect of Thermo-mechanical Treatment on Texture Evolution of Polycrystalline Alpha

Titanium. Employment: Computer Programmer (1989 - 1991) and Systems Analyst (1991 -

1992) at DataPrev, Rio de Janeiro, Brazil, from 1989 to 1991. Adjunct Professor in the Eletronic

Engineering Department and from 1991 to 1992 in the Applied Mathematics Department at

Severino Sombra University, Vassouras, Rio de Janeiro, Brazil. Assistant professor at Centro

Federal de Educação Tecnológica – CSF- Rio de Janeiro, Brazil, College of Engineering,

Mechanical Engineering Department, from 1994 to 2000. Visiting Professor from 1995 to 1999

in the Mechanical Engineering Department at the Instituto Militar de Engenharia, Rio de Janeiro,

Brazil. Adjunct Professor at Centro Federal de Educação Tecnológica – CSF- Rio de Janeiro,

Brazil, College of Engineering, Mechanical Engineering Department, from 2000 to 2003

(Sabbatical Leave). Research assistant at the National High Magnetic Field Laboratory, Florida

State University from 1999 to 2003. Adjunct professor at Centro Federal de Educação

100

Tecnológica – CSF- Rio de Janeiro, Brazil, College of Engineering, Mechanical Engineering

Department, from 2003 to present.

Professional Experience: Programmer Skills (COBOL, FORTRAN, PASCAL, BASIC,

ALGOL) in Main frame and Personal Computer environments. CAD programmer certified by

AutoDesk. Developed experience in sample preparation for characterization of Titanium alloys,

Interstitial-Free Steels and Silicon Steel samples by X-ray diffraction and Orientation Imaging

Microscopy (OIM). Micro-characterization using Scanning Electron Microscopy (SEM), OIM,

optical microscopy and X-ray diffraction. Texture analysis; magnetic annealing; magnetic

properties measurements and macro, micro and nano-hardness tests. Teaching at Centro Federal

de Educação Tecnológica – CSF, Severino Sombra University and Instituto Militar de

Engenharia, Rio de Janeiro, Brazil.

101