Вы находитесь на странице: 1из 50

Ti51111: TRIP Titanium

Entry based on a design project in the Northwestern University, Materials Science and Engineering undergraduate capstone course: MatSci 390 Clients: Office of Naval Research and General Motors Group Members: Pitichon Klomjit, Frank Lin, Kelsey Stoerzinger, Wenhao Sun, Allison Weil, Tian Zhou Graduate Advisor: Jamie Tran Instructor: Dr. Gregory Olson Correspondence address: 2220 Campus Drive Evanston, IL 60208

A.U. Navy Photo from Austrailian Defence Jobs Website

Executive Summary
Motivated by the prospect of lower cost Ti production processes, new directions in Ti alloy design are explored for naval and automotive applications. Using the near- Ti5111 alloy as a reference alloy, the feasibility is assessed for application of transformation toughening to maintain the high toughness of Ti5111 at the higher 120ksi (827MPa) yield strength of Ti-6Al-4V. Principles established in steels are used to optimize the phase transformation stability for toughening. First principles quantum mechanical calculations are combined with available atomic volume data to model the composition dependence of the transformation volume change. Experimental measurement of the mechanical transformation stability of the phase in Ti5111 provides a calibration of transformation models. Combined with models of solution and grain refinement strengthening, a modification of the 5111 composition is designed meeting the transformation stability requirement while increasing the transformation volume change by a factor of 3 for efficient toughening.

Table of Contents
EXECUTIVE SUMMARY .......................................................................................2 1 MOTIVATION AND BACKGROUND ...............................................................5
1.1 Motivation ........................................................................................................................................ 5

1.2 Alternatives ...................................................................................................................................... 5 1.2.1 Steel and Aluminum ..................................................................................................................... 5 1.2.2 Titanium alloys ............................................................................................................................. 6 1.3 1.4 Limitations of Current Titanium Alloys........................................................................................ 6 Titanium Phase Relations ............................................................................................................... 6

1.5 Ti5111................................................................................................................................................ 7 1.5.1 Ti5111 Processing/Structure/Properties........................................................................................ 7

2
2.1

PROPERTY OBJECTIVES FOR TRIP TITANIUM ........................................8


Application Requirements .............................................................................................................. 8

2.2 Property Objectives ......................................................................................................................... 8 2.2.1 Specific Strength and Toughness.................................................................................................. 8 2.2.2 Corrosion Resistance .................................................................................................................. 10 2.2.3 Weldability.................................................................................................................................. 10 2.2.4 Alloying Cost .............................................................................................................................. 10

3 4 5

TEAM ORGANIZATION...............................................................................11 SYSTEM STRUCTURE................................................................................12 DESIGN APPROACH ..................................................................................13

5.1 Strength Model............................................................................................................................... 13 5.1.1 Hall-Petch Strengthening ............................................................................................................ 13 5.1.2 Solid Solution Strengthening ...................................................................................................... 14 5.2 Transformation Toughening......................................................................................................... 15 5.2.1 TRIP Steels and Ceramics .......................................................................................................... 15 5.2.2 Martensitic Transformation in Titanium Alloys ......................................................................... 16 5.2.3 Modeling the Martensitic Transformation in Titanium .............................................................. 17 5.2.4 Ab-initio Calculations of Molar Volume for Control of Transformation Volume Change........ 19 5.2.5 Transformation Stability Measurements..................................................................................... 21 5.3 Cost Model...................................................................................................................................... 21

DESIGN MODELING....................................................................................22

6.1 Hall-Petch Grain Size Refinement ............................................................................................... 22 6.1.1 Lineal Intercept Measurement .................................................................................................... 23 6.1.2 Grain Size Modeling ................................................................................................................... 24 6.2 Solid Solution Strengthening ........................................................................................................ 25

6.3 Martensitic Transformation Toughening.................................................................................... 27 6.3.1 Volume Change of Martensitic Transformation in Titanium ..................................................... 27 6.3.2 DFT Atomic Volume Calculations ............................................................................................. 27 6.3.3 Redlich-Kister Fitting to Atomic Volume Data.......................................................................... 30 6.3.4 Interfacial Friction of the Martensitic Transformation in Titanium ........................................... 33 6.3.5 Ms calculations and Measurement ............................................................................................. 34 6.3.6 Ms calculations for Ti5111 and Model Calibration ................................................................... 37

DESIGN INTEGRATION ..............................................................................40

7.1 Design Constraints......................................................................................................................... 40 7.1.1 Composition constraints.............................................................................................................. 40 7.1.2 -phase fraction constraints ........................................................................................................ 42 7.1.3 Annealing temperature constraints ............................................................................................. 42 7.1.4 Mo partitioning interaction with Fe ............................................................................................ 43 7.2 Final Alloy Composition................................................................................................................ 44 7.2.1 Composition Refinement ............................................................................................................ 44 7.2.2 Final Composition....................................................................................................................... 44 7.2.3 Annealing Temperature and Ms ................................................................................................. 45 7.2.4 phase composition.................................................................................................................... 45 7.3 Processing ....................................................................................................................................... 46

7.4 Strengthening Results.................................................................................................................... 46 7.4.1 Solid Solution Strengthening ...................................................................................................... 46 7.4.2 Hall-Petch Grain Refinement...................................................................................................... 46 7.5 Toughening Results ....................................................................................................................... 47

8 9 10

CONCLUSION AND RECOMMENDATIONS ..............................................47 ACKNOWLEDGEMENTS ............................................................................48 REFERENCES..........................................................................................49

Motivation and Background

1.1 Motivation
Recent innovations have led to low cost production methods such as the FFC/Cambridge and the Armstrong/ITP methods which may reduce the cost of titanium by amounts on the order of 30% to 50%.1 Spurred by the prospect of more affordable Ti alloys, the United States Navy and the automotive industry are re-examining titanium alloys in the search for reliable structural materials that are lightweight, strong, tough, and corrosionresistant. Titanium alloys are well known for their high strength-to-weight ratio. This property first rendered them ideal for aerospace applications, but their use has since been extended to other realms where high strength and low density are desired. In fact, titanium alloys may achieve the high strength of certain steels at a fraction of the weight. Titanium is also very corrosion-resistant, forming a very thin, stable layer of titanium oxide that serves to chemically protect the base metal underneath. The principal drawback of titanium has been high cost. Realizing the potential for low cost titanium in military, naval, and aerospace applications, the United States Office of Naval Research (ONR) has supported research into tailoring alloy properties to a variety of needs. The Navy in collaboration with Titanium Metals Corporation (Timet) has recently developed an alloy, Ti5111 (Titanium Five Triple One), for structural marine applications. This alloy has a composition of 5wt%Al, 1wt%V, 1wt%Sn, 1wt%Zr, 0.8wt%Mo, and 0.1wt%Si. It possesses intermediate strength, high toughness, excellent stress-corrosion cracking resistance, and weldability. In addition to support from ONR2, General Motors (GM) has awarded a grant to Northwestern3 in response to recent government regulations that require an automobile to run at 35 miles per gallon by the year 2020.4 Many high performance components of an automobile engine require the use of heavy steel. The high strength-to-weight ratio of titanium has the potential to reduce the energy consumption of an automobile by 25% when replacing steel products.5 A new alloy combining the strength of commercial alloy Ti-6Al-4V with the toughness of Ti5111 will be designed by exploiting transformation toughening. Grain size refinement and solid solution strengthening models are used to obtain the increase in strength. Building on principles established in high-performance steel design, a new model for transformation toughening of titanium is developed to compensate for the concomitant decrease of toughness with increase in strength; resulting in a TRansformation Induced Plasticity (TRIP) titanium alloy design.

1.2 Alternatives
1.2.1 Steel and Aluminum High-strength alloyed steel is the primary material currently used for submarine hulls and other shipbuilding applications. However, the low corrosion resistance and strength-toweight ratios of steel limit its use from many marine applications.6

The low density, non magnetic nature and lower cost of aluminum (Al) compared to titanium (Ti) make it desirable for a wide variety of applications. However, the low absolute strength and low corrosion resistance of Al alloys make them unsuitable for many structural applications in naval applications. 1.2.2 Titanium alloys The commercial titanium alloy Ti-6Al-4V (Ti64) accounts for more than 50% of total Ti alloy tonnage worldwide: the aerospace industry accounting for 80% of this usage, and the automotive, marine, and chemical industries also use small amounts.7 Although the strength of Ti64 (Table 1) is high, the fracture toughness and stress corrosion cracking resistance of the material are insufficient for structural naval applications. The alloy Ti100 with a composition of Ti-6Al-2Nb-1Ta-0.8Mo (values in wt%) was originally developed by the Navy for high toughness and moderate strength applications. While the properties of the alloy met the Navys strength requirements (Table 1), the expensive alloying elements and manufacturing difficulties prevented its use.
Table 1: Mechanical and physical properties of Ti, Al, and steel alloys8

Yield Strength (ksi) Density (lb/in3) Elongation at Break Strength/Density Kic Fracture Toughness (ksi in1/2)10 Salt Water Corrosion Resistance10

Ti5111 100 0.160 15% 625 107-11311

Ti100 100 0.162 11% 617 100-11012

Ti64 120 0.160 14% 813 74.6-91 Very good

Al 7075-T6 73 0.102 11% 716 27.3-30 Good

HSLA1009 100 0.284 11-13% 352 Average

Very good11 Very good12

1.3 Limitations of Current Titanium Alloys


While the excellent corrosion resistance and strength-to-weight ratio of current Ti alloys make them particularly desirable for Navy applications, there are still several drawbacks. In addition to relative cost, the high affinity of Ti for oxygen dictates that Ti alloys cannot be produced and welded in atmosphere. Production and fabrication steps require the use of vacuum or inert gas atmospheres which is difficult when large sections of Ti are needed. Furthermore, available fundamental data and research on Ti and its alloys are limited compared to other materials such as steel with a longer history as a structural material.

1.4 Titanium Phase Relations


There are two different crystal structures of pure titanium, a low temperature HCP alpha () phase, and a high-temperature BCC beta () phase. The transition between these phases occurs at 882oC, termed the -transus temperature. This temperature can be modified by alloy composition, which can be used to engineer combinations of and titanium phases tailored to desired microstructures.

Titanium alloys fall into two main categories, -Ti and -Ti alloys, denoting the dominant phase. Notable mechanical properties of alpha alloys include high resistance to fracture, as well as fatigue. They also tend to be easier to weld than other titanium alloys, and are highly resistant to corrosion. Beta alloys can achieve higher strengths than alpha alloys, but are also less tough. In general, specific properties can be attained by tuning the different fractions of alpha and beta phases. These alloys are called + alloys, generally with beta phase concentrations over 10%, featuring properties between the -Ti alloy and -Ti alloy extremes. Alpha phase stabilizers include aluminum and tin, which replaces titanium substitutionally, and oxygen, which occupies the titanium lattice interstitially. Beta phase stabilizers include substitutional molybdenum (Mo), vanadium (V), silicon (Si), iron (Fe), and interstitial hydrogen (H). There are also neutral alloying elements, such as zirconium (Zr), that affect the properties of the material without significantly changing the beta transition temperature. Titanium can transform into a martensite phase through a Bain transformation of the BCC -Ti lattice to the HCP '-Ti lattice. There is also an orthorhombic ''-Ti martensitic phase, which can be described by an incomplete BCCHCP Bain transformation associated with high alloying.13

1.5 Ti5111
The U.S. Naval Research Laboratory worked in conjunction with the Titanium Metals Corporation (Timet) to develop an improved structural alloy as an alternative to Ti100 for use primarily in ship hulls. Requirements of the material include high strength (especially a high strength-to-weight ratio), high fracture toughness, and superior marine environment corrosion resistance. Ti5111, met all of these requirements set by the Navy (Table 1) and is more cost efficient and easier to manufacture than Ti100. 1.5.1 Ti5111 Processing/Structure/Properties The microstructure and properties of an alloy are controlled though optimized processing. Hot rolling of Ti5111 above the transus temperature (1093C) followed by a two-step duplex annealing treatment in the phase field (1010C) and + field (954C) is used to obtain a lamellar microstructure with a limited retained phase content (termed a near- alloy). During both annealing steps the material is held for 1 hour per 25 mm of thickness with an air cooling of ~15-20C/min after each stage. The optimized microstructure of Ti5111 is characterized by a basket weave or Widmansttten microstructure (Figure 1). When the high temperature phase is cooled at slow rates below the transus, the low temperature incoherent phase begins to nucleate at grain boundaries. Further cooling leads to the nucleation of additional phase at both and grain boundaries that grow into the grains as parallel plates11.

Figure 1: Widmansttten Microstructure of Ti511111

The average yield strength of Ti5111 in the longitudinal direction is 100ksi, with minimum yield strength of 85ksi. The fracture toughness values are also excellent, as are stress corrosion resistance properties14. Tests show that Ti5111 had no visible crevice corrosion after a full year of exposure to seawater15. This alloy also costs less to produce than other Ti alloys of comparable properties.

2 Property Objectives for TRIP Titanium 2.1 Application Requirements


The first step in determining the property objectives of TRIP Titanium is to understand the requirements of the application it will be used for. A structural alloy must have sufficient strength and toughness to withstand the physical forces acting upon it. The alloy must also have a high corrosion resistance for use in corrosive environments. The ability of components to be welded in air would also be desirable for ship building.

2.2 Property Objectives


2.2.1 Specific Strength and Toughness In this project, we seek to design an alloy with the toughness of Ti5111 (110ksi in1/2) and the strength of Ti64 (120ksi). A large toughness protects against projectiles, explosions, and general usage and a high strength allows the use of the design alloy as a structural hull material. Furthermore, using high specific strength and specific toughness (values normalized by density) criteria allows for the design of strong and fracture resistant materials that enable lightweight structures. As previously seen in Table 1, titanium alloys can provide higher specific strength and specific toughness than the other high performance engineering alloys. Cambridge Engineering/Materials Selector (CES/CMS) is used to plot and compare property objectives to currently available materials. As lightweight structures are important to the Navy, material properties normalized by density are plotted. To limit the CES plots to relevant materials, restrictions/minimum criteria were applied. A minimum elongation of 10% is placed on the materials as this is a criterion of weldable metals in the Navy. Also the mechanical property constraint of a minimum fracture toughness of 50ksi in is placed to eliminate materials that would not be appropriate for structural applications for the Navy. Additional constraints of a material having average, good,

or very good freshwater properties are included. Most importantly the plot was obtained with the restriction that all materials had to perform very well in seawater. These five criteria were entered into the CES software and the following plot was produced:

Figure 2: CES Plot of Specific Toughness vs Specific Strength

Figure 2 cross plots the specific fracture toughness and specific strength of all materials that have passed the aforementioned selection parameters. The main material competitors in this plot are titanium (purple), nickel (red), and stainless steels (teal). Ti alloys offer good specific fracture toughness to specific strength values, evidenced by their location in the upper right hand corner of the plot. This makes Ti alloys a good starting point for the development of a material that best meets our applications needs. The two Ti alloys previously mentioned, Ti5111 and Ti-6Al-4V, are detailed in the CES plot. As the project goal location shows, the property goals for the alloy being developed are in a range that no current alloy achieves.

The lines in Figure 2 represent values for the critical flaw size (acrit) at the yield stress, (YS), which can be defined as:

Equation 1

The minimum fracture toughness to yield strength (KIC/YS) ratio allowed by the Navy is 0.5, while a value of 1 is ideal. The minimum and ideal ratios correspond to critical flaw sizes of 0.08 and 0.32 respectively. Ti5111 surpasses the minimum ratio and closely approaches the ideal ratio. 2.2.2 Corrosion Resistance Naval materials have continuous exposure to seawater, and therefore corrosion resistance is very important. Elemental Ti and Ti alloys form a layer of TiO2 upon exposure to oxygen, which acts as a native oxide passivation layer that provides corrosion resistance. This property gives Ti alloys an advantage over other alloys that do not form such a layer, for example nonstainless Fe alloys, which need additional protective coatings prior to use in marine applications. Relevant to many current real-world industrial and naval applications, Ti alloys resist both unpolluted and polluted seawater corrosion up to 80C, in atmospheric corrosion and microbiologically induced corrosion.16 Crack tips resulting from material deformation are plagued by a lack of sufficient TiO2 buildup, threatening the material integrity. For this reason it is important to restrict certain alloying elements which increase the susceptibility of Ti alloys to stress corrosion cracking, including Al, Zr, oxygen (O), and tin (Sn). Other alloying elements, Mo and V, increase resistance to stress corrosion cracking. Part of the stress corrosion resistance of Ti5111 is due to the low weight percentages of Al (5 wt%) and O (0.09wt%).17,18,19 2.2.3 Weldability Titanium alloys can be welded by gas tungsten arc welding, plasma arc welding, or a number of other techniques. Due to the high temperatures during welding, the weld site must be isolated from oxygen and nitrogen gas that can embrittle the area if these elements are dissolved into the metal. In addition, stresses from welding must be relieved to increase fatigue resistance in the weld area. Adding yttrium (Y) to the alloy and the weld filler helps to getter the oxygen introduced during the weld process by forming Y2O3 precipitates. These precipitates, formed in lieu of more brittle metal oxides, also help strengthen the welding point by pinning -phase grain boundaries during recrystallization to reduce the grain size.20 2.2.4 Alloying Cost Although titanium alloys exhibit many exceptional properties that make them desirable for a wide range of applications, cost remains a primary limitation. Alloy cost is controlled by both raw material costs and processing costs. Assuming the processing costs are similar for other Ti plate material, the cost of Ti alloys then is driven by the amount and type of alloying elements used. The most commonly used Ti alloy, Ti64, will serve as a relative alloying cost constraint for the project. 10

3 Team Organization
Pitichon Klomjit will assess strengthening mechanisms for Ti5111 type alloys. He is particularly interested in working with metals, and is currently working with the Olson group to improve the mechanical properties of alloys. The course MSE 332 Mechanical Properties of Materials, taught by Professor Panico, introduced him to many strengthening mechanisms, such as solid solution strengthening, precipitate hardening, and grain size refining. He is excited to apply this knowledge quantitatively to real world applications. Currently in her junior year, Kelsey Stoerzinger has worked for over a year in the research group of Prof. Teri Odom, researching the fabrication, self assembly, and optical properties of nanoparticles. The many classes she has taken at Northwestern University have prepared her for work in many other areas as well. She was introduced to the subject of solid solution strengthening in Professor Seidmans 316 course sequence on Microstructural Dynamics. She is very interested in material behavior at interfaces and is excited to investigate strengthening mechanisms and microstructure control. To control the volume change of martensite transformation, Wenhao Sun, a joint major in materials and applied mathematics, will focus on using Density Functional Theory to predict molar volumes of primary titanium phases, as well as their relative formation energies. Wenhao has considerable experience using DFT and other first-principles quantum mechanical simulations in his research in the research group of Prof. Chris Wolverton, and has also taken Professor Wolvertons MSE 510 graduate-level Computational Materials Science course. Wenhao has always been interested in predicting the behavior of novel high-performance materials through first-principles, which is why he is particularly interested in predicting the properties of this new hightoughness titanium alloy. Allison Weil, currently a pre-senior, is a co-op at Federal-Mogul who has worked on materials development and failure analysis for metallic gaskets, including research in high temperature materials for exhaust gasket applications. This experience has further familiarized her with metallic mechanical properties. Several courses have prepared her for this project, and she gained an interest in toughening mechanisms due to Professor Dunands MSE 332 Mechanical Properties of Materials course, and in this project will be focusing on this aspect. Currently a senior, Tian Zhou has worked for the past year in the Shull group researching strengthening mechanisms in polymeric materials. In this project, he will explore the concept of transformation toughening, applying the mechanism of TRIP steels to Ti alloys. Professor Dunands Mechanical Properties of Materials course has prepared him well for analyzing this particular system. Senior Project student Frank Lin is currently working with the Olson group to measure the mechanical properties of titanium alloys. Frank has coordinated his experimental research with the TRIP titanium design project, building on his experience as a member

11

of last years MatSci 390 Materials Design titanium project. He will perform BollingRichman tensile tests to measure the Ms temperature of the reference Ti5111 alloy to provide an experimental calibration point for toughening models.
Table 2: Responsibility Allocation Matrix (X = primary, O = secondary) Stoerzinger Responsibilities Klomjit Sun Weil Zhou Strength Hall Petch Strengthening X O Solid Solution O X Strengthening SCC Constraints X O Toughening Frictional work Volume change Density Functional Theory phase stability Ti5111 stability measurement Cost

Lin

O X X X O X

X O O X

4 System Structure

Figure 3: System Chart

12

Systems engineering has had great success in materials design21. This method of engineering works by optimizing a hierarchy of subsystems where the requirements are set at the highest level: property priorities are optimized by materials structure, which is controlled by processing, see Figure 3. For the design of this particular alloy, the primary performance requirements are a high strength (equivalent to that of Ti64), high fracture toughness (equivalent to Ti5111), and resistance to stress corrosion cracking. Alloy strength and corrosion resistance is provided by grain refinement and the solid solution strengthening of the phase. Toughness is increased by optimized martensitic transformation of the phase during deformation. A balance of both properties is maintained with the prevention of the formation of embrittling phases such as Ti3Al. A sequence of processing steps is used to control microstructure. The initial cooling rate and annealing provide grain refinement and solution strengthening for the phase, while the secondary anneal and subsequent cooling affect the final composition determining how effectively strain induced martensitic transformations will act as a source of toughening.

5 Design Approach 5.1 Strength Model


There are several approaches to strengthening a titanium alloy: grain size refinement (Hall-Petch strengthening), interstitial solid solution strengthening, substitutional solid solution strengthening, precipitation, and the control of crystallographic texture. Based on the Ti5111 type of alloy as our starting point, our initial approach will emphasize grain size refinement and solid solution strengthening. 5.1.1 Hall-Petch Strengthening The strength of a crystalline material is determined by the force required to move dislocations through the lattice, resulting in macroscopic deformation. Creating more obstacles for dislocation movement thus improves strength. One such obstacle is grain boundaries. In their two classic papers, Hall and Petch studied two different material behaviors, but arrived at the same conclusion. Hall published three papers about the influence of grain size on mechanical properties, while Petch focused on the brittle failure of steels by measuring cleavage strength with respect to grain size. The relationship that the two scientists developed is:22

Equation 2

where D is the average grain diameter, y is the yield stress, 0 is a material constant for the base stress for dislocation movement, and k is the strengthening coefficient. It is noted that the Hall-Petch relationship holds true only for grain sizes between 1 millimeter and 1 micrometer. Studies on many nanocrystalline materials with grain sizes less than 100 nanometers have shown that the yield strength either remains constant or

13

decreases with further decrease of grain size. This phenomenon has been termed reverse or inverse Hall-Petch behavior as shown in Figure 4:

Figure 4: Schematic of the variation of hardness H with grain size D23

5.1.2 Solid Solution Strengthening The solid solution of atoms within the matrix can also act as obstacles to dislocation propagation. When solid solution atoms are dissolved into the lattice, local stress fields are formed that interact with the dislocations. The behavior can differ for the two types of solid solution alloying: substitutional and interstitial. The strengthening increment, , arising from solid solution strengthening can be estimated as:
Equation 3

where G is the shear modulus, c is the concentration of the solute atoms, b is the magnitude of the dislocation Burgers vector, and is the atomic misfit strain. The G, b, and values are material dependent terms and can be grouped into a constant K. This allows an estimate of the strengthening effectiveness of a solute from its atomic size misfit, and the relationship has been experimentally verified for interstitial-element additions in titanium systems.24 Experimentally measured solution strengthening coefficients by Collings24 and Hammond25 show a more accurate description of strengthening increment due to substitutional metal additions can be made using a linear relationship of strengthening increment, , to concentration:
Equation 4

These solid-solution relationships, experimentally determined, are summarized in Table 3 for transition elements and in Table 4 for common substitutional and interstitial-element additions to titanium alloys. While Table 4 provides coefficients for hardness, Collings notes that hardness values in Ti alloys are linearly related to the strength values26, the

14

experimental relationship between which is determined for the solid-solution strengthening model in Section 6.2.
Table 3: Solid-Solution Strengthening of Titanium by Transition Elements (Recreated from Collings, Table 5-4)25

Solid-solution strengthening rate, MPa Per wt. % Per at. %

V 19 20

Cr 21 23

Mn 34 39

Solute Element Fe Co 46 48 54 59

Ni 35 43

Cu 14 18

Mo 27 54

Table 4: Solid-Solution Hardening of Titanium by Substitutional and Interstitial Elements (Recreated from Collings Table 5-3)24 Hardening rate, Slope, b, dHv/dc, Conc kg mm-2 Intercept, Correlation kg mm-2 at. %-1 range, Condition* Law** at. %-1/2 or a, coefficient, at% kg mm-2 kg mm-1/2 % At 0.1 At 1.0 -1 at. % at. % at%

(a) Substitutional additions Al 0-10 100h/850C/I BQ Ga 0-5 As-cast Si 0-2 1h/100C/IB Q Ge 0-5 1h/100C/IB Q Sn 0-7 As-cast (b) Interstitial-element additions B 0-0.2 120h/800C/I BQ C 0-0.5 120h/800C/I BQ N 0-5 120h/800C/I BQ O 0-3 120h/800C/I * IBQ=Ice Brine Quenched BQ

c c c c c c1/2 c1/2 c1/2 c1/2

15 24 39 33 24 218 170 239 194

102 108 125 120 112 110 104 98 100

99.6 99 83 98 99 92 99.9 99.8 99.9

344 269 678 307

15 24 39 33 24 109 85 120 97

**Data fitted to either Hv = a + bc or Hv = a + bc1/2

5.2 Transformation Toughening


5.2.1 TRIP Steels and Ceramics TRansformation-induced plasticity (TRIP) steels have demonstrated great promise by offering good ductility and toughness at high strength. Transformation plasticity results from mechanical transformation of a metastable austenite phase at operating temperatures. Strain at a crack tip triggers the austenite to martensite transformation, enhancing strain hardening to stabilize flow against localization. Overall toughness can be greatly improved by this mechanism. Utilized for ceramics as well as metals, transformation toughening is also effective in brittle fracture, but not to the same extent as ductile fracture. In brittle fracture, transformation shear and dilation greatly increases the fracture process zone volume and

15

subsequently the toughening interaction. The transformation toughening in brittle solids scales with the square of the transformation dilation.29 In ductile solids, transformation toughening occurs on three structural scales. At the largest scale, the transformation zone is enlarged because transformation hardening near the crack tip delays shear localization in this region. At a smaller scale, the transformation also delays microvoid-softening. These microvoids assist in crack propagation, so in limiting their growth, the fracture is restricted. This is the largest toughening effect on the ductile material. Finally, the inhibited shear localization in the transformation zone can promote crack branching, which in turn increases the fracture process zone size. The combination of these factors leads to a greater toughening effect in ductile versus brittle solids, with JIC toughness increment scaling with volume change to the third power.29 5.2.2 Martensitic Transformation in Titanium Alloys The concept of transformation-induced plasticity may be extended to titanium alloys, exploiting martensitic transformation of the parent phase. Although this process may occur spontaneously on cooling, it is desirable to induce the transformation via plastic strain to enhance toughness.27 Beta-stabilizers, such as Mo and V, may be added to control the stability of a metastable parent phase at operating temperatures. Figure 5 shows the dependence of stress-assisted and strain-induced martensitic transformations on stress and temperature. Ms is the temperature at which martensitic transformation begins upon cooling; Ms is the temperature at which stress-assisted transformation occurs at a stress level equal to the yield stress of the parent phase, and Md the temperature above which no deformation-induced transformation occurs.29 A stressassisted martensitic transformation may occur when the temperature lies between Ms and Ms where transformation controls yielding. In addition to lowering the yielding strength, large martensitic plates form in this region which can reduce toughness. In contrast, formation of fine strain-induced transformation increases toughness, and will occur at a temperature between Ms and Md. The largest increase in toughness occurs at just above Ms. These two regimes of transformation are delineated in Figure 5. Due to the interaction of hydrostatic stress with the transformation volume change, the characteristic mechanical transformation temperature, Ms, strongly depends on stress state. According to the equivalent stress plot of Figure 6, Ms under uniaxial tension differs from that for the triaxial stress at a crack tip. Since uniaxial tension determines uniform ductility and the crack tip stress state dictates fracture toughness, it is difficult to optimize transformation stability for both uniform ductility and toughness at the same temperature. Because improving fracture toughness is a primary goal of this project, Ms will be optimized for the crack-tip stress state.

16

Figure 5: Stress and temperature relations for stress-assisted and strain-induced martensitic transformations.28

Figure 6: Martensitic transformation temperature for uniaxial tension

and crack-tip

.29,30

5.2.3 Modeling the Martensitic Transformation in Titanium A quantitative model of martensitic transformation in titanium may be derived by adapting principles established in steels. The general Olson-Cohen model describes the heterogeneous nucleation of martensite. The driving force for transformation of the parent phase to martensite is composed of mechanical and chemical components.31

17

Equation 5

Here Gtot is the total free energy change of the martensitic transformation, Gchem is the chemical free energy difference between parent and martensite phases at a given temperature in units J/mol. Gchem values are computed by ThermoCalc using the licensed Thermotech Ti-DATA-v3 thermodynamic database. Gmech is given by the following equation:34

Equation 6

The three terms in Equation 6 have been developed over many years from steel research. Patel and Cohen32 derived the first term in Equation 6 from the known transformation shear strain for martensitic transformations in steel alloys where is the equivalent stress. Since the transformation strains in steels are ~20% and those in Ti are ~14%,33 the coefficient in the first term is reduced by a factor of 0.7 to accommodate the reduction in transformation strain in Ti. The second term describes the effect of hydrostatic stress, h, and parent to martensite volume change V/V on the driving force. The h term is described by for a crack tip stress state and for a uniaxial tension stress state; all stress units in Equation 6 in MPa. The volume change may be found by fitting lattice parameter measurements of parent and martensite phases in Ti with changes in composition. This composition dependence can be described by the same Redlich-Kister polynomials employed in the ThermoCalc database. The third term in the equation is based on analysis of the Cech-Turnbull small particle experiments, defining a potency distribution of the heterogeneous nucleation sites controlling martensitic transformations. Cohen et al. extended this exponential nucleation site distribution, adding the effect of applied stress to the effective potency distribution.34 By combining the statistical model of structural potency distribution with a driving-force distribution for random orientation distribution of nucleation sites, relations were obtained for stress-assisted transformation plasticity, and the nonlinear expression for Gmech() was determined. The Olson-Cohen model31 further predicts that martensitic nucleation occurs at a critical free energy change:
Equation 7

Equation 8

18

Here Gn is a function of nucleation site potency, G0, particle volume Vp of the transforming phase, a reference V0 and a proportionality factor K. The ln term in Equation 8 reflects the increased stability of the parent phase when small particle volume reduces the characteristic potency of available nucleation sites. The boundary between microscopic and macroscopic behaviors is determined by the reference volume, defined as the volume of a sphere of diameter 100m.34 As the particle volume increases stability of the phase decreases. Wf is the frictional work opposing glide of the martensitic interface in the solid solution. As discussed in Section 6.3.4, analysis of transformation data in Ti alloys suggests this friction is a linear function of composition:

Equation 9

Ki serves as the proportionality constant and Xi is the atomic fraction of alloying element i. This linear relationship is equivalent to the linear relationship of solid solution strengthening for slip measured by Collings.24 When the testing or use temperature is T= Ms and the stress applied energy change may be set equal to Gcrit, yielding: =ys, the total free

Equation 10

as the basis for predicting Ms. The desired Ms for this design project is room temperature (300K or 27C) and the target yield strength is 120ksi or 827MPa. 5.2.4 Ab-initio Calculations of Molar Volume for Control of Transformation Volume Change Application of transformation toughening principles established in steels to Ti alloys has already been demonstrated in a system consisting of -Ti-Al-V particles in a -Ti-Al intermetallic alloy by Grujicic.35 He assumed that toughening follows the same dependence on thermodynamic stability of the parent phase and volume change during the martensitic transformation as in steels. Applying these same concepts to our design, the phase should be well-stabilized to promote a strain-induced transformation. The transformation itself should generate as large a volume change as possible. One of the many computational methods now used by materials engineers is Density Functional Theory (DFT), a quantum mechanical method that has proven extremely successful in describing a wide range of solid-state systems. The foundation of DFT is formed by the two Hohnberg-Kohn theorems, developed in 1964. The first theorem states that the ground-state properties of a many-electron system can be uniquely described by its three-dimensional electron density. Finding the total electronic-energy of a complicated system using this approach is significantly cheaper computationally than solving for the energy of a many-body wave function. The second theorem states that the 19

correct ground-state electron density will effectively minimize the energy functional of the system. One of the most elementary tasks in DFT is determining the molar volume of a material. However, this is hindered by one of the fundamental difficulties in DFT - that is, determining a suitable exchange-correlation for approximation of the energy contributions from the electron density. Two most prominent methods are the Local Density Approximation (LDA) and the Generalized Gradient Approximation (GGA) exchange correlations. LDA approximates the contribution to electronic energy by assuming that a volume element constitutes a uniform homogeneous electron gas, whereas GGA also considers the gradient of the electron density within that element. GGA tends to give results more consistent with experiment, and so we will perform molar volume calculations of different binary titanium phases using the GGA exchange correlation. Because there are many calculations, we will reduce time by using pseudopotentials by Perdew and Wang36 instead of full-potential LAPW calculations. These DFT calculations will first be applied to cases where experimental data are available to assess model accuracy. The technique will then be applied to assemble a molar volume database for all components of interest for this project. As mentioned earlier, the database will employ the same Redlich-Kister polynomial format employed by the ThermoCalc software. In addition, DFT can be utilized to map relative energies at 0K of the phases of interest for comparison with ThermoCalc database predictions. Because the electronic totalenergy of a material is a direct function of its volume and lattice parameters, it is possible to determine energetic minima and metastable diffusionless phases by relaxing an inherently unstable structure to a ground state minimum, where lattice parameters can be determined. The martensitic lattice transformation of titanium is described by the Bain distortion, whereby BCC titanium deforms uniformly into an HCP structure. Although highly alloyed titanium is observed to exhibit an orthorhombic structure (termed ), which corresponds to an incomplete BCCHCP transformation13, the volume change in a martensitic transformation can be well-approximated by modeling the HCP cell.* The primitive-cell to be run in DFT is generated by placing titanium and its alloying atoms on a HCP lattice. DFT relaxation of this transformed cell should allow prediction of the composition dependence of the axial ratios as well as the molar volumes of these structures, pictured below in Figure 7.

Originally the orthorhombic cell was modeled by applying a transformation to the unit BCC cell, although it was found to naturally relax back to either the BCC or HCP cell energetically.

20

Figure 7: DFT will model the volume change of a Bain deformation of a BCC phase of binary Ti-X alloys (left) to an orthorhombic structure (right), approximated by an HCP lattice.

5.2.5 Transformation Stability Measurements As represented in Figure 5, a reversal in the temperature of yield strength defines the Ms temperature in uniaxial tension. The test for MS temperature is done following the Bolling-Richman test method.37 Tensile tests were performed on a single flat dogbone specimen in an enclosed thermal temperature controlling device. The temperature of the chamber is set to 180C with a loading rate of 0.5mm/min. When the stress-strain curve extends into the plastic region where the slope decreases, the sample is unloaded. After each loading and unloading sequence, the temperature is reduced by 20C and the sequence is repeated. Cooling is achieved either by air injection or nitrogen gas injection to the lowest attainable temperature of -90C is reached.

5.3 Cost Model


As discussed earlier in Section 2.2.4 the cost of producing Ti alloys is dictated by the alloying elements once the processing costs are assumed to be similar to current Ti plate alloys. The cost of elements generally used in Ti alloys is listed in Table 5, along with their price ranges and average prices. These values were obtained from the CES 2008 database10. A second case for cost of Ti is also listed, corresponding to projected cost reduction of the new Ti production processes. It is readily apparent that the cost of Ti alloys tends to be high by virtue of the fact that the majority of the composition is Ti, a relatively expensive metal. Final alloy cost will therefore be sensitive to the effectiveness of the new Ti production processes. The use of expensive alloying elements, such as vanadium, can also dramatically increase the cost of the alloys. The calculated raw material cost of Ti alloys of interest for marine and automotive applications is listed in Table 6. These costs are calculated directly from the compositions by weight:

Equation 11

21

where Wi and Ci are the weight fractions and costs per weight of alloying element i, respectively. As expected, Ti64, with high V content, is the most expensive alloy, whereas Ti5111 is 30-40% lower in cost. In order to design a cost-efficient Ti alloy this cost model will be applied.
Table 5: Costs of elements in Titanium Alloys10

Price Range (USD/lb) Element Average Ti (1) $36.60 $40.20 $38.40 Ti (2) $24.40 $26.80 $25.60 V $527.00 $658.00 $592.50 Ta $188.00 $254.00 $221.00 Nb $84.70 $122.00 $103.35 Ni $19.20 $21.10 $20.15 Mo $12.00 $15.00 $13.50 Zr $12.40 $13.60 $13.00 Sn $6.02 $6.62 $6.32 Si $4.14 $6.87 $5.51 Cr $3.49 $3.84 $3.67 Cu $3.05 $3.36 $3.21 Al $1.14 $1.25 $1.20 Fe $0.35 $0.38 $0.37
Table 6: Costs of Titanium Alloys

Titanium Alloys Ti5111 (Ti-5Al-1Sn-1V-1Zr-0.8Mo) Ti64 (Ti-6Al-4V-0.25Fe)

Cost (USD/lb) With Ti(1) With Ti(2) $41.26 $29.59 $58.24 $46.75

6 Design Modeling 6.1 Hall-Petch Grain Size Refinement


The Hall-Petch relationship was used to first assess if the yield strength of a Ti alloy could be increased by reducing the grain size to meet the desired final yield strength of 120ksi (827MPa) through an adjustment of the processing. To quantify the relationship between grain size and yield strength in the system of interest, the values for the reference alloy Ti5111 processed with different schemes (slow cooled, fast quenched of wrought Ti and SPS Ti), resulting in different microstructures, are used (Table 7). The wrought material was provided by the Navy and the SPS Ti5111 alloys were powder processed in the doctoral research of Jamie Tran. The slow cooled alloys are first brought to 30C above the -transus for 30minutes and cooled at 15C/min to room temperature. The alloy is then held at 30C below the -transus for another 30min and cooled to room temperature at 15C/min. For fast quenched alloys, the alloy is first 22

brought to 30C above the -transus for 30minutes and rapidly quenched to room temperature with helium jets, before being held at 30C below the -transus for another 30min and cooled to room temperature at 15C/min. The fastest quenching rate for the production of Ti plates is estimated around 1000C/min.38 Because the spark plasma sintered (SPS) alloys are powder processed, there is excess oxygen when compared to the wrought alloys. The excess oxygen increases the hardness due to solid solution hardening. To compare wrought data with SPS data, the increase in hardness due to excess oxygen is calculated knowing the coefficient for oxygen solid solution strengthening is 194 kg mm-2 at%-1/2 (Table 4) and the amount of oxygen in the SPS and wrought alloys. The hardness from oxygen is calculated and subtracted from the overall hardness. The hardness (HVN) is converted to yield strength (Y.S.) with models presented in section 6.2.
Table 7: Strength and Average Lineal Intercept Measurements for Different Processing Routes Hardness reduction to Strength Ti5111 HVN O at% compensate (m) (MPa) for excess oxygen

wrought slow cooled wrought quenched SPS slow cooled* SPS quenched* *Spark Plasma Sintered

4.10 3.45 5.12 2.72

249 297 365 405

0.245 0.245 1.014 1.014

0 0 99.33 99.33

583 695 622 715

6.1.1 Lineal Intercept Measurement The average lineal intercept length ( ) of plates in these structures was determined in the doctoral research of Jamie Tran.11 Following ASTM E 112 9639, Standard Test Methods for Determining Average Grain Size, a circular intercept procedure was used to measure the average lineal intercept length of plates. As stated in the standard, the use of circular test lines rather than straight test lines automatically compensates for departures from equiaxed grain shapes, without biasing any direction. For this project the Hilliard Single-Circle Procedure is used. Several digital optical micrographs of processing route are obtained at 50X magnification where is the light phase and is the dark phase (Figure 8 left). Imaging software is then used to threshold the image to clearly distinguish phase boundaries (Figure 8 right). In this case, the phases are converted to red and phases are converted to white. A circle is randomly applied to the microscope image at 50X magnification, making sure the circle diameter is not smaller than the largest observed grains to ensure a representative distribution of sizes. Every point on the circle intersecting the phase is deleted leaving an image similar to Figure 9. Each lineal intercept length is calculated by counting the intercepts on the test line and tabulated to calculate an average. The measurement is taken as our measure of grain size in these microstructures.

23

Figure 8: (Left) Optical micrograph Ti5111 wrought fast quenched 50x magnification. (Right) Threshold image with phase colored red and phase colored white.

Figure 9: Example of circle with i ntercept region erased

6.1.2 Grain Size Modeling With the data from Table 7, yield strength (MPa) can be plotted versus average lineal intercept value -1/2 in m-1/2 (Figure 10) to determine the Hall-Petch coefficients.

Figure 10: Graph of yield strength versus average linear intercept of the plates-1/2 for Ti5111

24

These four data points give the following constants for the Hall-Petch relations:

Equation 12

Rearranging the variables

Equation 13

From the above equation, in order to increase yield strength of Ti5111 to 120 ksi or 827 MPa using only the mechanism of grain size refinement, the value would have to be reduced to 1.72 m. The Hall-Petch coefficients are calculated from plots of stresses in commercially pure (CP) Ti with grain size measurement .40 However, to compare data from this report to these values, the is converted to an value. According to ASTM E3112-9639, the conversion is . The measured 0 for CP Ti ranges from 424-715 MPa, while the value for wrought Ti5111 is lower at 281 MPa. For CP Ti the Hall-Petch coefficient k ranges from 0.136-0.233 MPa m1/2, while the calculated strengthening coefficient k for Ti5111 is 0.716 MPa m1/2. The difference in results may be due to the differing microstructures where Ti5111 has a basket weave widmanstatten microstructure and CP Ti has equaixed grains.

6.2 Solid Solution Strengthening


A contribution to the 20 ksi, or138 MPa, strength increase desired of this alloy can also be provided by increased solution strengthening of the phase. Levels of stabilizing elements such as V, Fe, and Mo, will be highly constrained by composition control of the -phase for transformation toughening, subsequently.24 Another option is increasing the content of phase substitutional elements, such as Al, Sn, and Si by a fraction of a percent, but with the consequence of reducing the corrosion resistance of the alloy. Values are available for the hardness coefficient per atomic percent of elements Al, Si, and Sn and strength coefficient per atomic percent of elements Mo, V, and Fe in Ti alloys as shown in Section 5.1.2..24 A linear proportionality between microhardness and strength was determined using Vickers hardness and yield strength measurements from the doctoral research of Jamie Tran (Figure 11). This relationship was used to convert between hardness and strength in obtaining the strength coefficients of Table 8. The effect of Zr was assumed negligible. Table 8 lists the hardness and strength coefficients for all alloying elements in Ti5111.

25

Figure 11: Yield Strength vs. MicroHardness for Ti5111

Table 8: Slopes of Solid Solution Effects of Alloying Elements in Ti24,26

Element Al Fe Mo Si Sn V

SSH slope (kg at%-1mm-2) 15* 23* 23* 39 24 9

SSS slope (*106N at%-1m-2) MPa/at% 35 54 54 91* 56* 20*

* Coefficient was converted with the empirical relationship of Figure 11

Using the above values from the literature, a model was created to estimate the hardness and strength of near Ti alloys, from solid solution hardening, from the content of V, Fe, Mo, Al, Sn, and Si approximating the alloy as 100% phase. Linear relationships and linear superposition of element effects were assumed. A spreadsheet to calculate the solid solution strength is created with the overall composition in at% as an input, shown below in Figure 12.

26

Figure 12: Model Predicting Solid Solution Strengthening

The total strength of the design alloy will be a superposition of the Hall-Petch strengthening effects with the solid solution strengthening effects41:

Equation 14

6.3 Martensitic Transformation Toughening


6.3.1 Volume Change of Martensitic Transformation in Titanium The composition dependence of the martensitic transformation volume change can be described via a Redlich-Kister equation as employed in the ThermoCalc database structure. Coefficients for the Redlich-Kister fit have been calculated using available literature data35, 42,434445,46, and calculated DFT data, described in Section 5.2.4. Because literature values of the atomic volume are only available for Ti-Mo, Ti-V, and Ti-V-Al alloys, DFT calculations are performed to estimate the atomic volumes of other elements for the binary interaction parameters with Ti. 6.3.2 DFT Atomic Volume Calculations DFT calculations are used to make first-principles predictions of the volume change. After these atomic volumes are found, the Redlich-Kister model can be fit to the data to interpolate between points. As proof of concept, preliminary DFT modeling on the volume change in the Ti-V system was shown to confirm good trend fits between DFT calculations and experimental values, albeit with a displacement (Figure 17). This displacement is not unexpected, as the DFT calculations are carried out at T=0K while the data represents 300K. In addition, there is expected to be some element of tightbinding between atoms in DFT calculations, leading to lower bulk volume. However, because the curve shape of the DFT data has strong correlation to experiment, a calibration increment can be applied to volumes uniformly to approximate the thermal expansion from 0 to 300K. The systematic increment employed is 0.62083/atom and 0.57343/atom for the and martensite phase, respectively.

27

Figure 13: Redlich-Kister Curve Fit to the Beta Phase of Experimental and DFT values for Ti-V

Table 9 and Figure 14 show values of phase atomic volume calculations for binary TiX alloys, where X is Al, Fe, Mo, Si, Sn, Ti, V, and Zr. Alloy solutions at the 25% and 50% concentrations are represented by ordered BCC structures corresponding to the A3B L21, (Heusler) and AB B2 (CsCl) structures, respectively. The heusler phase for Mo was found to be energetically unstable, and so that point was removed. Table 10 and Figure 15 show the calculations for the martensitic volume approximations with HCP phase. For HCP, only the equiatomic mixed phase was modeled. Also, calculations for pure Fe ended up being extremely unstable energetically, and so the volume data was taken from empirical measurements.
Table 9: Atomic Volume Calculations (Angstroms3/Atom)

Al Fe Mo Si Sn Ti V Zr

Heusler (Ti3X) xi = 25% 17.14 15.12 16.04 18.36 16.32 18.89

BCC-B2 (TiX) xi=50% 16.73 13.35 16.28 15.69 20.39 15.65 20.39

Pure BCC Phase xi=100% 17.74 10.87 16.28 15.39 28.43 17.47 13.81 23.52

28

Figure 14: DFT atomic volume values for Phase

Table 10: Martensite Atomic Volume Calculations (Angstroms3/Atom)

Al Fe Mo Si Sn V Zr Ti

HCP (TiX) xi=50% 17.12 13.67 16.78 15.54 20.56 15.81 20.89 -

Pure HCP Phase xi=100% 17.33 14.14 15.11 28.42 14.29 23.88 17.65

29

3 0 2 8 2 6 2 4 2 2 2 0 1 8 1 6 1 4 1 2 1 0 0 0. 2 0. 4 F e

S n Zr

A l S i M o V 0. 6 0. 8 1

Atomic Volume

X
i

Figure 15: DFT atomic volume values for Martensite Phase

6.3.3 Redlich-Kister Fitting to Atomic Volume Data All volume data used in fitting is available in the Error! Reference source not found.. Equation 15 and Equation 16 below both show the Redlich-Kister equation used to model the molar volume of both phases for Ti-V. In the equation, Vi is the molar volume of the alloy with pure component i, Xi is the atomic fraction, and is a an interaction parameter equivalent to a regular solution form. Similar equations were used for the other alloying components.

Equation 15

Equation 16

The solver function in excel is used to iterate through possible values of to find the coefficient for the least squared fit to the experimental data. The values of these parameters are given in Table 11 and Table 12.

30

Ti-Mo Atomic Volume Parameters (3) VTi 17.45 VTi 17.66 VMo 15.58 -1.76
Table 12: Fitting Constants for Ti-V.

Table 11: Fitting Constants for Ti-Mo.

VMo

14.14 -2.97

Ti-V Atomic Volume Parameters (3) VTi 17.45 VTi 17.66 VV 13.84 0.37 VV 14.29 -0.069

The available experimental data points and appropriate DFT calculations of the atomic volume of Ti-V and Ti-Mo binary alloys in both and martensitic phases are shown in Figure 16 and Figure 17 respectively. The Redlich-Kister fit for each phase is shown to compare with the data. Binary interaction parameters for all other elements with Ti are computed from DFT values.

Figure 16: Dependence of Atomic Volume in Ti-V on Composition for and Martensite Phases45

31

Figure 17: Dependence of Atomic Volume in Ti-Mo on Composition for and Martensite Phases46

Through the research of Grujicic,35 there is available ternary, Ti-Al-V, atomic volume data. This data is used to compute the binary interaction parameter and ternary interaction parameter . The form of the Redlich-Kister equation used for ternary Ti-V-Al atomic volume data is as follows:

Equation 17

where the binary interactions of Ti-V and Ti-Al are fixed from calculations with the respective binary data. This approach is taken rather than first combining all data to calculate all interaction parameters at once, as the ternary data is heavily weighted towards Ti lean sections of the Ti-Al-V ternary diagram.
Table 13: Fitting Constants for Ti-V-Al

Ti-V-Al Atomic Volume Parameters 17.45 17.66 13.84 17.74 0.37 -2.96 -4.59 -12.29 14.29 17.34 -0.07 -1.50 -3.82 -22.83

32

6.3.4 Interfacial Friction of the Martensitic Transformation in Titanium Following Equation 8 and Equation 9 discussed in Section 5.2.3, the critical free energy change for martensitic transformation can be expressed as:

Equation 18

Compiling data for the composition dependence of Ms in binary Ti alloys, the Wf interfacial friction term for martensitic transformation in Ti is found to show a linear dependence on the atomic fraction of most alloying elements agreeing with the linear relationship found for solid solution strengthening. The following graphical method is used to solve for the proportionality constant Ki (for Wf calculations) and G0. ThermoCalc is first used to generate values for Gchem. at Ms. A plot of -Gchem versus atomic fraction Xi results in a line for every element with the exception of tin (Figure 18), which follows the x0.5 parabolic behavior observed for martensitic transformation in ferrous alloys. The value of G0 is determined by averaging the y-intercepts of all two-component systems. Ki is then estimated to be the slope of a line constructed between G0 and the center of each data set. Having found G0 and Ki for each binary system, it is possible to predict the critical chemical driving force for martensitic transformation at a given atomic fraction of each alloying element.

Figure 18: Chemical driving force of transformation at Ms as a function of atomic fraction

G0 was determined to be 48 J/mol, and a table of Ki values is given in Table 14. Note that the proportionality constant for Sn satisfies Equation 19, where the work of friction

33

depends on atomic fraction to the power. The full model for the work of friction is given in Equation 20.

Equation 19

Equation 20 Table 14: Ki values of alloying elements in Ti5111

Alloying Element Zr V Fe Al Mo Si Sn

Ki (J/mol) 405 2530 4260 7640 9557 13000 3360

All Ms values for relevant binary Ti-X alloys are available except for binary Ti-Si. The Ki value for Si is estimated using an alternative method. As the mechanism of solid solution strengthening and work of friction are the same, a correlation between the known solid solution strengthening parameter to the measured Ki values is found and applied to estimate a KSi value. 6.3.5 Ms calculations and Measurement Figure 19 shows a flow chart of the design process proposed to optimize the composition, volume fraction, and dispersion for the design alloy. The optimization of the phase parameters is important in order to obtain an increase in toughness as a result of volume change upon martensitic transformation. The iterative nature of this process stems from determining an initial composition and annealing temperature, yielding phase parameters and inputs of the Olson-Cohen model, from which Ms is determined, and the process is continued until a final composition is obtained from the highlighted results shown below. To provide an important calibration point for the current Ti5111 alloy, flat dog bone tensile samples with cross section 5mm x 1mm and gauge length of 10mm were machined and tensile tested. The Bolling-Richman tensile tests for the reference Ti5111 alloy is shown in Figure 20 where the Ms temperature is estimated to be -67C. In addition this data can be used to obtain a YS vs T trendline. The Ms value for a material is dependent upon its yield strength, which is itself dependent on temperature. The reversal of the temperature dependence of the yield strength defines the Ms temperature of the phase in Ti5111 denoted by the vertical arrow in Figure 20 and Figure 21. The

34

yield stress vs temperature dependence is depicted in Figure 21 for data measured on Ti5111 1 thick plate giving a d/dT of -1.101 MPa K-1. For the design alloy with a desired strength value of 120ksi or 827 MPa at 300K, this relationship is shifted upwards to use for Ms calculations as shown in Figure 21.

Figure 19: Iterative Flow Chart for Alloy Design, Highlighted Sections Lead to Final Composition

35

Ms(ut) = -67C

Figure 20: Bolling-Richman stress-strain curves for conventionally processed Ti5111.

Figure 21: Yield Strength vs. Temperature for Ti5111: flat tensile samples with a gauge length of 10mm, width of 4.98mm, and thickness of 1.12mm47

36

6.3.6 Ms calculations for Ti5111 and Model Calibration For Ms calculations, the composition of the phase in Ti5111 is needed and can be calculated and validated by LEAP microanalysis. The nonequilibrium and phase compositions formed during nonisothermal processing can be modeled using the DIffusion Controlled TRAnsformation (DICTRA) software. DICTRA with a proper mobility database for the titanium system can solve the flux balance equations of an alloy system during various processing phases such as cooling from an annealing temperature. Using DICTRA with a mobility database previously validated by diffusion couple experiments in the senior project of Eric Simpson,48 a 1-D simulation of transformation in Ti5111 by plate thickening during the cooling process is modeled. Figure 22 compares the retained phase fraction upon cooling between DICTRA cooling simulations and equilibrium predictions of the Thermotech Ti-DATA-v3 database. The retained phase fraction from the DICTRA simulations correspond well with the 10-20% measured in conventionally processed Ti5111. An estimate of the amount of element partitioning that takes place during cooling is also modeled and compared to experimental data from the Local Electrode Atom Probe (LEAP) (Figure 23). A three dimensional image of individual atoms is obtained as the LEAP removes atoms from an atomically sharp tip. The atom probe microscope uses time-of-flight mass spectrometry to identify individual atoms, and uses point-projection microscopy to identify original atom locations in three dimensions. This technique was used in the doctoral research of Jamie Tran to analyze a Ti5111 specimen, shown in Figure 23.11

Figure 22: phase fraction with temperature from DICTRA simulations compared to equilbrium predictions

37

From composition profiles simulated by DICTRA (Figure 24 left) it is confirmed that Mo is a very strong partitioning element and can be used to define the and regions in the LEAP analysis. A proximity histogram (proxigram), a function in the IVAS atom-probe analysis software used to find the composition profiles in LEAP tomographic reconstructions, is then created (Figure 24 right).

Figure 23: Reconstruction of Ti5111 LEAP sample showing thin platelet in retained region11

Figure 24: Positional composition of Ti5111 alloy from DICTRA simulations (left) and LEAP microanalysis (right)11

38

The composition of the phase of conventionally processed Ti5111 material (Table 15) is used as an input to our transformation model to predict the Ms temperature of the alloy in uniaxial tension.
Table 15: composition measured from LEAP microanalysis in decreasing concentration composition Element (Xi: at fract) V 0.0642 Mo 0.0482 Al 0.0315 Fe 0.0111 Sn 0.0071 Zr 0.0058 Si 0.0019

Assuming a macroscopic particle size where Vp > Vo: Gn=Go=48J/mol, the Ms(ut) of Ti5111 was predicted to be 74C. As presented in section 6.3.5 Bolling-Richman49 tensile testing was preformed to measure the Ms temperature of Ti5111 with a single tensile specimen. The tensile specimen is taken into a chamber at 180C, tension is applied to create a stress-strain curve until a yield strength can be measured. The chamber temperature is then lowered 20 degrees to 160C and strained to measure the yield strength at that temperature. This process is continued to -90C. The measured Ms temperature is the temperature where the drop in yield stress with temperature is measured. As shown in Figure 21, the Ms(ut) of Ti5111 is measured to be -67C; 141C lower than that of the predicted model due to the assumption of a macroscopic particle size. This data was used as a calibration for the transformation toughening model in section 6.3.5 for the proportionality constant, K, in Equation 8 reproduced here:

Equation 21

For the model to predict an Ms temperature of -67C for Ti5111, a Gn value of 778 J/mol is needed. Since G0 remains 48, the second term in the equation is equal to 730J/mol. The reference volume V0 is 5.24x105 m3 which is the volume of a 100 m diameter sphere. Vp, is estimated from micrographs of Ti5111. The phase is modeled as a square plate with lengths of 40m, and thickness of 1m for a Vp of 1600 m3. To calibrate the model to Ti5111, Gn=778J/mol, K is equal to -4230 J/mol. This value is comparable to that evaluated for steels, when normalized by the shear modulus.34

39

Figure 25: Ti 5111 Ms as a function of annealing temperature for crack tip and uniaxial tension stress states

Figure 25 shows Ms calculations using the calibrated transformation model for both crack tip and uniaxial tension stress states vs. annealing temperatures. To achieve a room temperature Ms (300K or 27C), the Ti5111 must be annealed at 737C and 740C for crack-tip and uniaxial tension stress states respectively, until equilibrium composition is achieved and subsequent quenching will retain the appropriate -phase composition. The lower Ms temperature of the conventionally processed Ti5111 material reflects the slow cool of the alloy from a temperature of 954C at 15C/min. The conventionally processed (slow cool) Ti5111 composition is equivalent to an equilibrium composition 707C corresponding to the uniaxial tension stress state measured Ms.

7 Design Integration
The project objective, to design a Ti alloy with 120 ksi yield strength (comparable to that of Ti64), high fracture toughness and stress corrosion resistance (comparable to that of Ti5111), at a cost lower than that of Ti64, can be achieved using the models developed over a range of compositions.

7.1 Design Constraints


7.1.1 Composition constraints Initial designs employ the same alloying components as our reference alloy Ti5111. Ti5111 has seven alloying elements and the models for strengthening and transformation toughening are developed for these elements. In Table 16 a summary of the relative

40

effects of each element on , volume fraction, SCC resistance and cost is presented where the + and symbols indicate the affect on the alloy if the element is increased. Increases in all values except cost are desired.
Table 16: Summary of Alloying Element Effects

Element Al Fe Mo Si Sn V Zr Ti (1) Ti (2)

V/V ++ ++ -++

volume fraction + + + -

SCC resistance (<5wt%) + + -

Cost ($/lb) $1.20 $0.37 $13.50 $5.51 $6.32 $592.00 $13.00 $38.40 $25.60

Affect if alloying element is increased

The content of Al, Si, Sn, and Zr are first considered relative to values in Ti5111. With an increase in these alloying elements there are two parameters that are affected negatively, and two parameters that are affected positively, therefore these contents will initially not be changed. For example, an addition of Al will increase the volume change and decrease cost but will also decrease the fraction and SCC resistance. As explained in 5.2.1, the martensitic transformation in TRIP alloys inhibits the propagation of cracks on three size scales. The larger the volume change in the martensitic transformation, the greater the effect of these toughening mechanisms. As a result, positive volume change in martensitic transformation is the primary aim in toughening the alloy. According to a plot of the change in volume due to alloying element fraction in Figure 26, an increase in volume change can be achieved by increasing the amount of Fe in the phase. In addition, Mo should be reduced in the phase as it decreases the volume change. However, reducing Mo sacrifices SCC resistance. To compensate, V is increased to replace Mo to maintain SCC resistance. However, the amount of V added needs to be restricted to <4wt% due to the high cost of V. Therefore, we can initially vary Fe, Mo, and V while keeping all other alloying elements constant in preliminary designs.

41

Figure 26: Effect of atomic fraction on volume change

7.1.2 -phase fraction constraints Additionally, the goal of this design is for the -phase fraction to be similar to that in Ti5111; 10-20%. The phase provides our principal mechanism to promote higher toughness in the alloy. It is also important to limit the amount of phase because as the phase fraction increases into the + alloy region, the segregation from stabilizers is harder to control during heating processes such as welding. This is why near- alloys are considered weldable. As a result, in determining the -phase fraction in our final alloy, the goal is 15% phase. 7.1.3 Annealing temperature constraints Another constraint to the design is that Ms(ct) is set at room temperature (300K) to allow for optimal martensitic transformation at use temperature. This constraint determines the annealing temperature associated with each alloy composition. In other words, each overall design composition will have one associated annealing temperature to obtain the desired room temperature Ms(ct) value of 300K. To allow for practical designs, the annealing temperature is constrained to >600C due to the slow kinetics of alloy partitioning at low temperatures. In addition annealing temperatures <600C risk forming the embrittling Ti3Al phase. To further quantify this annealing temperature constraint, the coarsening rate for Ti5111 at the minimum annealing temperature of 600C is calculated as a reference for future designs. The rate constant for coarsening was determined using Computational Materials Dynamics (CMD)50 design interface to ThermoCalc. The Ti thermodynamic database and Ti mobility database were used in conjuncture with CMD to calculate the normalized multicomponent coarsening rate constant, i.e. K/(Vm), using the Lee model.51 The matrix phase was set to BCC and the precipitate phase to HCP. This yielded a KLee of 4.0411*10^-20 m2mol/Js. This provides a constraint to the design, because the Ti alloy 42

under consideration should have a rate constant higher than that of Ti5111, where precipitation rate constants increase with increasing temperature. 7.1.4 Mo partitioning interaction with Fe Analysis of the change in volume provided from each component give the direction of chemical change necessary to increase toughness. The two components with the largest effect on this parameter are Fe and Mo. To maximize volume change, the Fe content of the phase should be increased, while Mo of the phase should be decreased. As expected, ThermoCalc calculations showed that increasing the amount of Fe in the overall composition increased the amount of Fe in the phase. Unexpectedly, these ThermoCalc calculations also showed that increasing Fe content cause Mo to strongly partition out to the phase thereby decreasing the Mo content in the phase as shown in Figure 27. This is desirable, as decreasing the overall Mo content was previously shown to decrease the total volume fraction of the phase, moving away from the target 15% phase by volume. If this interplay between components was not present, the Mo content would have to be decreased to reduce the composition in the phase in order to increase the volume change, consequentially decreasing the volume fraction of and SCC resistance of the alloy. To compensate, V content would have been increased to achieve the 15% volume fraction of and maintain the SCC resistance of the alloy, thereby increasing the cost of the alloy. Fortunately, because of the partitioning of Mo with Fe content, changes to Mo and V content can be avoided and an efficient optimization can be achieved with an increase in Fe content.

T=730C

Figure 27: phase partitioning of elements with Fe in alloy content

43

7.2 Final Alloy Composition


7.2.1 Composition Refinement

T=730C

Figure 28: Effect of Fe in atomic percent on volume fraction of beta

Because iron has been identified as having the most significant effect on the volume fraction beta, we varied Fe in order to determine the composition that would provide a 15% volume. The data is collected by varying atomic composition of iron while holding all other alloying components constant (Figure 28). We fit a third order curve to the resulting beta volume compositions, and interpolated the point where beta volume is 15%. This value is 0.858 at% Fe. 7.2.2 Final Composition The final overall composition of the proposed design alloy is listed in Table 17. The amount of Fe in the design is about 1wt%, suggesting the design alloy name should be Ti51111 (Titanium five quadruple one).
Table 17: Final Composition of Design Alloy

Element Al Fe Mo Si Sn V Zr Ti

at% 8.67 0.86 0.44 0.17 0.39 0.92 0.51 88.04

wt% 4.99 1.02 0.90 0.10 1.00 1.00 1.00 89.98

44

7.2.3 Annealing Temperature and Ms To find the annealing temperature for our desired Ms, we generated three different phase compositions at different annealing temperatures. Figure 29 shows Ms as a function of annealing temperature for the two stress states of interest. Since our goal is transformation toughening, the crack tip stress state is of primary interest. For a room temperature transformation, corresponding to an Ms(ct) of 27C, the annealing temperature must be 717C or 990K. This corresponds to an Ms(ut) of -5C which can be easily validated experimentally.

crack tip

uniaxial tension

Figure 29: Final composition Ms as a function of annealing temperature for crack tip and uniaxial tension stress states

7.2.4 phase composition The phase composition at an annealing temperature of 717C is found using ThermoCalc. The alloy is ~14 at% phase, and the compositions of alloying elements within the phase are shown in Table 18.

45

Table 18: phase composition of the design alloy at 717C

Element Al Fe Mo Si Sn V Zr Ti

at% 6.35 5.77 2.04 0.22 0.11 2.89 0.60 82.02

wt% 3.54 6.66 4.05 0.13 0.28 3.04 1.14 81.16

The rate constant for coarsening of the designed alloy, determined using CMD at its annealing temperature of 717C, was found to be 5.45E-19 m2mol/Js. This KLee is greater than that of Ti5111 at its minimum annealing temperature, where KLee is 4.04E-20 m2mol/Js. This verifies increased precipitation rate for the newly designed alloy in comparison to Ti5111.

7.3 Processing
The proposed processing treatment of the design alloy is similar Ti5111. After homogenization of the alloy and hot rolling deformation, there is two-step annealing. In Ti5111 the first annealing temperature is 30C above the transus temperature for 30 min. Ti alloys annealed above the transus can fully recrystallize in the phase field to produce a Widmansttten microstructure upon cooling. This annealing temperature must not be too high as the grains will rapidly coarsen which leads to lower yield strengths. The transus is calculated using ThermoCalc and found to be 1212K (939oC) for the current alloy design compared to 1253K (980C) for Ti5111. Therefore the design alloy Ti51111 (Ti5111 + 1wt%Fe) is brought to 969oC, 30oC above the transition temperature, for 30 minutes. Unlike conventionally processed Ti5111, the Ti51111 design alloy will be quenched after the first annealing step to produce refined martensitic plates and brought to the second annealing step at 731C to develop the desired 15% phase fraction and composition for the fracture toughness model. After the second annealing step the alloy is cooled at 15oC/min to room temperature similar to conventionally processed Ti5111.

7.4 Strengthening Results


7.4.1 Solid Solution Strengthening The final alloy composition with Fe content of 0.86at% is input into the solid solution strengthening model, detailed in section 6.2, to determine the change in strength compared to the reference composition of Ti5111 as a result of a change in the alloying elements. Using this model the strength is found to increase by 41.83MPa. 7.4.2 Hall-Petch Grain Refinement Hall-Petch strengthening supplements SSS to achieve the end strength goal. The difference of strength between the goal of 827MPa and the strengthening obtained from 46

SSS (41.83MPa) is 785.17 MPa, the strength which must be achieved as a result of grain refinement. Using Equation 13, the lineal intercept length corresponding to this strength was found to be 2.02 m which should be achievable by fast cooling. This microstructure, along with the composition and processing previously mentioned, will achieve the desired strength of 120ksi.

7.5 Toughening Results


Having determined the design composition of the alloy and the annealing temperature needed for Ms(ct) at 300K, the Gmech, Gchem, Gn, and Wf values of the martensitic transformation model are reported here. Gmech is given by Equation 6, where is our target yield strength of 120 ksi (827 MPa). Furthermore, h= 3 for the crack tip stress state, and V/V based on our composition is 1.21%. This volume change compares to 0.4% in Ti5111, corresponding to a factor of 3 increase. These values result in a Gmech =-359 J/mol. ThermoCalc is used to determine Gchem, which is -1562 J/mol. Gn, calibrated to experimental results, is 778 J/mol. The total Wf can be found using the composition and Ki values calculated in section 6.3.4, and is 1143 J/mol. These values of Gmech, Gchem, Gn, and Wf sum to zero when Ms is at room temperature for that specific stress state. A similar procedure can be used to calculate Gmech for uniaxial tension. Table 19 shows values for both uniaxial tension and crack tip stress states. Note that the total does not sum to zero for the uniaxial stress state as the Ms temperature cannot be the same for both crack tip and uniaxial stress states.
Table 19: Martensitic Transformation Model for Uniaxial Tension and Crack Tip Stress States

Gmech (J/mol) Gchem. (J/mol) Gn (J/mol) Wf (J/mol) Total

Crack Tip -359 -1562 778 1143 0

Uniaxial Tension -219 -1645 778 1143 57

8 Conclusion and Recommendations


A new alloy, Ti51111(Ti5111+1wt%Fe) was designed, capable of a high strength of 120ksi, an increase in volume change upon martensitic transformation of the phase with optimized stability for transformation toughening, a lower cost of $29.34 to $40.86, without sacrificing corrosion resistance. This alloy was designed with the constraint of having a volume fraction of ~15%, enough to utilize the toughness provided by phase volume change, but not high enough to exceed the weldable near- phase regime. Table 20 compares important values predicted for Ti5111 and the design alloy, Ti51111.

47

Table 20: Comparison of Ti5111 and design alloy

Ms (ct) Ms(ut) phase fraction V/V

Ti5111 -59C -67C 10-20% 0.4%

Ti51111 design 27C -5C 14% 1.2%

Several models were generated and used in this design, including one of volume change, utilizing Redlich-Kister curve fits of density functional theory calculations, work of friction models employing both thermodynamic and experimental data, -partitioning models utilizing atom probe tomography data, solid solution strengthening models, HallPetch and grain size predictions, and a cost model. The use of all of these design tools allowed for all aspects of current titanium alloys to be improved: Ti51111 (Ti5111+1wt%Fe) promises greater strength, improved toughness, and lower cost, making it an attractive candidate to evaluate for naval and automotive applications. Next step recommendations include testing the prediction of Figure 25 that an optimum stability of can be achieved with higher temperature annealing. The corresponding Ms in uniaxial tension can be tested experimentally as further validation of the transformation models. The new composition can be made first as small arc melted buttons to test phase relations and predicted Ms(ut) with the single tensile test technique. Validated stability can be followed by larger scale melts for toughness tests. Arrangements are being made to conduct experiments this summer.

9 Acknowledgements
Doctoral candidate Jamie Tran provided significant data, discussion, and guidance for this work, without which such a well-explored solution would not have been possible. Professor Olson also provided significant insight and interpretation, for which we are very grateful. Also the senior project of Eric Simpson gave useful experimental insight into the microstructure and properties of the reference alloy Ti5111, and the contributions of previous 390 teams provided a solid foundation upon which this work was built.

48

10 References
1

EHKTechnologies; Summary of Emerging Titanium Cost Reduction Technologies. US Department of Energy Oak Ridge National Laboratory (2004). 2 Cyberalloys 2020: Naval Materials by Design Award #: N00014-07-1-0279 3 GM High Toughness, High Strength Titanium Wrought Alloys Made by Powder Consolidation; Contract #: P.O. TCS14233 4 U.S. Congress. P.L. 110-140: Clean Energy Act of 2007. (2007). 5 C. Kim and M.L. Holly. Low Cost Titanium for Automobiles. GM Research and Development Center 10 (2006) 618. 6 Brickey, J.; Chen, C.; Chastain, M.; Lin, F. MSE 390 Report: Ti120 Marine Titanium. Northwestern University: Evanston, IL (2008). 7 Henry, S.D.; Dragolich, K.S.; DiMatteo, N.D. Fatigue Data Book: light structural alloys ASM International (1995) 264. 8 Delmedico, R., J. Fakonas, K. Peter, G. Scott and E. Simpson MSE 390 Report: Marine Ti120 High Performance Ti Alloy. Northwestern University: Evanston, IL (2007). 9 Davis, J.R. ASTM A812 HSLA Steel Grade 80 Sheet ASM Specialty Handbook - Carbon and Alloy Steels, ASM International, Metals Park, OH, (1996). 10 CES EduPack 2008 software, Granta Design Limited, Cambridge, UK (2008). 11 Tran, J.; Research Proposal: Ti by Design. Northwestern University Evanston, IL (2007). 12 Aerospace Structural Metals Handbook, Code 3720, Battelle Columbus Laboratories 4 (1969). 13 Cahn, J. W; "Symmetry Changes Expected From Deformation Twinning and martensite Transformations." Martensite a tribute to Morris Cohen. ASM International: Metals Park, OH (1992) 97102. 14 Czyryca, E.J.; Wells, M.E.; Tran, K. Titanium Alloy Ti-5111 for Naval Applications. Advanced Marine Materials: Technology & Applications (2003), 41-49. 15 Been, J.; Davis, D. Corrosion Property Evaluation of Ti 5111in Marine Environments. NACE International Corrosion 2000 (2000), 00641.1-00641.10 16 Been, J., Titanium Alloy 5111 brings intermediate strength, excellent toughness, and corrosion resistance to Naval operating environments. Corrosion (1999) 55. 17 Boyer, R. R. a. W. F. S., Effect of Composition, Microstructure, and Texture on Stress-Corrosion Cracking in Ti-6Al-4V Sheet. Metallurgical Transactions A 9 (1978) 1443-1448. 18 Curtis, R. E.; Boyer, R. R.; Williams, J. C., Relationship between Composition, Microstructure, and Stress Corrosion Cracking (in Salt Solution) in Titanium Alloys. ASM Transactions Quarterly 2 (1969) 457. 19 Kolachev, V. A. a. V. V. T., The Influence of Alloying On The Tendency Of Titanium Alloys Toward Salt Corrosion. Fiziko-Khimicheskaya Mekhanika Materialov 17 (1981) 35-38. 20 Donachie, J. M., Selection of Titanium Alloys for Design. Handbook of Materials Selection, John Wiley and Sons, Inc.: New York, NY (2002) 201-233. 21 Jenkins, G. M.; Beishon J.; Peters G.., Systems Behavior. Open University Press (1972). 22 Meier M. The Hall-Petch Relationship Department of Chemical Engineering and Material Science, University of California, Davis (2004). 23 Conrad, H.; Narayan, J. , On the Grain Size Softening in Nanocrystalline Materials Scripta Materialia, 42 (2000) 1025-1030. 24 Collings, E. W., The Physical Metallurgy of Titanium Alloys. Materials Properties Handbook: Titanium Alloys, American Society for Metals: Metals Park, OH (1984) 94-111. 25 Hammond, C.; Nutting, J., The Physical Metallurgy of Superalloys and Titanium Alloys, in Forging and Properties of Aerospace Materials, The Metals Society (1978) 75-102. 26 Collings, E. W., The Physical Metallurgy of Titanium Alloys. Materials Properties Handbook: Titanium Alloys, American Society for Metals: Metals Park, OH (1984) 134-143. 27 Aurelio, G.; Fernandez Guillermet, A.; Cuello, G.J.; Campo, J. Metastable Phases in the Ti-V System: Part I. Neutron Diffraction Study and Assessment of Structural Properties. Metallurgical and Materials Transactions A. 33A (2002) 1307-1317.

49

28

Olson, G.B; Cohen, M. A mechanism for the strain-induced nucleation of martensitic transformations. Journal of Less Common Metals. 28 (1972) 107-118. 29 Olson, G.B. Transformation Plasticity and Toughening. Journal de Physique IV. 6 C1 (1996) 407-418. 30 Olson, G.B.; Cohen, M. Martensitic Transformation as a Deformation Process Mechanical Properties and Phase Transformations in Engineering Materials. Eds. Antalovich, S.D.; Ritchie, R.O.; Gerberich, W.W: TMS-AIME, (1986) 367-390. 31 Olson, G.B.; Cohen, M. General Mechanism of Martensitic Nucleation. 1. General Concepts and FCCHCP Transformation. Metallurgical Transactions A-Physical Metallurgy and Materials Science. 7 (1976) 1897-1923. 32 Patel, J. R.; M. Cohen, Criterion for the Action of Applied Stress in the Martensitic Transformation.Acta Metallurgica 1 (1953) 531. 33 Bhattacharjee, A.; S. Bhargava. Effect of beta grain size on stress induced martensitic transformation in beta solution treated Ti-10V-2Fe alloy. Scripta Materialia. 53 (2005) 195-200. 34 Olson, G.B.; Tsuzaki, K.; Cohen, M. Statistical Aspects of Martensitic Nucleation. Mat. Res. Soc. Symp. Proc. 57 (1987) 129-148. 35 Grujicic, M. Design of Ti-Al-V -phase for Transformation Toughening of -Titanium Aluminide. Materials Science and Engineering. A154 (1992) 75-78 36 Perdew, J. P.; Wang Y., Phys. Rev. B 45, 13 244 (1992). 37 Bolling. G.; Richman, R.; The plastic deformation-transformation of paramagnetic FCC Fe-Ni-C alloys, Acta Metallurgica 18 (1970). 38 Shong; Simon. Effect of Cooling Rate on the -Ti to -Ti Transformations in a Ti-43at%Al Intermetallic Alloy. Scripta Metallurgica 23(1989) 1181-1186. 39 ASTM, E112-96 Standard Test Method for Determining Average Grain Size (1996). 40 Ltjering, G.; Williams, J.C. Titanium Springer, Berlin ; New York, (2003). 41 Bhadeshia, H.K.D.H; Honeycombe, R.W.K. Steels: microstructure and properties. ButterworthHeinemann, (2006) 27. 42 Blackburn, M.; Williams, J, Transactions of the Metallurgical Society of AIME 242 (1968) 2461. 43 Smithells, C; Brandes, E.; Brook, G; Smithells metals reference book Butterworth-Heinemann, Oxford; Boston, (1992). 44 Davis, R.; Flower, H.M.; West, D.R.F.; Martensitic Transformations in Ti-Mo Alloys Journal of Materials Science 14 (1979) 712. 45 Aurelio,G.; Guillermet A. F.; Cuello G. J.; and J. Campo, Metallurgical and Materials Transactions Physical Metallurgy and Materials Science 33 (2002) 1307. 46 Dupouy, J. M.; Averbach B. L., Atomic Arrangement in Titanium-Molybdenum Solid Solution Acta Metallurgica 9 (1961) 755. 47 Lin, F. Senior Project. Northwestern University Evanston, IL (2009). 48 Simpson, E. Diffusion Couple Analysis for Ti5111 Alloy Northwestern University: Evanston, IL (2008). 49 Bolling, G; Richman, R. The Plastic Deformation-transformation of Paramagnetic FCC Fe-Ni-C Alloys.ACTA Metallurgica Vol 18, (1970) 50 Computational Materials Dynamics, Questek, (2007). 51 Lee, Hyuck M., Samuel M. Allen, and Mica Grujicic. "Stabiliity and Coarsening Resistance of M2C Carbides in Secondary Hardening Steels." Proceedings of the 34th Sagamore Conference. Innovations in Ultrahigh-Strength Steel Technology, Lake George, New York. 127-46.

50

Вам также может понравиться