effect of liquid metal distribution on the flow field …

84
Purdue University Purdue e-Pubs Open Access eses eses and Dissertations January 2015 EFFECT OF LIQUID METAL DISTRIBUTION ON THE FLOW FIELD AND MACROSEGREGATION DURING DIRECT CHILL CASTING John Steven Coleman Purdue University Follow this and additional works at: hps://docs.lib.purdue.edu/open_access_theses is document has been made available through Purdue e-Pubs, a service of the Purdue University Libraries. Please contact [email protected] for additional information. Recommended Citation Coleman, John Steven, "EFFECT OF LIQUID METAL DISTRIBUTION ON THE FLOW FIELD AND MACROSEGREGATION DURING DIRECT CHILL CASTING" (2015). Open Access eses. 1052. hps://docs.lib.purdue.edu/open_access_theses/1052

Upload: others

Post on 08-Apr-2022

1 views

Category:

Documents


0 download

TRANSCRIPT

Purdue UniversityPurdue e-Pubs

Open Access Theses Theses and Dissertations

January 2015

EFFECT OF LIQUID METAL DISTRIBUTIONON THE FLOW FIELD ANDMACROSEGREGATION DURING DIRECTCHILL CASTINGJohn Steven ColemanPurdue University

Follow this and additional works at: https://docs.lib.purdue.edu/open_access_theses

This document has been made available through Purdue e-Pubs, a service of the Purdue University Libraries. Please contact [email protected] foradditional information.

Recommended CitationColeman, John Steven, "EFFECT OF LIQUID METAL DISTRIBUTION ON THE FLOW FIELD AND MACROSEGREGATIONDURING DIRECT CHILL CASTING" (2015). Open Access Theses. 1052.https://docs.lib.purdue.edu/open_access_theses/1052

Graduate School Form 30Updated 1/15/2015

PURDUE UNIVERSITYGRADUATE SCHOOL

Thesis/Dissertation Acceptance

This is to certify that the thesis/dissertation prepared

By

Entitled

For the degree of

Is approved by the final examining committee:

To the best of my knowledge and as understood by the student in the Thesis/Dissertation Agreement, Publication Delay, and Certification Disclaimer (Graduate School Form 32), this thesis/dissertation adheres to the provisions of Purdue University’s “Policy of Integrity in Research” and the use of copyright material.

Approved by Major Professor(s):

Approved by:Head of the Departmental Graduate Program Date

John S. Coleman

EFFECT OF LIQUID METAL DISTRIBUTION ON THE FLOW FIELD AND MACROSEGREGATION DURING DIRECTCHILL CASTING

Master of Science in Materials Science Engineering

Matthew J.M. KraneChair

David R. Johnson Co-chair

Kevin P. TrumbleCo-chair

Matthew J.M. Krane

David F. Bahr 10/16/2015

i

i

EFFECT OF LIQUID METAL DISTRIBUTION ON THE FLOW FIELD AND

MACROSEGREGATION DURING DIRECT CHILL CASTING

A Thesis

Submitted to the Faculty

of

Purdue University

by

John S. Coleman

In Partial Fulfillment of the

Requirements for the Degree

of

Master of Science in Materials Engineering

December 2015

Purdue University

West Lafayette, Indiana

ii

ii

This work is dedicated to my family

iii

iii

ACKNOWLEDGEMENTS

I would like to thank my family, friends, and colleagues for their support and guidance

throughout this process. First, I would like to thank my advisor Dr. Matthew J.M. Krane,

who has mentored me throughout my academic and professional development. I am

forever grateful for the opportunities that he has provided me. Second, I would like to

thank Kyle Fezi and Alex Plotkowski for their patience and willingness to help progress

my understanding of the numerical methods and physical sciences used in this work. This

work would not have been possible without them. Finally, I would like to thank my

mother, Dorothy, and my father, Steve, for their endless love, support, and belief in me.

The funding of this research by Shandong Nanshan Aluminum Co., Beijing Nanshan

Institute of Aeronautical Materials is gratefully acknowledged.

iv

iv

TABLE OF CONTENTS

Page

LIST OF TABLES ............................................................................................................. vi

LIST OF FIGURES .......................................................................................................... vii

LIST OF SYMBOLS ......................................................................................................... xi

ABSTRACT ..................................................................................................................... xiii

CHAPTER 1. INTRODUCTION .................................................................................... 1

1.1 Direct Chill Casting ................................................................................................... 1

1.1.1 Process Description ......................................................................................... 1

1.1.2 Literature Review of Liquid Metal Distribution in DC Casting ...................... 7

1.2 Research Objectives ................................................................................................ 13

CHAPTER 2. NUMERICAL MODEL DESCRIPTION ............................................... 14

2.1 Fluid Mechanics ...................................................................................................... 16

2.2 Energy Conservation ............................................................................................... 18

2.3 Species Conservation .............................................................................................. 19

2.4 Solidification Model ................................................................................................ 20

2.5 Domain .................................................................................................................... 21

2.6 Boundary Conditions ............................................................................................... 26

2.7 Summary of Numerical Model ................................................................................ 29

CHAPTER 3. RESULTS AND DISCUSSION ............................................................. 30

3.1 Base Case ................................................................................................................ 32

3.2 Effect of Casting Speed ........................................................................................... 41

v

v

Page

3.3 Effect of Distribution System .................................................................................. 48

3.4 Effect of Nozzle Diameter ....................................................................................... 56

3.5 Results Summary ..................................................................................................... 61

CHAPTER 4. SUMMARY AND FUTURE WORK..................................................... 62

4.1 Summary ................................................................................................................. 62

4.2 Future Work ............................................................................................................ 63

LIST OF REFERENCES .................................................................................................. 64

vi

vi

LIST OF TABLES

Table .............................................................................................................................. Page

Table 2.1: Chemical Specification Range of Aluminum Alloy 7050 From ASTM B247 20

Table 2.2: Thermophysical Properties of Aluminum Alloy 7050 .................................... 21

Table 3.1: Summary of the Parametric Study Cases ......................................................... 30

vii

vii

LIST OF FIGURES

Figure ............................................................................................................................. Page

Figure 1.1: Schematic of the vertical DC casting process .................................................. 1

Figure 1.2: Typical composition profile for a grain refined DC cast billet. ....................... 3

Figure 1.3: Typical mold assembly for DC casting for the (a) bi-level transfer method and

(b) level pour method [12] .................................................................................................. 6

Figure 1.4: Geometry of melt feeding distributors for the (a) semi-horizontal melt feeding

scheme and (b) vertical melt feeding scheme. .................................................................... 9

Figure 2.1: Schematic of the numerical domain representing a hot top DC caster with a

submerged nozzle.............................................................................................................. 22

Figure 2.2: Schematic of the numerical domain for the (a) combo bag and (b) flow

diffuser plate ..................................................................................................................... 24

Figure 2.3: Schematic of the control volume expansion and splitting method used to grow

the domain: (a) start of expansion, (b) during expansion, (c) just before splitting, and (d)

after splitting. .................................................................................................................... 25

Figure 2.4: Schematic of the velocity boundary conditions used in the numerical model.

........................................................................................................................................... 26

Figure 2.5: Schematic of the thermal boundary conditions in the hot top mold ............... 27

viii

viii

Figure Page

Figure 3.1: For Case 1 (base case), the temperature and flow fields are shown on the left

and the mixture composition, the solidus and liquidus lines (solid), and the packing

interface (dashed) are shown on the right. The solid streamlines show the clockwise flow

with 0.1< ρΨ < 1 and ΔρΨ=0.1 kg/s and dotted streamline show the counter-clockwise

flow with -0.005< ρΨ < -0.000025 and ΔρΨ=0.0004975 kg/s. ........................................ 33

Figure 3.2: Close-up view of the rigid mushy zone for both (a) the ingot frame of

reference with -0.000025< ρΨ <-0.005 and ΔρΨ=0.0004975 kg/s and (b) the mold frame

of reference with -0.1< ρΨ < -0.001 and ΔρΨ=0.001 kg/s. The temperature and flow

fields are shown on the left and the mixture composition, the solidus and liquidus lines

(solid), and the packing interface (dashed) are shown on the right. ................................. 36

Figure 3.3: Axial component of the solid velocity along the bottom of the sump for the

base case............................................................................................................................ 37

Figure 3.4: Predicted composition deviation of each component for the 30mm/min base

case 1.5 m from the bottom block ..................................................................................... 39

Figure 3.5: Steady state volume weighted composition distributions for Case 1 (base case)

of (a) Zn, (b) Cu, and (c) Mg . .......................................................................................... 40

Figure 3.6: Effect of casting speed through a submerged nozzle for (a) 30 mm/min and (b)

60 mm/min. The temperature and flow fields are shown on the left and the mixture

composition, the solidus and liquidus lines (solid), and the packing interface (dashed) are

shown on the right. The solid streamlines show the clockwise flow with 0.1< ρΨ < 1 and

ΔρΨ=0.1 kg/s and dotted streamline show the clockwise shrinkage driven flow with -

0.005< ρΨ < -0.000025 and ΔρΨ=0.0004975 kg/s........................................................... 42

ix

ix

Figure Page

Figure 3.7: Close up of the rigid mushy zone for a casting speed of 60 mm/min, showing

the two phases of flow field oscillation: (a) entrainment by the thermosolutal bouyancy

and (b) penetration into the slurry region. The temperature and flow fields are shown on

the left and the mixture composition, the solidus and liquidus lines (solid), and the

packing interface (dashed) are shown on the right. .......................................................... 43

Figure 3.8: Comparison of the rigid mushy zone profiles for casting speeds of 30 and 60

mm/min. ............................................................................................................................ 45

Figure 3.9: Normalized radial Zn segregation profile at 1.5m from the bottom block,

comparing the effects of casting speed. ............................................................................ 46

Figure 3.10: Steady state volume weighted composition distributions for Cases 1 and 2 of

(a) Zn, (b) Cu, and (c) Mg ................................................................................................ 47

Figure 3.11: Effect of a combo bag for Vcast = (a) 30 and (b) 60 mm/min. The

temperature and flow fields are shown on the left and the mixture composition, the

solidus and liquidus lines (solid), and the packing interface (dashed) are shown on the

right. .................................................................................................................................. 51

Figure 3.12: Effect of diffuser plate for Vcast = (a) 30 and (b) 60 mm/min. The

temperature and flow fields are shown on the left and the mixture composition, the

solidus and liquidus lines (solid), and the packing interface (dashed) are shown on the

right. .................................................................................................................................. 52

Figure 3.13: Steady state volume weighted composition distributions for Cases 1,3, and 5

of (a) Zn, (b) Cu, and (c) Mg . .......................................................................................... 54

x

x

Figure Page

Figure 3.14: Steady state volume weighted composition distributions for Cases 2,4, and 6

of (a) Zn, (b) Cu, and (c) Mg . .......................................................................................... 55

Figure 3.15: Plot of the normalized velocity as a function of the nozzle to billet area ratio

........................................................................................................................................... 56

Figure 3.16: Effect of nozzle diameter for Vcast = 30 mm/min and the distribution

systems: (a) nozzle only, (b) combo bag, and (c) diffuser plate. The temperature and flow

fields are shown on the left and the mixture composition, the solidus and liquidus lines

(solid), and the packing interface (dashed) are shown on the right. ................................. 59

Figure 3.17: Steady state volume weighted composition distributions for Cases 1,8, and 9

of (a) Zn, (b) Cu, and (c) Mg . .......................................................................................... 60

xi

xi

LIST OF SYMBOLS

Latin Symbols

cp Specific heat (J/kg K)

C Composition (mass fraction)

D Mass diffusivity (m2/s)

f Phase mass fraction

g Phase volume fraction

k Thermal conductivity (W/m K)

K Permeability (m2)

P Pressure (N/m2)

T Temperature (K)

u Velocity in z direction

v Velocity in r direction

𝑉 Velocity vector (m/s)

Greek Symbols

β Expansion coefficient

Secondary dendritic arm spacing (m)

μ Dynamic viscosity (kg/m s)

Density (kg/m3)

Φ Generic transport variable

Ψ Stream function

xii

xii

Subscripts

l Liquid

s Solid

S Solutal component

T Thermal component

xiii

xiii

ABSTRACT

Coleman, John S. M.S.M.S.E., Purdue University, December 2015. Effect of Liquid

Metal Distribution on the Flow Field and Macrosegregation During Direct Chill Casting.

Major Professor: Matthew J.M. Krane.

A fully transient, 2-D axisymmetric numerical model is used to investigate the

effects of liquid metal distribution on the flow field and macrosegregation during direct

chill (DC) casting. The model is applied to 500 mm diameter billets of aluminum alloy

7050 cast at 30 and 60 mm/min. The predicted flow fields and compositions are

compared for three metal feeding systems; one containing only a submerged nozzle,

another with the addition of a rigid combo bag, and a final with the addition of a flow

diffuser plate. The effects of nozzle diameter, casting speed, and liquid metal distribution

system type are studied. The results from the numerical model show that the flow field

and macrosegregation are strongly influenced by casting speed and nozzle exit velocity.

At a 60 mm/min casting speed, oscillations in the flow field and composition banding in

the axial direction develop due to the increased competition between the nozzle flow and

the thermosolutal buoyancy. For nozzle to billet area ratios less than 5%, the nozzle exit

velocity is large enough to jet down the centerline, resulting in higher levels of

macrosegregation. The addition of either a low permeability combo bag or a flow diffuser

plate is shown to prevent oscillations at higher casting speeds and jetting of the nozzle

flow down the centerline for low nozzle to billet area ratios.

1

1

CHAPTER 1. INTRODUCTION

1.1 Direct Chill Casting

1.1.1 Process Description

Semi-continuous direct chill (DC) casting is the primary casting method for

wrought aluminum alloys. Presently, more than 90 percent of wrought aluminum alloy is

cast using this technology. [1,2] Typical DC cast products include rectangular ingots that

are further rolled into plate, sheet, and foil, and cylindrical billets that are further

extruded into rods, bars, tubes, and wires, or forged into complicated shapes. A schematic

of the vertical DC casting process is presented in Figure 1.1.

Figure 1.1: Schematic of the vertical DC casting process

2

2

During DC casting, hot liquid metal is fed into a water-cooled mold that is

initially closed with a bottom block. Once the liquid metal freezes on the bottom block

and a solid shell is formed near the mold walls, the bottom block is lowered into a casting

pit using a hydraulic ram. This first stage of solidification is known as primary cooling.

At this stage, the inner core of the billet still has semi-solid and liquid regions. In order to

freeze the inner core of the billet, the solid shell is quenched directly with water jets as

the billet descends out of the bottom of the mold. This secondary cooling stage accounts

for 80-90% of the total heat extraction during the DC casting process. [1,3] Casting stops

when the hydraulic ram reaches the bottom of the casting pit, making the process semi-

continuous.

The addition of TiB2 grain refiner during DC casting allows for the nucleation of

equiaxed grains, which improve the uniformity in mechanical properties of the billet and

reduces hot cracking susceptibility.[4,5] These equiaxed grains form a slurry of solid

particles suspended in the liquid metal. As the particles grow and interact, they connect to

form a permeable rigid solid structure. The demarcation between the slurry of solid

particles and the permeable rigid solid structure is represented by a critical solid fraction,

typically 0.3, and is referred to as the packing interface. [4,6,7] Shown schematically in

Figure 1.1, the region between the liquidus isotherm and the packed interface is known as

the slurry zone, and the region between the packed interface and solidus isotherm is

known as the rigid mushy zone.

A major irreversible defect that occurs in DC casting is macrosegregation, defined

as the spatial variation of alloying elements on the length scale of the billet. [4,7]

Macrosegregation impairs the quality of the billet through nonuniform mechanical

3

3

properties, caused by inhomogeneous solid solution strengthening and secondary phase

fractions. Unlike microsegregation, macrosegregation cannot be alleviated by subsequent

heat treatments, such as homogenization, so it is important to control during the casting

process.

A commonly observed radial macrosegregation profile in a DC cast billet reveals

positive (solute-rich) segregation at the billet surface, and negative (solute-depleted)

segregation at the billet centerline, as shown in Figure 1.2. [7] Enrichment near the billet

surface is primarily attributed to exudation of solute-rich liquid as the partially solidified

shell shrinks and pulls away from the mold. [8] Macrosegregation over the rest of the

billet radius is a result of the fluid motion during solidification. This motion is influenced

by thermosolutal buoyancy, solidification in the mushy zone, free floating solute depleted

particles, and forced convection from the liquid metal distribution system.

Figure 1.2: Typical composition profile for a grain refined DC cast billet.

4

4

Thermosolutal buoyancy is attributed to the uneven cooling associated with DC

casting resulting in temperature and concentration gradients in the solidifying billet.

Temperature differences between the center and outer surface of the billet cause a density

difference in the liquid. The liquid cooled at the outer radius contracts and acquires the

momentum that causes it to sink and the liquid in the center to rise. In the all liquid

region, only thermal convection is active. Below the liquid isotherm however, some solid

is formed and the consequent partitioning of alloying elements produces variations in

liquid composition, altering the local density and adding a solutal component to the

buoyancy driven convection. [4] In general, thermosolutal convection will contribute to

positive centerline segregation for alloying elements with a partition coefficient less than

one.

Shrinkage during solidification due to the density change from liquid to solid also

causes liquid flow. In the slurry region, the shrinkage-driven flow is compensated by the

melt flow and has little effect. However, in the rigid mushy zone solidification causes a

volume deficit that is filled by the motion of liquid towards the solid surfaces. Although

the velocity of shrinkage-driven flow is much smaller than the flow in the slurry, it

involves the transport of highly enriched liquid. Shrinkage-driven flow is perpendicular

to the solidification front, so any horizontal component transports solute away from the

center of the billet, resulting in depletion at the centerline. [9,10] As the sump deepens

and the solidification front becomes more vertical, negative centerline segregation

increases.

Free-floating equiaxed grains form just below the liquidus temperature in the

grain-refined liquid. These solute-depleted grains are transported by the thermosolutal

5

5

convection along the rigid mushy zone interface where they settle. Yu and Granger [11]

and Jacoby and Chu [12] credit the movement and sedimentation of solute-depleted free-

floating grains forming near the top, outer region of the slurry for the appearance of a

duplex structure at the center of a DC cast ingot, and the associated negative centerline

macrosegregation. Various other studies have attributed the transport of free floating

solid grains to an overall increase in negative centerline segregation. [13–17]

Forced convection from the liquid metal distribution system affects the flow field

in the liquid and slurry zones, and therefore the transport of solute-rich liquid and solute-

depleted free-floating grains in the solidifying billet. In DC casting, liquid metal is fed to

the mold by either the level-pour method or the bi-level transfer method, shown

schematically in Figure 1.3. In the level pour method, liquid metal is transferred to the

mold directly from a trough that is level with the top of the mold. In the bi-level transfer

method, liquid metal is transferred to the mold by a nozzle extending from an elevated

trough. Due to the constriction of flow through a nozzle in the bi-level method, the

magnitude of the forced convection is greater than in the level-pour method. The level of

macrosegregation in a DC cast billet has been attributed to the magnitude of convection

in the liquid and slurry zones, though the mechanisms describing this phenomenon are

debated. [11,12] In order to damp or redirect the forced convection from the nozzle, a

permeable fiberglass cloth known as a combo bag is typically placed around the nozzle.

[18–22]

6

6

Figure 1.3: Typical mold assembly for DC casting for the (a) bi-level transfer method and

(b) level pour method [12]

The process parameters determine the thermal conditions and flow field of the

solidifying billet, so it is important to understand the sensitivity of the process to a certain

parameter. The controllable process parameters in DC casting are the cooling water flow

rate, level of liquid metal superheat, casting speed, and the design of the liquid metal

distribution system. Out of these parameters only the casting speed and the design of the

liquid metal distribution system have a significant effect on the thermal conditions and

flow field. [4] While casting speed controls the flow field indirectly by influencing the

thermal gradients in the billet, liquid metal distribution directly affects the flow field by

introducing forced convection. Therefore, proper design of the liquid metal distribution

system is a critical parameter for controlling the fluid flow and macrosegregation during

DC casting

7

7

1.1.2 Literature Review of Liquid Metal Distribution in DC Casting

For a general overview of the DC casting process, the reader is referred to Eskin’s

book, “Physical Metallurgy of Direct Chill Casting of Aluminum Alloys” [2], and

Grandfield’s book, “Direct-Chill Casting of Light Alloys: Science and Technology”. [1]

The scope of this study requires a more specific literature review focused on the effects

of liquid metal distribution on the flow field and macrosegregation during DC casting. In

the following section, descriptions of the previous work related to liquid metal

distribution in DC casting are presented in order to develop a better understanding of the

process and to establish the need for the research presented in this study.

The effect of a submerged nozzle on the flow field during DC casting was

investigated by Flood et al. [23] using a numerical model to compare the bi-level transfer

method to the level pour method. The numerical model used in this study was a 2-D

axisymmtric modified version of the CFD package PHOENICS. A 550 mm diameter

cylindrical billet of the binary aluminum-copper alloy, Al-4 wt.% Cu, was modeled for

two casting speeds, 45 and 50 mm/min, and a fixed inlet temperature of 968K. The flow

through the nozzle for the bi-level transfer method was restricted to 10% of the billet

diameter, which resulted in an entry velocity that was 100 times greater than the entry

velocity of the level pour method. The authors found that at low entry velocitites,

bouyancy effects dominated the flow pattern, producing flat isotherms and a recirculating

bouyancy flow cell in the liquid. As the entry velocity increased, so did the competition

with the thermally driven bouyancy. When the entry velocity was large enough, the

thermal bouyancy was overriden. The result was the penetration of the rigid mushy zone

by the nozzle flow, which produced both a different sump shape and a significant change

8

8

in the slurry flow field. While the authors provided insight into the effects of the entry

velocity on the flow field, no insight into the resulting macrosegregation was provided.

Furthermore, the authors did not predict transient effects and the solidification model was

limited to a binary alloy.

The effect of a submerged nozzle on the macrosegregation and microstructure

during DC casting of 1270 x 406 mm Al-Zn-Mg-Cu alloy rectangular ingots was

expirementally observed by Chu and Jacoby [12]. Similar to Flood et al. [23], the bi-level

transfer method was compared to the level pour method. The results show that the level

of centerline segregation is strongly affected by the magnitude of the convection in the

slurry. Due to the increased forced convection in the bi-level transfer method, a 15%

increase in negative centerline segregation compared to the level pour method was

observed. Unfortunately, the authors do not provide the dimensions of the nozzle, which

controls the strength of the forced convection. Also, because no numerical work was

performed, the mechanics of the flow field could only be speculated.

A combined numerical and experimental investigation on the effect of the

orientation of forced convection on the flow field, macrosegregation, and microstructure

of 315 mm diameter cylindrical billets of aluminum alloy 7050 was conducted by Zhang

et al. [24]. The authors compared a vertical feeding scheme to a semi-horizontal feeding

scheme (Figure 1.4) and concluded that the vertical melt feed scheme caused a jetting

flow from the nozzle that penetrated the rigid mushy zone, resulting in a cliff-shaped

sump at the centerline of the billet. While the numerical model presented was able to

predict the flow field and macrosegregation for different orientations of forced

9

9

convection and the results were experimentally validated, the transient flow field was not

investigated.

Figure 1.4: Geometry of melt feeding distributors for the (a) semi-horizontal melt feeding

scheme and (b) vertical melt feeding scheme.

In order to diffuse the downward momentum from a submerged nozzle, Gruen et

al. [25] proposed surrounding the nozzle with a distributor bag. A distributor bag, also

referred to as a combo bag, is made of a permeable material that diffuses the incoming

nozzle flow and traps inclusions that may have formed in the trough feeding the liquid

metal. To investigate the influence of the distributor bag on the temperature gradients and

sump shape during DC casting of aluminum alloy 1050 rectangular ingots, a steady state

3-D coupled turbulent fluid flow, heat transfer and solidification model was used. Two

dimensions of rectangular ingots were studied, 1730 x 660 m and 1200 x 600 mm, and

two different aspect ratio bags were simulated for the casting speeds 40 and 60 mm/min.

The authors report that the distributor bag diffused the nozzle momentum, and that the

fluid pattern and size of the mushy zone were both goverened by the apect ratio of the

distributor bag. While the predicted velocity profiles from the CFD model were

10

10

compared qualitatively to an isothermal water model and found to be in good agreement,

the effects of natural convection were not included in the model. Because DC casting is

not an isothermal process, water models do not realistically predict the physics of the

flow field.

Ilinca et al. [22] expanded upon the previous investigations of combo bags by

including the effects of natural convection using a 3-D finite element model. The model

domain and material properties were taken from Tremblay et. al [21], who modeled the

effects of a Thermally Formed (TF) combo bag on an aluminum alloy 5182 rectangular

ingot with the dimension 596 x 1660 mm. The rigid TF combo bag was introduced by

Pyrotek in 2002, and claims to produce a less turbulent flow field in the ingot resulting in

more consistent castings, better surface finish, and less inclusions compared to flexible

combo bags. [20] While Tremblay et al. attempted to study the effect of the permeability

of fabric cloth on the distribution pattern in the metal, computational cost only allowed

for 22 seconds of process simulation. Ilinica et al. showed that while natural convection

drove the flow outside of the combo bag, the forced convection still affected the flow

field in a narrow region around the combo bag which influenced the temperature

distribution in the upper portion of the pool where defects are most likely to occur. Thus,

the authors concluded that a numerical model including both natural convection and

forced convection is necessary to analyze the flow field during DC casting with a combo

bag.

Another method for preventing the downward jetting of flow from the nozzle is

the use of a floating (steel) diffuser plate. The diffuser plate redirects the incoming flow

toward the outer surface of the ingot. Reddy and Beckermann [9] used a two-phase model

11

11

developed by Ni and Beckermann [26] to investigate the flow field and macrosegregation

during DC casting of a 500 mm diameter cylindrical billet of Al-4.5 wt.% Cu. The model

includes the effects of thermosolutal bouyancy, shrinkage-driven flow in the rigid mushy

zone, heat transfer, and solute macrosegregation. The authors showed that as the liquid

metal enters the domain through the inlet it is rediected by the diffuser plate and flows

along the upper liquid surface toward the mold. Although the scope of the paper was to

investigate the grain density and permeability in the mush, the entrainment of the nozzle

flow into the thermosultal bouyancy flow cell was demonstrated. The concept of using a

diffuser plate will be further investigated in this study.

In the most comprehensive experimental study on the effects of liquid metal

distribution during DC casting to date, Gariepy and Caron [27] investigated the effect of

grain refinement, the design of the metal distribution system, and casting speed on the

extent of negative centerline segregation in DC cast slabs of aluminum alloy 3104 and

5182. The authors conducted 51 different trials and concluded that grain refining

practices and metal distribution system design had the strongest influence on the extent of

centerline segregation for a given alloy cast in a narrow range of casting speeds. The

authors also concluded that distribution systems such a combo bags and channel bags

reduce the extent of bulk segregation in bi-level transfer systems, in some cases by 30%

to 40%. It is noted however, that the experimental study did not allow for the

visualization of flow fields so the mechanisms by which different metal distribution

systems would inhibit segregation, was not determined.

From the presented literature review, it is apparent that numerous attempts have

been made to understand the mechanisms leading to macrosegregation for different liquid

12

12

metal distribution systems in DC casting. It is well understood that flow through a nozzle

in the bi-level method increases the forced convection in the ingot compared to level pour

methods. It is also understood that distribution systems such as combo bags and diffuser

plates prevent the nozzle flow from jetting down the centerline of the ingot. While

experimental studies of full scale castings provide data regarding the macrosegregation

for different liquid metal distribution systems, they are expensive, time consuming, and

do not allow for the visualization of the flow field. In order to experimentally visualize

the flow fields for different metal distribution systems, isothermal water models with

particle tracking have been used. This method does not replicate the physics of the DC

casting process however, as the effects of natural convection are not included. To

overcome the restrictions of experimental studies, numerical models have become

standard to investigate the flow field and macrosegregation of DC casting.

13

13

1.2 Research Objectives

High strength, precipitation strengthened aluminum alloy 7050 is desired for

aerospace applications requiring the best combination of strength, stress corrosion

cracking resistance and toughness. [28] Similar to other 7XXX series aluminum alloys,

alloy 7050 is recognized as difficult to cast to due to its susceptibility to hot cracking.

This is a result of the relatively high concentration of Cu in alloy 7050 impairing hot

tearing resistance by widening the rigid mushy zone. [5] To prevent hot cracking, alloy

7050 ingots are typically heavily grain refined, which increases macrosegregation in the

billet. [4,5,27] Combined with the segregation tendencies of Zn, Cu, and Mg in Al,

macrosegregation is commonly observed at relatively high levels in alloy 7050 ingots.

Macrosegregation in DC cast billets is an irreversible and undesirable

phenomenon that results in spatial variation of metal quality and mechanical properties.

In order to examine the effect of liquid metal distribution on macrosegregation during the

DC casting process, a transient numerical model with coupled momentum, energy, and

composition equations is used. While it has been show that modifying the design of the

liquid metal distribution system can improve macrosegregation, the fluid mechanics of

the system are still not fully understood. To the best of the author’s knowledge no other

numerical work in the present literature investigates the effect of metal distribution

system design, including a permeable combo bag, on the macrosegregation present in

high solute DC cast billets.

14

14

CHAPTER 2. NUMERICAL MODEL DESCRIPTION

This chapter provides a description of the numerical model used to simulate the

transient solidification process of aluminum alloy 7050 billets during DC casting. The

model description includes the governing equations, numerical methods, domain, and

boundary conditions. The following assumptions were considered when developing the

numerical model:

A fixed Eulerian coordinate system in the ingot frame of reference

Local thermodynamic equilibrium exists during solidification

All fluid flow is laminar

Flow in the mushy zone acts as flow through a porous medium

The conservation equations used in this model are based on the continuum

mixture approach developed by Bennon and Incropera. [29,30] The general advection-

diffusion equation for a single phase is:

i i i i i i iV St

(2.1)

where i denotes the phase, is the density, V is the velocity field, is the general

transport quantity, Γ is the diffusion coefficient, and S denotes any source terms.

15

15

The continuum equation is obtained by summing equation (2.1) over each phase:

i i i i i i i i i i i

i i i i

g g V g g St

(2.2)

where g is the volume fraction of each phase. Equation (2.2) can be rearranged to the

general advection-diffusion form for use in standard solver algorithms:

V St

(2.3)

where i i

i

f is mass averaged mixture transport quantity, and S is the summation

of the advection and diffusion source terms for each phase, plus any terms created in the

derivation of equation (2.3) from equation (2.2).

16

16

2.1 Fluid Mechanics

The conservation equations for mass (2.4) and momentum (2.5, 2.6) are taken

from the axisymmetric continuum mixture model of Vreeman et al. [6]. The source terms

acting on the momentum equations include the thermal and solutal buoyancy, the

restriction of flow through the rigid mushy zone due to dendritic solidification, and the

effect of free-floating solid in the slurry zone. The momentum equations are solved on a

staggered grid using the SIMPLER algorithm [31] and upwind differencing for the

advection terms.

0Vt

(2.4)

l l

l l

u Vu u ut

s sl s l s l s s

l l

f fu u u g

ss s s s s z

l

f Pg u V V u u

f z

(2.5)

2l l

l l

vv Vv v

t r

2

s sl l l

l l l

f f vv v

r

2

sl l s s s s s

l

f v Pv v g g v g

r r

(2.6)

17

17

In this model, solid is assumed to nucleate as free-floating solid grains, forming a

slurry of particles suspended in the liquid metal. As the particles grown and interact, the

volume fraction solid, gs , increases to a critical value known as the packing fraction,

gs,crit, at which point a connected permeable structure is formed. The packing fraction is

assumed to be 30% in this model. [6,7] In the slurry region, (gs < gs,crit), the effect of free-

floating solid is included in the mixture equations. The relative vertical velocity of the

solid and liquid in the slurry region is related by Stokes law,

gd

gVV ls

m

sls

2

18

1

(2.7)

where d is the average particle diameter and μm is the mixture viscosity. This

approximation assumes the equiaxed dendrites are uniform solid spheres of a specified

diameter.

In the rigid mushy zone (gs > gs,crit), the free-floating solid grains pack into a

connected permeable structure, where there is no solid motion (us=vs=0), and the terms

accounting for the transport of solid are removed from equations (2.5) and (2.6). To

represent the liquid drag in the rigid mushy zone, the terms uK l

and v

K s

are added to

equations (2.5) and (2.6), respectively. The resulting momentum equations in the rigid

mushy zone for the axial and radial directions are:

z

PBuu

KuuVu

tzs

l

l

l

l

,

(2.8)

r

P

r

vv

KvvVv

tl

l

l

l

l

l

2

(2.9)

18

18

where K is the permeability, modeled using the Blake-Kozeny equation,

2

322 1

180s

s

g

gK

,

which is a function of the secondary dendrite arm spacing (2 ) and the volume fraction

solid ( sg ).

Thermosolutal buoyancy is included in the mixture equations as a source term in

the axial direction, ρBz. The expanded form of the buoyancy source term is expressed in

equation (2.10), as a function of the differences in temperature and composition for both

the solid and liquid phases.

, 0 , 0

i i

z s s l s s T s S s sg g g g T T C C

, 0 , 0

i i

l l T l S l lg g T T C C

(2.10)

2.2 Energy Conservation

The energy conservation equation is based on the mixture temperature:

p p l fc T c TV k T f Lt t

l f s pl ps f sf L V f c c T L V V

(2.11)

The first three terms represent the sensible energy storage, advection, and conduction,

respectively. The last two terms are advection-like source terms, products of the mixture

conservation derivation. [29,30] The remaining term is the transient latent heat generation

term which is linearized using the general source-based method proposed by Voller and

Swaminathan [32] for use in the finite volume discretization described by Patankar [31],

where p is the specified control volume, nb denotes the neighbors of the (2.12)

19

19

specified control volume, o denotes values from the previous time step, and S denotes the

source terms. This method was extended for multicomponent solidification by Fezi et al.

[33] Assuming the liquid and solid specific heats are equal, the source terms for the latent

heat generation term are

and

where F(T) is the liquid fraction as a function of temperature, and m

lfF 1 is its inverse

at the previous time step. The development of the solution techniques with this latent heat

release term is further described by Fezi et al. [33]

2.3 Species Conservation

A species conservation equation is provided for N-1 elements of an N component

alloy:

iCil

CVsfiCi

lCDiCDlf

iCVlfiC

t

(2.15)

In the present study, three components are tracked: Zn, Cu, and Mg. This equation tracks

the mixture composition, i i i

s s l lC f C f C , and the individual phase compositions are

found using the alloy phase equilibria described in the next section.

pfp VdT

dFLS

(2.13)

m

lm

lo

lpfc fFdT

dFffVLS 1

(2.14)

20

20

2.4 Solidification Model

Aluminum alloy 7050 was assumed to have a nominal composition based on the

mid-range of the specifications in ASTM B247 (Table 2.1).

Table 2.1: Chemical Specification Range of Aluminum Alloy 7050 from ASTM B247

Alloying Element Min. wt. fr. Max. wt. fr. Nominal wt. fr.

Zn 0.057 0.067 0.062

Cu 0.02 0.026 0.023

Mg 0.019 0.026 0.0225

Zr 0.0008 0.0015 0.0012

Fe - 0.0015 0.00075

Si - 0.0012 0.0006

Mn - 0.001 0.0005

Ti - 0.0006 0.0003

Cr - 0.0006 0.0002

In this study, only three alloying elements (Zn, Cu, and Mg) are modeled, chosen

based on their relative amount and tendency to segregate on the macroscopic scale during

casting. To link the energy conservation equation to the thermodynamic phase diagram

for multicomponent systems, the liquidus temperature as a function of the liquid

composition and the equilibrium partition coefficients for each element were determined

using ThermocalcTM

with the TCAL1 database. [33] The liquidus temperature function is

933.15 0.63 174.16 271.55 494.68Zn Cu Mg

liq l l lT C C C (2.16)

where the first term is the liquidus temperature of pure Al, and the second term is an

adjustment to due to the elements in ASTM B247 not tracked in this model (Zr, Fe, Si,

Cr, and Mn), assumed to have constant composition. The partition coefficients (ki = Cs

i /

Cli) for the alloying elements were determined to be 39.0Znk , 09.0Cuk , and 29.0Mgk .

21

21

The metal is assumed to be in local thermodynamic equilibrium. Depending on

the nominal composition, the metal can undergo either primary solidification, where most

macrosegregation occurs, or primary solidification followed by a secondary eutectic

reaction. The relevant thermophysical properties for the aluminum alloy 7050

solidification model are summarized in Table 2.2.

Table 2.2: Thermophysical Properties of Aluminum Alloy 7050

Thermophysical Property Value Reference

Liquid Density [kg/m3] 2515.0 [34]

Solid Density [kg/m3] 2744.1 [34]

Liquid Thermal Conductivity [W/m K] 83.2 [34]

Solid Thermal Conductivity [W/m K] 149.4 [34]

Specific Heat [J/kg K] 1141.0 [34]

Latent Heat [J/kg] 3.76x105 [34]

Liquid Thermal Expansion [1/K] 1.5x10-4

[35]

Solid Thermal Expansion [1/K] 2.29x10-5

[34]

Liquid Solutal Expansion [1/K]

(Zn, Cu, Mg) -0.65, -0.75, 0.53

[35]

Solid Solutal Expansion [1/K]

(Zn, Cu, Mg) -1.43, -2.01, 0.31

[36]

Liquid Viscosity [kg/m s] 0.0013 [36]

Average Solid Viscosity [kg/m s] 4.53μl [37]

2.5 Domain

The mixture model described in the previous sections is capable of predicting the

transport phenomena during the solidification of aluminum alloy 7050 in a grain refined

liquid melt. To simulate DC casting, proper domain and boundary conditions are applied.

In this study, 500 mm diameter billets are cast to a centerline solid height of 1.5 m. A

cylindrical nozzle submerged 20 mm into the liquid melt supplies hot liquid metal from a

feeding height of 35 mm. The grid is constructed in axisymmetric coordinates with a

22

22

uniform spacing of Δr = 5 mm and Δz = 10 mm. The radial grid spacing is chosen based

on the refinement needed for the thickness of the nozzle walls, 5 mm. To decrease the

computational expense of the simulations, the momentum equations are only solved in

the liquid and mixture regions. Because there is no velocity in the all solid region, the

ingot frame of reference is used, and therefore, the domain is fixed to the bottom block.

In order to accommodate liquid metal feeding through a submerged nozzle, the volume of

fluid (VOF) method is used. The VOF tracks the interface between two fluids, in this

case, liquid metal and air. A detailed description of the VOF method is provided by

Yanke et al. [38] A schematic of the numerical domain is shown in Figure 2.1.

Figure 2.1: Schematic of the numerical domain representing a hot top DC caster with a

submerged nozzle

23

23

To simulate a hot top mold, an isothermal boundary condition, equal to the

superheat temperature (950 K), was used at the top of the domain and the conductivity of

the air was set artificially high so that no heat was dissipated during diffusion to the

liquid melt. Two nozzle diameters, 20 and 40 mm, were simulated at the top-center of the

domain. The nozzle wall has the same thermophysical properties as aluminum alloy 7050,

however the viscosity of the nozzle wall control volumes are set arbitrarily high, 1015

,

extending from a no slip boundary condition. This method ensures proper alignment

between the coupled energy and momentum equations on a staggered grid. [31] Two

different metal distribution systems were modeled, a combo bag and a flow diffuser plate.

The dimensions of the combo bag are 50 x 75 mm, and the dimensions of the flow

diffuser plate are 50 x 50 mm in the axial and radial directions, respectively. Both

distribution systems where modeled by setting the permeability on the designated

momentum control volume faces to a specified values. For the combo bag, two different

permeabilities were investigated, one typical of a foam filter, 9.0E-8 m2, and another

typical of a Thermally Formed (TF) combo bag supplied by Pyrotek, 8.2 E-10 m2.

Schematics of both distribution systems are shown in Figure 2.2.

24

24

(a) (b)

Figure 2.2: Schematic of the numerical domain for the (a) combo bag and (b) flow

diffuser plate

The liquid metal is initially set to a height of 10 cm, with the bottom block in

contact with the bottom of the mold. The mold design consists of a hot top (adiabatic

region), a graphite ring, and an aluminum mold. Initially, the temperature of the liquid is

set to the pouring temperature, 950 K (677 oC), and the composition is the nominal alloy

composition. In order to allow time for the formation of a solid shell near the mold wall

before the bottom block is withdrawn, the liquid metal is held in the mold for 60 seconds.

After the specified hold time, the casting speed is set to 50% of the steady state casting

speed, Vcast, and is increased linearly until a diameter of metal has been added to

existing domain. In this study two casting speeds are investigated, Vcast = 30 and 60

mm/min.

25

25

In order to grow the domain in the axial direction, the top row of control volumes

were expanded to a grid spacing twice the original size and split at its center (Figure 2.3).

The transport values of the split control volumes were linearly interpolated. Expanding

the control volumes at the top of the domain was chosen to avoid expanding solidifying

control volumes. Because a control volume can only have a single fraction solid value,

expanding a solidifying control volume over predicts the local volume fraction solid. It is

also noted that if control volumes in the interior of the domain are expanded, an

advection term needs to be added to the momentum equations to account for the moving

mesh. The introduction of this source terms led to the smearing the composition field.

Figure 2.3: Schematic of the control volume expansion and splitting method used to grow

the domain: (a) start of expansion, (b) during expansion, (c) just before splitting, and (d)

after splitting.

26

26

2.6 Boundary Conditions

The velocity boundary conditions for the numerical model are shown

schematically in Figure 2.4. A symmetry boundary condition is assumed at the centerline,

and no slip boundary conditions are assumed at the outer radius and bottom of the

domain. At the top of the domain, a nonuniform boundary condition consisting of a

velocity inlet inside of the nozzle, a no slip condition above the nozzle wall, and a

velocity outlet over the air is used. The inlet velocity accounts for the metal addition

necessary to feed the cast speed and shrinkage in the correct reference frame (2.17).

cast billetin shrinkage cast

in

V AV V V

A

(2.17)

The outlet velocity is determined using the volume conservation method described by

Versteeg (2.18). [39]

in in

out

out

V AV

A

(2.18)

Figure 2.4: Schematic of the velocity boundary conditions used in the numerical model.

27

27

The thermal boundary conditions for the mold are shown in Figure 2.5. An

isothermal boundary condition equal to the pouring temperature, 950 K, is applied to the

top of the domain with an adiabatic region at the top, outer surface to simulate the hot top

region of the mold. A heat transfer coefficient boundary condition is applied at the

bottom of the domain to simulate the heat transfer between the actively cooled bottom

block and billet. The heat transfer coefficient for the bottom block is difficult to

determine, but its effects become negligible as the solidification front moves away from

the bottom block. The heat transfer coefficients in the mold are treated as a function of

fraction solid, which accounts for the decrease in contact due to shrinkage. Since the

mold is water-chilled, it is assumed to be isothermal.

Figure 2.5: Schematic of the thermal boundary conditions in the hot top mold

Below the mold, the billet surface is quenched by the impingement of water jets

and cooled further by the water falling along the billet surface. The boundary conditions

for the water jet are taken from Weckman and Niessan [3], where the heat transfer

28

28

coefficient is a function of the volumetric water flow rate (Q), billet diameter (D), billet

surface temperature (Twall), and water temperature (TH20). In order to determine if the

water was boiling on the surface of the billet, the incipient boiling criterion in equation

(2.20) is used.

16.2" 39100 boilwallIB TTq . (2.19)

If incipient boiling criterion is greater than the heat flux predicted for the no-boiling

condition in equation (2.20), which calculated using the heat transfer coefficient in

equation (2.21), then the heat transfer coefficient is replaced with the heat transfer

coefficient for boiling shown in equation (2.22).

02" Hwall TThq , (2.20)

31

51067.12

704 2

D

Qx

TTh

OHwall

(2.21)

OHwall

boilwallOHwall

bTT

TT

D

Qx

TTh

2

2

33

1

5 8.201067.1

2704

(2.22)

29

29

2.7 Summary of Numerical Model

In this chapter, a numerical model to simulate the DC casting process for aluminum

alloy 7050 was presented. The presented model includes thermosolutal buoyancy effects,

shrinkage-driven flow in the rigid mushy zone, and the effect of free-floating solid during

multicomponent solidification. This model was modified to simulate forced convection

through a submerged nozzle. To investigate the transient effects of DC casting, a cell

expansion and splitting method to grow the domain in the axial direction was developed.

This model was used to perform a parametric study on the process parameters associated

with liquid metal distribution during DC casting.

30

30

CHAPTER 3. RESULTS AND DISCUSSION

Using the numerical model described in Chapter 2, a parametric study was

performed to investigate the relative effects of casting speed, liquid metal distribution

system type, and nozzle width on the flow field and macrosegregation during DC casting.

The base case for the parametric study consists of a 500 mm diameter billet cast at 30

mm/min through a 40 mm diameter submerged nozzle. A summary of the test cases used

for the parametric study are provided in Table 3.1.

Table 3.1: Summary of the Parametric Study Cases

Case # Casting Speed

[mm/min]

Nozzle Diameter

[mm]

Distribution System

1* 30 40 None

2 60 40 None

2 30 40 High Permeability Combo Bag

3 30 40 Low Permeability Combo Bag

4 60 40 Low Permeability Combo Bag

5 30 40 Diffuser Plate

6 60 40 Diffuser Plate

7 30 20 None

8 30 20 Low Permeability Combo Bag

9 30 20 Diffuser Plate

* Base Case

31

31

To analyze the predicted flow fields, the stream function, ψ, is used. The

stream function is everywhere tangent to the velocity field; uz

and vr

. The

mass flow rate between any two streamlines is the difference in stream function

values, Δ(ρψ). If Δ(ρψ) is constant, a decrease in streamline spacing indicates a local

increase in velocity.

To analyze the predicted macrosegregation for the different test cases, radial

segregation profiles and composition distribution are used. The radial segregation

profile shows the composition deviation, 0

0

iC CC

C

, at z = 1.5 m. The composition

distribution shows the volume weighted frequency of a composition value in the

steady state region of casting, defined as the axial position when the centerline

composition variations are less than 1x10-5

.

32

32

3.1 Base Case

The base case for the DC casting parametric study consists of a 500 mm

diameter billet of alloy 7050 cast at 30 mm/min. The hot liquid metal enters the mold

through a 40 mm diameter nozzle submerged in the liquid melt to a depth of 20 mm.

The numerical results for the temperature, composition, flow field, and fraction solid

are shown in Figure 3.1. The steady state sump depth, defined as the distance from

the top of the liquid melt to the top of the rigid mushy is 24 cm. The streamlines

indicate that as the liquid metal enters the mold through the submerged nozzle, it is

entrained toward to outer surface. At the outer surface the high rate of heat extraction

from the impinging water jets creates a momentum source due to negative thermal

buoyancy effects. As the liquid drops below the liquidus temperature, a slurry of

solute-depleted grains is formed adding a solutal component to the thermal buoyancy.

The cooled slurry falls along the edge of the rigid mushy zone towards the centerline.

At the centerline, the slurry collides with a symmetric flow field from the opposing

side of the billet and is driven upwards. The net effect of the thermosolutal buoyancy

is a clockwise recirculating flow cell in the slurry region with a maximum flowrate of

0.355 kg/s. The temperature field in the slurry is stratified, and the composition field

is relatively uniform due to convective mixing.

33

33

Figure 3.1: For Case 1 (base case), the temperature and flow fields are shown on the

left and the mixture composition, the solidus and liquidus lines (solid), and the

packing interface (dashed) are shown on the right. The solid streamlines show the

clockwise flow with 0.1< ρΨ < 1 and ΔρΨ=0.1 kg/s and dotted streamline show the

counter-clockwise flow with -0.005< ρΨ < -0.000025 and ΔρΨ=0.0004975 kg/s.

34

34

As the slurry flows along the rigid mushy zone, it becomes entrained by

shrinkage-driven flow which is oriented perpendicular to the solidification front. The

shrinkage-driven flow in the rigid mushy zone can be visualized in either the ingot

frame of reference (Figure 3.2a) or the mold frame of reference (Figure 3.2b).

Because the ingot frame of reference allows for the visualization of the horizontal

component of shrinkage-driven flow, which contributes to macrosegregation by

pulling solute rich liquid towards the outer surface, it is used in this study.

As seen in Figure 3.2b, the free-floating solute-depleted grains formed the

near surface travel a short distance before becoming entrained by the shrinkage-

driven flow. This results in the negative subsurface segregation shown in Figure 3.3.

The free-floating grains that form away from the outer surface are free to move with

the thermosolutal buoyancy cell. For the base case, most of these free-floating grains

settle along the rigid mushy zone before reaching the bottom of the sump. Along the

rigid mushy zone, the horizontal component of shrinkage-shriven flow pulls solute-

rich liquid towards the outer surface. The negative segregation increases in magnitude

towards the centerline as there is less volume of solute-rich liquid to fill the depletion

caused by the shrinkage-driven flow. No grains are predicted to settle at the bottom of

the sump, indicated by the positive axial free-floating solid velocity at the bottom of

the sump (Figure 3.4). Because the solidification front is horizontal at the bottom of

the sump, the shrinkage-driven flow does not have a significant horizontal component

to pull enriched liquid away from the centerline, and therefore, there is no mechanism

for negative centerline segregation. The transport of solute-rich liquid toward the

35

35

centerline by thermosolutal buoyancy is the dominate mechanism for segregation in

the base case, resulting in positive segregation that extends ~ 5 cm from the centerline.

36

36

(a)

(b)

Figure 3.2: Close-up view of the rigid mushy zone for both (a) the ingot frame of

reference with -0.000025< ρΨ <-0.005 and ΔρΨ=0.0004975 kg/s and (b) the mold

frame of reference with -0.1< ρΨ < -0.001 and ΔρΨ=0.001 kg/s. The temperature and

flow fields are shown on the left and the mixture composition, the solidus and

liquidus lines (solid), and the packing interface (dashed) are shown on the right.

37

37

Figure 3.3: Axial component of the solid velocity along the bottom of the sump for

the base case.

-0.004

-0.002

0

0.002

0.004

0.006

0.008

0.01

0.012

0.014

0 0.02 0.04 0.06 0.08 0.1

Axia

l S

oli

d V

eloci

ty [m

/s]

R (m)

38

38

The segregation profiles shown in Figure 3.3 include each of the modeled

components (Zn, Cu, and Mg). The level of centerline macrosegregation is known to

be a function of the partition coefficient, not necessarily the relative weight percent of

the nominal composition. [27] As seen in segregation profile, Mg exhibits the least

amount of segregation and Cu exhibits the greatest amount of segregation. Because

each of the alloying components will follow similar segregation trends, only the Zn

segregation profiles will be compared for simplicity in the interpretation of the results.

The composition distributions for the base case are shown in Figure 3.5. Each

of the distributions have a peak near the initial composition. Because of the presence

of both positive and negative segregation in the base case, the distribution tails off in

both directions from the peak. The specification range for ASTM 247 is indicated by

the dotted lines. Each of the components lie well within the specification range,

occupying 15%, 26%, and 11% of the specification range for Zn, Cu, and Mg

respectively. Therefore, the macrosegregation for the base case is not a significant.

39

39

Figure 3.4: Predicted composition deviation of each component for the 30mm/min

base case 1.5 m from the bottom block

40

40

(a)

(b)

(c)

Figure 3.5: Steady state volume weighted composition distributions for Case 1 (base

case) of (a) Zn, (b) Cu, and (c) Mg .

0

0.05

0.1

0.15

0.2

0.25

0.3

0.055 0.057 0.059 0.061 0.063 0.065 0.067 0.069

Vo

lum

e F

ract

ion

Zn (wt.fr.)

00.05

0.10.15

0.20.25

0.30.35

0.40.45

0.5

0.018 0.02 0.022 0.024 0.026 0.028

Vo

lum

e F

ract

ion

Cu (wt.fr.)

0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.018 0.02 0.022 0.024 0.026

Vo

lum

e F

ract

ion

Mg (wt.fr.)

41

41

3.2 Effect of Casting Speed

High solute alloys such as 7050 are typically cast in the range 30-60 mm/min

to prevent hot cracking. [5] To investigate the effect of casting speed for aluminum

alloy 7050 billets, the upper and lower limits of the casting speed range are compared.

The numerical results for the temperature, composition, flow field, and fraction solid

are shown in Figure 3.6, comparing Vcast = 30 and 60 mm/min. As the casting speed

increases, there is less time for heat to be extracted along the outer surface, resulting

in more vertical isotherms. As a result, the steady state sump depth increases from 24

cm to 59 cm. The increase in the temperature gradients and sump depth increases the

thermosolutal buoyancy effects in the slurry region. The maximum flow rate of the

thermosolutal buoyancy increases from 0.356 kg/s to 1.142 kg/s.

In addition to the increase in convection in the slurry from thermal buoyancy

effects, the magnitude of the forced convection from the nozzle also increases. The

primary effect of the forced flow is that the super heat of the incoming liquid metal is

driven into the solidification region, pushing the liquidus line downwards. Due to

increased competition with the upward thermal buoyancy at the centerline, the flow

field becomes destabilized and long range oscillations develop. The flow field is now

unsteady, oscillating between penetration of the superheated liquid metal into the

slurry region (Figure 3.7a) and entrainment of the forced convection by the

thermosolutal convection along the top of the liquid region (3.7b). The oscillating

flow field causes composition banding on the outer surface of the billet. Composition

variations in the axial direction are known to occur for unstable flow fields. [24,40]

42

42

The maximum composition variation in the axial direction is ~3% for Vcast = 60

mm/min.

(a) (b)

Figure 3.6: Effect of casting speed through a submerged nozzle for (a) 30 mm/min

and (b) 60 mm/min. The temperature and flow fields are shown on the left and the

mixture composition, the solidus and liquidus lines (solid), and the packing interface

(dashed) are shown on the right. The solid streamlines show the clockwise flow with

0.1< ρΨ < 1 and ΔρΨ=0.1 kg/s and dotted streamline show the clockwise shrinkage

driven flow with -0.005< ρΨ < -0.000025 and ΔρΨ=0.0004975 kg/s.

43

43

(a) (b)

Figure 3.7: Close up of the rigid mushy zone for a casting speed of 60 mm/min,

showing the two phases of flow field oscillation: (a) entrainment by the thermosolutal

bouyancy and (b) penetration into the slurry region. The temperature and flow fields

are shown on the left and the mixture composition, the solidus and liquidus lines

(solid), and the packing interface (dashed) are shown on the right.

44

44

The transition from positive to negative centerline segregation is possible

depending on relative transport of solute rich liquid toward the center by

thermosolutal buoyancy and away from the center by shrinkage-driven flow. The

amount of solute rich liquid transported by the shrinkage-driven flow depends on the

thickness of the rigid mushy zone, and the direction of the transported liquid depends

on the orientation of the solidification front. The shape of the rigid mushy zones are

compared in Figure 3.8 for Vcast = 30 and 60 mm/min. The top line represents the

solidification front (gs,p = 0.3) and the bottom line represents the solidus (gs = 0.0). As

casting speed increases the thickness of the rigid mushy zone increases from 8 cm 11

cm, and the solidification front becomes more vertical. Therefore, as casting speed

increases a larger volume of solute rich liquid is transported away from the centerline

by shrinkage driven flow. The net result is transition from positive to negative

centerline segregation as shown in Figure 3.9.

45

45

Figure 3.8: Comparison of the rigid mushy zone profiles for casting speeds of 30 and

60 mm/min.

The increase in both thermosolutal buoyancy and shrinkage-driven flow result

in an overall increase in macrosegregation. The composition distributions are

compared in Figure 3.10 for the two casting speeds. At Vcast = 60 mm/min, there is

no longer positive segregation at the centerline so there is only one tail to the

distribution. The volume of the specification range occupied increases to from 15%,

26%, and 11% at Vcast = 30 mm/min to 37%, 42%, and 24% at Vcast = 60 mm/min

for Zn, Cu, and Mg, respectively. The increase in negative centerline segregation due

to increasing casting speed is not large enough to predict the composition field to be

outside of the specified range. Therefore, in the casting range typically used for alloy

7050 macrosegregation is not predicted to be a limiting factor.

46

46

Figure 3.9: Normalized radial Zn segregation profile at 1.5m from the bottom block,

comparing the effects of casting speed.

47

47

(a)

(b)

(c)

Figure 3.10: Steady state volume weighted composition distributions for Cases 1 and

2 of (a) Zn, (b) Cu, and (c) Mg .

0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.055 0.057 0.059 0.061 0.063 0.065 0.067 0.069

Vo

lum

e F

ract

ion

Zn (wt.fr.)

30 mm/min

60 mm/min

0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

0.5

0.018 0.02 0.022 0.024 0.026 0.028

Vo

lum

e F

ract

ion

Cu (wt.fr.)

30 mm/min

60 mm/min

0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.018 0.02 0.022 0.024 0.026

Vo

lum

e F

ract

ion

Mg (wt.fr.)

30 mm/min

60 mm/min

48

48

3.3 Effect of Distribution System

The use of a submerged nozzle increases the velocity of the incoming hot

liquid metal, the strength of which may be strong enough to destabilize the flow field,

causing long range oscillations and composition banding. In order to reduce the

flowing momentum into the slurry region, diffuser plates and combo bags are often

used. The purpose of a combo bag is to preferentially distribute the flow at the top of

the liquid metal pool and to trap incoming inclusions. [18] The purpose of a diffuser

plate is to redirect the nozzle flow towards the outer surface. In both cases, the main

goal is to aid in the entrainment of the incoming flow by the thermosolutal buoyancy

momentum source on the outer surface. To investigate the effects of these liquid

metal distribution systems on the flow field and macrosegregation of alloy 7050

billets, liquid metal feeding through a 40 mm diameter submerged nozzle is studied

for Vcast = 30 and 60 mm/min.

The distribution of flow from a combo bag is a function of both its

permeability and aspect ratio. In this study two different permeabilities are

investigated; one typical of a thin foam filter, 9.0x10-8

m2, and another provided by

Pyrotek for their industrially available Thermally-Formed (TF) combo bag, 8.2x10-10

m2. The permeability of the foam filter was too high to have a significant effect of the

flow field at Vcast = 30 mm/min. With the addition of the thin foam filter, the

maximum flow rate of the thermosolutal buoyancy increased from 0.355 to 0.365 kg/s,

a relative difference of only 2.8%. The slight increase in the thermosolutal buoyancy

flowrate was not large enough to effect macrosegregation prediction. Although the

49

49

numerical results do not predict a change in the flow field or macrosegregation for a

thin foam filter, it is noted that it can still be used to trap incoming inclusions.

The permeability of the TF combo bag was low enough to have an effect on

the flow field. The numerical result of the temperature, composition, flow field, and

fraction solid are shown in Figure 3.11 for Vcast = 30 and 60 mm/min. For Vcast =

30 mm/min, the combo bag diverts the thermosolutal convection around the bottom

of the bag. By removing the competition of the downward flowing forced convection

and the upward flowing thermosolutal buoyancy, the nozzle momentum flows to the

bottom of the bag where is diverted towards the outer surface. This does not have a

significant effect on entrainment at Vcast = 30 mm/min, as the nozzle flow is

entrained regardless of the presence of the combo bag. Because the flow cell in the

slurry is constricted to a smaller area, the maximum flowrate in the slurry increases

from 0.355 to 0.670 kg/s. As the casting speed is increased to 60 mm/min, the

redirection of the nozzle flow towards the outer surface removes the oscillations in

the flow field, and therefore, the composition banding observed when a combo bag

was not used. Similar to Vcast = 30 mm/min, the constriction of the flow cell in the

slurry results in an increase in the maximum flow rate from 1.142 to 1.383 kg/s The

increase in forced convection in the slurry from the combo bag deepens the sump ~1

cm, and increases the height of the liquidus surface at the outer radius by forcing

cooled liquid transported by the recirculating flow cell into this region.

The redirection of the flow by a diffuser plate was modeled using a

permeability low enough to block all flow in the axial direction, 1.0x10-15

m2. The

numerical results for the temperature, composition, flow field, and fraction solid are

50

50

shown in Figure 3.12 for Vcast = 30 and 60 mm/min. Similar the low permeability

combo bag, the redirection of the nozzle flow towards the outer surface removes the

vertical component to the interaction of the nozzle flow and thermal buoyancy. The

result is the entrainment of the nozzle flow towards the outer surface, and the removal

of oscillations that develop in the flow field at higher casting speeds. The main

difference between the low permeability combo bag and the flow diffuser plate is

their aspect ratio. While the combo bag extends 7.5 cm from the centerline, the

diffuser plate only extends 5 cm. Due to the flow cell in the slurry not being

constricted to as small of an area, the maximum flowrates for the diffuser plate are

less than for the combo bag. The maximum flowrates for Vcast = 30 and 60 mm/min

are 0.573 and 1.246 compared to 0.670 and 1.383 kg/s for the combo bag. As a result,

the increase in sump depth and the suspension of the liquidus surface at the outer

surface are not observed for the diffuser plate.

51

51

(a) (b)

Figure 3.11: Effect of a combo bag for Vcast = (a) 30 and (b) 60 mm/min. The

temperature and flow fields are shown on the left and the mixture composition, the

solidus and liquidus lines (solid), and the packing interface (dashed) are shown on the

right.

52

52

(a) (b)

Figure 3.12: Effect of diffuser plate for Vcast = (a) 30 and (b) 60 mm/min. The

temperature and flow fields are shown on the left and the mixture composition, the

solidus and liquidus lines (solid), and the packing interface (dashed) are shown on the

right.

53

53

The composition distributions for the different liquid metal distribution

systems are shown in Figure 3.16 and 3.17 for Vcast = 30, and 60 mm/min ,

respectively. In order to observe the relatively small change in composition that the

liquid metal distribution system cause for a 40 mm diameter nozzle, the scale of the

composition rage is reduced, so the specification limits are not shown. It is seen that

there is no significant change to the composition field at Vcast = 30 mm/min for the

different liquid metal distribution systems. This is an expected result as the nozzle

flow is entrained for the base. At Vcast = 60 mm/min, the only significant effect is

that the combo bag increases the positive segregation at the outer surface of the billet.

This increase the volume of the specification range occupied from 37%, 42% and 24%

for the nozzle only configuration to 42%, 49%, and 27% for the low permeability

combo bag.

54

54

(a)

(b)

(c)

Figure 3.13: Steady state volume weighted composition distributions for Cases 1,3,

and 5 of (a) Zn, (b) Cu, and (c) Mg .

0

0.1

0.2

0.3

0.4

0.0605 0.061 0.0615 0.062 0.0625 0.063 0.0635

Vo

lum

e F

ract

ion

Zn (wt.fr.)

Nozzle Only

Combo Bag

Diffuser Plate

0

0.1

0.2

0.3

0.4

0.5

0.6

0.0215 0.022 0.0225 0.023 0.0235 0.024 0.0245

Vo

lum

e F

ract

ion

Cu (wt.fr.)

Nozzle Only

Combo Bag

Diffuser Plate

0

0.1

0.2

0.3

0.4

0.0215 0.022 0.0225 0.023 0.0235

Vo

lum

e F

ract

ion

Mg (wt.fr.)

Nozzle Only

Combo Bag

Diffuser Plate

55

55

(a)

(b)

(c)

Figure 3.14: Steady state volume weighted composition distributions for Cases 2,4,

and 6 of (a) Zn, (b) Cu, and (c) Mg .

0

0.1

0.2

0.3

0.4

0.5

0.058 0.059 0.06 0.061 0.062 0.063

Vo

lum

e F

ract

ion

Zn (wt.fr.)

Nozzle Only

Combo Bag

Diffuser Plate

0

0.1

0.2

0.3

0.4

0.5

0.02 0.021 0.022 0.023 0.024

Vo

lum

e F

ract

ion

Cu (wt.fr.)

Nozzle Only

Combo Bag

Diffuser Plate

0

0.1

0.2

0.3

0.4

0.5

0.6

0.02 0.021 0.022 0.023

Vo

lum

e F

ract

ion

Mg (wt.fr.)

Nozzle Only

Combo Bag

Diffuser Plate

56

56

3.4 Effect of Nozzle Diameter

It is reported in previous work that the forced convection from a vertical

nozzle leads to the penetration of the rigid mushy zone by the superheated liquid

metal. [24,40] The magnitude of the forced convection from the 40 mm diameter

nozzle was not large enough to cause this effect in the investigated range of casting

speed. To investigate the effects of the increased momentum flowing down the

centerline, a 20 mm diameter nozzle is used. The nozzle diameter was chosen based

on the ratio of the nozzle exit velocity to cast speed as a function of the nozzle to

billet area ratio, shown in Figure 3.18. From this plot, it is seen that the nozzle exit

velocity is highly sensitive to the nozzle to billet area ratio below values of 0.1.

Figure 3.15: Plot of the normalized velocity as a function of the nozzle to billet area

ratio

0

100

200

300

400

500

600

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Vn

ozzle /

Vca

st

Anozzle / Aingot

20 mm Diameter Nozzle

40 mm Diameter Nozzle

57

57

The nozzle to billet area ratio decreases from 6.4% for the 40 mm diameter

nozzle to 1.6% for the 20 mm diameter nozzle. This change results in a 400%

increase in the nozzle exit velocity for the 20 mm diameter nozzle case. Due to this

significant increase in nozzle exit velocity, the nozzle flow jets down the centerline of

the billet to the rigid mushy zone, resulting the formation of a cliff-shaped sump,

shown in Figure 3.19a. The solidifcation front of the cliff-shaped sump is completely

vertical, so there is a large horizontal component to the shrinkage-flow, resulting in a

solute depleted region ~2 cm from the centerline, which varies along the axial

direction.

Currently, the nature of the solid packing model used to define the

demarcation between the slurry and rigid mushy region cannot resolve the interaction

of the nozzle jet and thermosolutal buoyancy at the top of the sump cliff. When the

volumetric fraction solid in a control volume reaches the critical packing fraction, the

control volume becomes packed and the sudden drop in permeability restricts the

fluid flow in the control volume. When this occurs in a control volume near the

centerline, the superheated fluid from the inlet nozzle can be cut off at the top of the

sump cliff. Because there is no condition for unpacking, the control volumes at the

top of the sump cliff pack permenantely. Without superheated liquid restricting the

solidification rate in the rigid mushy zone, the solidus reaches the top of the rigid

mushy zone and the build up of the sump cliff begins again. This sequence repeats

itself throughtout the simulation, resulting in the banded composition observed. While

the predicted composition field resembles reported results [24] its nature here is a

numerical artifact, and not a true model of the physical phenomenon.

58

58

In order to prevent the jetting of the forced convective flow down the

centerline, the low permeability combo bag and diffuser plate were added to the

domain. As seen in Figure 3.19, both the low permeability combo a bag and diffuser

plate prevent the downward jetting of the nozzle flow. The redirection of the nozzle

flow aids in the entrainment by the thermosolutal buoyancy momentum source on the

outer surface, and the resulting temperature, composition, flow field and fraction solid

are similar to the base case. As seen in Figure 3.21,the composition distributions are

similar to the base case indicating that the magnitude of forced convection does not

have a significant effect on the macrosegregation of as long as it is prevented from

flowing down the centerline.

59

59

(a)

(b) (c)

Figure 3.16: Effect of nozzle diameter for Vcast = 30 mm/min and the distribution

systems: (a) nozzle only, (b) combo bag, and (c) diffuser plate. The temperature and

flow fields are shown on the left and the mixture composition, the solidus and

liquidus lines (solid), and the packing interface (dashed) are shown on the right.

60

60

(a)

(b)

(c)

Figure 3.17: Steady state volume weighted composition distributions for Cases 1,8,

and 9 of (a) Zn, (b) Cu, and (c) Mg .

0

0.1

0.2

0.3

0.4

0.5

0.06 0.061 0.062 0.063 0.064

Vo

lum

e F

ract

ion

Zn (wt.fr.)

NozzleOnlyComboBag

0

0.1

0.2

0.3

0.4

0.5

0.022 0.023 0.024

Vo

lum

e F

ract

ion

Cu (wt.fr.)

NozzleOnlyComboBag

0

0.1

0.2

0.3

0.4

0.5

0.0215 0.0225 0.0235

Vo

lum

e F

ract

ion

Mg (wt.fr.)

NozzleOnlyComboBag

61

61

3.5 Results Summary

A 500mm diameter billet is used to examine the effect that different liquid

metal distribution systems have on the flow field and macrosegregation during DC

casting. The results show that the sump depth and shape depended on the thermal

conditions in the billet, which is primarily a function of the casting speed. Increasing

the casting speed allows less time for heat extraction, resulting in the deepening of the

sump and a more vertical solidification profile. With more volume in the slurry zone,

the magnitude of the thermosolutal convection increases, resulting in more transport

of solute-rich liquid toward the centerline. The increased vertical component of the

solidification front increases the horizontal component of the shrinkage-driven flow

in the rigid mushy zone pulling solute-rich liquid away from the centerline. The

results show that for the exception of the downward jetting of forced convection at

the centerline, the metal distribution system design had little effect on the

macrosegregation in the billet

In past studies, it was reported that the use of a vertical submerged nozzle led

to a deeper sump and increased segregation across the ingot compared to a level pour

method or horizontal feeding. [24,40] These studies fail to show the effect that the

nozzle to billet area ratio has on the velocity at the nozzle exit. It was shown in Figure

3.16, that the jetting phenomenon is only observed for nozzle to billet area ratios

below 0.05. It was also shown, that the addition of combo bags and diffuser plates can

eliminate the downward jetting effect by redirecting the flow horizontally, resulting in

similar sump shapes and segregation profiles as nozzle diameters that allow for the

entrainment of the incoming liquid metal.

62

62

CHAPTER 4. SUMMARY AND FUTURE WORK

4.1 Summary

A fully transient numerical model for DC casting is used to investigate the

effects of liquid metal distribution on the flow field and macrosegregation of

aluminum alloy 7050. A method for growing the domain is developed so that the

transient effects of the process can be investigated. Numerical methods for including

a nozzle wall, a permeable combo bag, and an impermeable diffuser plate are

presented.

A 500 mm diameter billet cast at 30 mm/min through a 40 mm diameter

nozzle is used as the base case. As the liquid metal enters the domain it becomes

entrained by negative thermal buoyancy effects at the outer surface. At Vcast = 60

mm/min, the magnitude of the nozzle exit velocity increases, resulting in increased

competition with the thermal buoyancy. This competition produces oscillations in the

flow field which causes composition banding along the outer surface of the billet. The

redirection of flow by either a combo bag or diffuser plate aids in the entrainment of

nozzle flow, resulting in the elimination of flow field oscillations and composition

banding.

63

63

The nozzle diameter was shown to have a significant effect on the flow field, at

nozzle to billet area ratios < 0.05. If the nozzle diameter is low enough, jetting of the

nozzle flow down the centerline. Jetting of the nozzle flow produces a cliff shaped

sump at the centerline, which increases the extent of macrosegregation in the billet.

The redirection of flow by either a combo bag or diffuser plate prevents nozzle jetting

down the centerline.

4.2 Future Work

The main area of interest for liquid metal distribution in DC casting involves

the effects that distribution systems have on high aspect ratio slabs. In order to model

this geometry, a 3-D model is needed. Because the casting speed is higher for slabs

and the sump is deeper, turbulent effects are known to exist. To investigate liquid

metal distribution in high aspect ratio slabs, a turbulence model would need to be

included in the 3-D model.

64

LIST OF REFERENCES

64

64

LIST OF REFERENCES

[1] J.F. Grandfield, D.G. Eskin, I.F. Bainbridge, Direct-Chill Casting of Light Alloys,

1st ed., John Wiley & Sons, Inc., Hoboken, New Jersey, 2013.

[2] D.G. Eskin, Physical Metallurgy of Direct Chill Casting of Aluminum Alloys, 1st

ed., CRC Press, Boca Raton, 2008.

[3] D.C. Weckman, P. Niessan, A Numerical Simulation of the D. C. Continuous

Casting Process Including Nucleate Boiling Heat Transfer, Metall. Mater. Trans. B.

13 (1982) 593–602.

[4] R. Nadella, D.G. Eskin, Q. Du, L. Katgerman, Macrosegregation in direct-chill

casting of aluminium alloys, Prog. Mater. Sci. 53 (2008) 421–480.

doi:10.1016/j.pmatsci.2007.10.001.

[5] R. Nadella, D. Eskin, L. Katgerman, Effect of grain refining on defect formation in

DC cast Al-Zn-Mg-Cu alloy billet, TMS Light Met. (2007) 727–732.

[6] C.J. Vreeman, F.P. Incropera, The effect of free-floating dendrites and convection

on macrosegregation in direct chill cast aluminum alloys Part II : predictions for Al

± Cu and Al ± Mg alloys, Int. J. Heat Mass Transf. 43 (2000) 687–704.

[7] C.J. Vreeman, J.D. Schloz, M.J.M. Krane, Direct Chill Casting of Aluminum

Alloys: Modeling and Experiments on Industrial Scale Ingots, J. Heat Transfer.

124 (2002) 947. doi:10.1115/1.1482089.

[8] D.G.G. Eskin, J.J. Zuidema, V.I.I. Savran, L. Katgerman, Structure formation and

macrosegregation under different process conditions during DC casting, Mater. Sci.

Eng. A. 384 (2004) 232–244. doi:10.1016/j.msea.2004.05.066.

65

65

[9] A.V. V Reddy, C. Beckermann, Modeling of Macrosegregation Due to

Thermosolutal Convection and Contraction-Driven Flow in Direct Chill

Continuous Casting of an Al-Cu Round Ingot, Metall. Mater. Tranactions B. 28

(1997) 479–489. doi:10.1007/s11663-997-0115-2.

[10] D.G. Eskin, L. Katgerman, Macrosegregation Mechanisms in Direct-Chill Casting

of Aluminium Alloys, Mater. Sci. Forum. 630 (2009) 193–199.

doi:10.4028/www.scientific.net/MSF.630.193.

[11] H. Yu, D.A. Granger, Macrosegregation in Aluminum Alloy Ingot Cast by the

Semicontinuous Direct Chill (DC) Method, in: Alum. Alloy. Their Phys. Mech.

Prop., EMAS, 1986: pp. 17–29.

[12] M.G. Chu, J.E. Jacoby, Essential Readings in Light Metals: Cast Shop for

Aluminum Production., (2013).

[13] D.G. Eskin, Q. Du, L. Katgerman, Relationship between shrinkage-induced

macrosegregation and the sump profile upon direct-chill casting, Scr. Mater. 55

(2006) 715–718. doi:10.1016/j.scriptamat.2006.06.021.

[14] Q. Du, D.G. Eskin, L. Katgerman, Modeling Macrosegregation during Direct-Chill

Casting of Multicomponent Aluminum Alloys, Metall. Mater. Trans. A. 38 (2007)

180–189. doi:10.1007/s11661-006-9042-0.

[15] C.J. Vreeman, F.P. Incropera, Numerical Discretization of Species Equation

Source Terms in Binary Mixture Models of Solidification and Their Impact on

Macrosegregation in Semicontinuous, Direct Chill Casting Systems, Numer. Heat

Transf. Part B Fundam. 36 (1999) 1–14. doi:10.1080/104077999275749.

[16] C.J. Vreeman, M.J.M. Krane, F.P. Incropera, The effect of free-floating dendrites

and convection on macrosegregation in direct chill cast aluminum alloys Part I :

model development, Int. J. Heat Mass Transf. 43 (2000) 677–686.

[17] C.J.C.J. Vreeman, J.D.D. Schloz, M.J.M.M.J.M. Krane, Direct Chill Casting of

Aluminum Alloys: Modeling and Experiments on Industrial Scale Ingots, J. Heat

Transfer. 124 (2002) 947. doi:10.1115/1.1482089.

[18] L. Begum, 3-D TRANSPORT PHENOMENA IN VERTICAL DIRECT CHILL

CASTING PROCESSES, (2013).

66

66

[19] M. Hasan, L. Begum, IMECE2014-36763, (2015) 1–12.

[20] D. Larouche, TF Combo bag - Mathematical simulation results and up-date Dept .

of Mining , Metallurgy and Materials Engineering Adrien-Pouliot Bldg, (2005)

14–16.

[21] S. Tremblay, A. Arsenault, DC Casting Using a TF Combo Bag: 3D Modeling of

the Distribution of Molten Metal and Heat Transfer, Mater. Sci. …. 630 (2010)

223–232. doi:10.4028/www.scientific.net/MSF.630.223.

[22] F. Ilinca, J.F. Hetu, A. Arsenault, D. Larouche, S.P. Tremblay, 3D Modeling of the

Glow and Heat Transfer During DC Casting with a Combo Bag, NRC

Publucations Arch. (n.d.).

[23] S.C. Flood, L. Katgerman, a. H. Langille, S. Rogers, C.M. Read, Modellling of

Fluid Flow and Stress Phenomena During Dc Casting of Aluminium Alloys,

Essent. Readings Light Met. Cast Shop Alum. Prod. (1989) 862–866.

[24] L. Zhang, D.G. Eskin, A. Miroux, T. Subroto, L. Katgerman, Effect of Inlet

Geometry on Macrosegregation During the Direct Chill Casting of 7050 Alloy

Billets: Experiments and Computer Modelling, IOP Conf. Ser. Mater. Sci. Eng. 33

(2012) 1–8. doi:10.1088/1757-899X/33/1/012019.

[25] G.U. Gruen, I. Eick, D. Vogelsang, 3-D modeling of fluid flow and heat transfer

for DC casting of rolling ingots, Light Met. 1997. (1997) 863–869.

[26] J. Ni, C. Beckermann, A volume-averaged two-phase model for transport

phenomena during solidification, Metall. Trans. B. 22 (1991) 349–361.

doi:10.1007/BF02651234.

[27] B. Gariepy, Y. Caron, Investigation in the effects of casting parameters on the

extent centerline macrosegregation in DC cast sheet ingots, Light Met. 1991. (1991)

961–971.

[28] ALCOA Product Catalog: Alloy7050 Plate and Sheet, (2015).

https://www.alcoa.com/mill_products/catalog/pdf/alloy7050techsheetrev.pdf

(accessed October 6, 2015).

67

67

[29] W.D. Bennon, F.P. Incropera, A continuum model for momentum, heat and

species transport in binary solid-liquid phase change systems—II. Application to

solidification in a rectangular cavity, Int. J. Heat Mass Transf. 30 (1987) 2171–

2187. doi:10.1016/0017-9310(87)90095-0.

[30] W.D. Bennon, F.P. Incropera, A continuum model for momentum, heat and

species transport in binary solid-liquid phase change systems—I. Model

formulation, Int. J. Heat Mass Transf. 30 (1987) 2161–2170. doi:10.1016/0017-

9310(87)90094-9.

[31] S.V. Patankar, Numerical Heat Transfer and Fluid Flow, Hemisphere Publishing,

Washington DC, 1980.

[32] V.R. Voller, C.R. Swaminathan, GENERAL SOURCE-BASED METHOD FOR

SOLIDIFICATION PHASE CHANGE, Numer. Heat Transf. Part B Fundam. 19

(1991) 175–189.

[33] K. Fezi, J. Coleman, M.J.M. Krane, Macrosegregation During Direct Chill Casting

of Aluminum Alloy 7050, in: M. Hyland (Ed.), Light Met., Wiley, 2015: pp. 871–

87.

[34] M. Lalpoor, D.G. Eskin, D. Ruvalcaba, H.G. Fjær, A. Ten Cate, N. Ontijt, et al.,

Cold cracking in DC-cast high strength aluminum alloy ingots: An intrinsic

problem intensified by casting process parameters, Mater. Sci. Eng. A. 528 (2011)

2831–2842. doi:10.1016/j.msea.2010.12.040.

[35] T. Iida, R. Guthrie, The Physical Properties of Liquid Metals, Clarendon Press,

Oxford, 1988.

[36] W.F. Gale, Smithells Metals Reference Book, 8th ed., Elsevier, 2004.

[37] C.J. Vreeman, Modeling Macrosegregation in Direct Chill Cast Aluminum Alloys,

Purdue University, 1997.

[38] J. Yanke, K. Fezi, R.W. Trice, M.J.M. Krane, Simulation of Slag Skin Formation

in Electroslag Remelting Using a Volume-of-Fluid Method, Numer. Heat Transf.

Part A Appl. (2015).

68

68

[39] H.K. Versteeg, W. Malalaskekera, An Introduction to Computational Fluid

Dynamics, 2007. http://www.mie.utoronto.ca/labs/mussl/cfd20.pdf.

[40] D.G. Eskin, A. Jafari, L. Katgerman, Contribution of forced centreline convection

during direct chill casting of round billets to macrosegregation and structure of

binary Al–Cu aluminium alloy, Mater. Sci. Technol. 27 (2011) 890–896.

doi:10.1179/026708309X12578491814672.