174
Investigation of Sustainable Machining Strategies for Difficult-to-Cut Materials by Abdelkrem Ali Aboulqasim Eltaggaz A Thesis presented to The University of Guelph In partial fulfilment of requirements for the degree of Doctor of Philosophy in Engineering Guelph, Ontario, Canada © Abdelkrem Ali Aboulqasim Eltaggaz, August, 2019

Investigation of Sustainable Machining Strategies for

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Investigation of Sustainable Machining Strategies for Difficult-to-Cut Materials

by

Abdelkrem Ali Aboulqasim Eltaggaz

A Thesis

presented to

The University of Guelph

In partial fulfilment of requirements

for the degree of

Doctor of Philosophy

in

Engineering

Guelph, Ontario, Canada

© Abdelkrem Ali Aboulqasim Eltaggaz, August, 2019

ii

ii

ABSTRACT

Investigation of Sustainable Machining Strategies for Difficult-to-Cut Materials

Abdelkrem Eltaggaz Advisor:

University of Guelph, 2019 Professor I. Deiab Co-Advisor:

Professor H. Kishawy

Over the past few decades, the demand for employing sustainable machining has increased because of high

competition, and operator health, and environmental concerns. The generated heat is an immense challenge

when machining difficult-to-cut materials such as Austempered Ductile Iron (ADI) and Ti-6Al-4V alloy due

to their poor thermal conductivity. The amount of heat generated significantly affects their machinability.

Application sustainable cooling strategy instead of conventional flood for these materials has become more

desirable, so Minimum Quantity Lubrication (MQL) has been proposed. However, MQL has inefficient

cooling ability. To enhance the cooling and lubricating efficiency of an MQL, nanoparticles can be dispersed

into a base fluid, thereby creating an MQL-nanofluid.

This work attempts to improve the MQL machining performance by adding aluminum oxide (Al2O3)

nanoparticles which have superior tribological and thermal properties. Furthermore, to the best of the

author’s knowledge, there is a gap in the available literature regarding the numerical modelling of the ADI

machining process including the effect of coolant. ADI constitutive equation was coded into ABAQUS

software package and used for heat generation computations at the workpiece-tool interface. The results

were used in a Computational Fluid Dynamic (CFD) model to evaluate the heat removal ability of MQL in

the cutting zone. In this research, the Ti-6Al-4V alloy and ADI were machined at different cutting

iii

iii

parameters with different coolant approaches to optimize the combination of cutting parameters and

sustainable coolant methods. The findings showed that adding Al2O3 nanoparticles enhanced the

convection, wettability, and conduction characteristics of the proposed sustainable MQL-nanofluid

technique, and offered a significant improvement in tool wear behavior, power consumption, and surface

roughness in compared to regular MQL. The results of the CFD model for applying the MQL approach

during the machining of ADI were found to be in acceptable agreement with the experimental results.

This research determines an appropriate sustainable cooling technique to improve the machinability of ADI

and Ti-6Al-4V while considering both the operator's health and the environmental impacts. This work codes

the constitutive equation for ADI and develops both an effective FEM model and CFD simulation for the

machining of ADI using MQL cooling.

iv

iv

Acknowledgments

Thanks and gratitude is due to ALLAH for the continuous support I have felt throughout my life.

The author would like to thank all those who supported him during the course of this research.

Special thanks to my research supervisors Dr. Ibrahim M. Deiab and Dr. Hossam Kishawy for

their guidance and support throughout this study. Thanks to the member of the Supervisory

Committee, Dr. Marwan Hassan. Thanks to Zied Said for developing the project with regards to

the computational fluid dynamics model. Thanks are extended to Barry Verspagen, John Cloutier,

Ken Graham, Hong Ma, John Whiteside, David Wright, and Patricia Zawada for their assistance

in running the experimental tests, and the remainder of the Advanced Manufacturing Lab (AML)

team for their support.

In addition, I am deeply thankful to the Libyan Ministry of Higher Education and Scientific

Research, the Canadian Bureau for International Education (CBIE), and the Zawia Higher Institute

of Science and Technology for their financial scholarship. I would like also to thank the National

Science and Engineering Research Council of Canada (NSERC) and Ontario Centers of

Excellence (OCE) for their support funding this research. I would like to thank Wisam Al-Wajidi,

Dr. Hussien Hegab, and Nihad Alzuabidi for their friendship, their motivation and encouragement,

and their belief in me. Lastly but certainly not the least, I would like to acknowledge my wife and

my beautiful kids for their love and encouragement throughout the work on this research even

during the most trying of times.

v

v

TABLE OF CONTENTS

Abstract ...........................................................................................................................................ii

Acknowledgments……………………................................................................................iv

Table of Contents ………………………………………………………..………………….…..v

List of Tables …………..…………………………………………………………..……….......ix

List of Figures …………………………………………………………………………………...xi

Nomenclature …………………………………………………………………………………..xv

List of Appendices …………………………………………………………………………....xviii

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

1.1 Preamble …………….……………………………………………………….…………1

1.2 Thesis Scope and Aim …….…….…………………………….…………………….…2

1.3 Thesis Objectives ……………...…………………………………………………….…3

1.4 Thesis Layout ………………………….…………………………………….………….5

Chapter 2: Literature Review …………………………………..……………….………….…..6

2.1 Introduction ………………….……………………………………………………….…6

2.2 Literature Review ……………….……………………...……….…………………….11

Chapter 3: Research Methodology.………………………………………….………..…......30

3.1 Introduction …………………………………..……………………….………...….…30

3.2 Design of Experiments (DOE) ……………..…………………….……………....…30

3.3 Planning of Experiments ………………..…………..…………………………..…...31

3.3.1 Cutting Tool and Workpiece materials ………………..…..…………….….32

3.3.1.1 Austempered Ductile Iron (ADI) …………………………..…….……33

3.3.1.2 Ti-6Al-4V Titanium Alloy ………..………………………....…….……34

3.3.1.3 Turning Cutting Tool Insert ………..….………………..………..……34

3.4 Experimental Setup ………………..…….…………………...……….……….……36

3.5 Measurement of Machining Variables ………..…………………………….….….39

vi

vi

3.5.1 Cutting Forces …………………………..………………………….…....…...39

3.5.2 Tool Wear and Wear Mechanisms ……………………….…….…..……....40

3.5.3 Machined Surface Roughness (Ra) ……………….………….……...….....43

3.5.4 Power Consumption ……………………….……….…………..……............44

3.6 Numerical Methodology …………………………………….…………….………….45

3.6.1 Design of Experiment …………………………...…………………….…..…..46

3.6.2 Experimental Setup ………………………………...……….…….……….….47

3.6.3 Thermocouple and Temperature Arrangement ……………….……………48

Chapter 4: Results and Discussion …………………………………….……………...……..52

4.1 The Effects of Using Different Coolant Strategies on Tool Life and

Surface Roughness …………………………….…………………………………....52

4.1.1 Introduction …………………………………….………………….………...…52

4.1.2 Results and Discussions ……………………………………….………….....54

4.1.2.1 Tool Life and Surface Roughness ………………...……………...........54

4.1.2.2 ANOVA Analysis ……………………………………….………...……...59

4.1.2.3 Response Surface Methodology (RSM) …….………..…….………....63

4.1.2.4 Wear Mechanisms Analysis ………….….………….……….………....69

4.1.3 Conclusion ……………………………………….…….……………...........…73

4.2 Hybrid Nanofluid-MQL Strategy for Machining ADI …….………………………...75

4.2.1 Introduction ……………………………………………….………….…………75

4.2.2 Results and Discussions ……………………………….……………………...76

4.2.2.1 Tool Wear ….………………………………………………………….….76

4.2.2.2 Nanoparticle Effects ……………………….…………….……………...79

4.2.2.3 Scanning Electron Microscope (SEM) Investigation ……………..…..81

4.2.2.4 Nanofluid-MQL Mechanism ………………………………....……….…84

4.2.3 Conclusion ………………………………………………………..…………….85

4.3 - Application of MQL with Nano-Fluids during Machining of Ti-6Al-4V Alloy:

vii

vii

Experimental Investigation and Analysis ………………………………………...87

4.3.1 Introduction …………………….……………………………….…….…...……87

4.3.2 Experimentations and Methodology...……………………….……………….90

4.3.3 Results and Discussion …………………………………….….……………....93

4.3.4 Tool Wear Mechanisms …….………………………….…………..……….…99

4.3.5 Nanofluid-MQL Mechanisms …….………………….……….…….…..……102

4.3.6 Conclusion …………………………………………….….….………………..103

Chapter 5: Sustainability Assessment of Machining Difficult-to-cut Materials ……….…105

5.1 Introduction …………………………………………………………..……………..105

5.2 Assessment Results …………………………..…………….…………………….108

5.3 Conclusion …………………………………………………………….……………110

Chapter 6: Modeling and Numerical Simulation …………………………………….……..112

6.1 Introduction ……………………………………………………………….………...112

6.2 Experimental Setup …………………………………………….………………….114

6.3 Finite Element Modeling ……………………….…………………….……………118

6.3.1 ADI Model for Orthogonal Turning Process …………………....….……….118

6.3.2 FEM Results ……………………..……………………………...………..…..121

6.4 CFD Modeling …………….…………………………………………….….………123

6.4.1 CFD Results …………………………………………...…………......……….123

6.4.1.1 Geometry for CFD Model ….………………………………….……….123

6.4.1.2 Meshing ………………….….…………………….………………….…124

6.4.1.3 Meshing Approach ………….……………………………………....…125

6.4.2 CFD Results under Dry and MQL Conditions ………….………..…....…...125

6.5 Conclusion ….…………….………….……….……………………….…………...130

Chapter 7: Conclusions and Future Work ………………………….……………………...131

7.1 Conclusions .………………………..…………………………………….……......133

7.2 Future work …………….……………………….…………………….…….……...133

viii

viii

References ………………………………………………...……………………………....…135

Appendix I …………………………………………………………………………..………...151

Appendix II ……………………………………………………………….…………………...156

ix

LIST OF TABLES

Table 2.1: Summary of available research on MQL and MQCL coolants ……………..…..27

Table 2.2: (a) Summary of available research on nanofluid coolants ……………………….28

Table 2.2: (b) Summary of available research on nanofluid coolants ………………………29

Table 3.1: Percent weight chemical compositions of ADI grade 2 (wt %) …………….…..33

Table 3.2: ASTM A897/897M-02 mechanical properties of ADI grade 2 ………….……….33

Table 3.3: Chemical compositions of Ti-6Al-4V (wt %) …………………………………...….34

Table 3.4: Ti-6Al-4V titanium alloy mechanical properties (at room temperature) …….…34

Table 3.5: Specifications of the base vegetable oil utilized in the MQL system ………..…39

Table 3.6: Specifications of the Kistler dynamometer …………………………………..…...40

Table 3.7: The Mitutoyo tool maker microscope information ………………………….….…41

Table 3.8: Specifications of the Mitutoyo surface roughness tester SJ 201P ………..……44

Table 3.9: PS 2500 power data logger specifications ………………………………….….…45

Table 3.10: Cutting tool (CCMT 12 04 04-KM) thermo-mechanical properties …………….46

Table 3.11: Specifications of Omega HH 374 data logger ………………………………..…49

Table 4.1: Assignment of levels to the cutting process variables …………………………...53

Table 4.2 L24OA experimental results ……………………………………………………......55

Table 4.3: ANOVA results for average surface roughness (Ra) …………………………….60

Table 4.4: ANOVA results for tool life ……………………………………….……………..…..62

Table 4.5: The coefficient of correlation results …………………………………………..…..65

x

x

Table 4.6: The studied parameter levels and the plan of the experiments …………….….77

Table 4.7: L9OA with respect to the studied design variables ……………..……………….93

Table 4.8: The machining parameter levels and the output measurements ………………94

Table 4.9: Analysis of variance for maximum flank wear ………………………………..…..95

Table 4.10: Analysis of variance for surface roughness ………………………………..……96

Table 4.11: Analysis of variance for power consumption ………………………………..…..98

Table 5.1: Case 1: Equal weighting factors for all sustainable metrics and indicators…...106

Table 5.2: Case 2: Weighting factors for all sustainable metrics and indicators ……..…..107

Table 5.3: Case 3: Weighting factors for all sustainable metrics and indicators ……..…..107

Table 5.4: Case 4: Weighting factors for all sustainable metrics and indicators ….….…..108

Table 5.5: (a) Case (1): Sustainable factors and sustainable indexes ……………..……..109

Table 5.5: (b) Case (1): Weighted sustainable index ………………..……………….……..110

Table 5.6: Assessment algorithm outputs versus optimal machining experimental

outputs …………………………………………………………………………...….110

Table 6.1: Machining parameters and equipment specifications …………………….……115

Table 6.2: Thermo-mechanical properties of the ADI grade 2 …………………….……….118

Table 6.3: Thermo-mechanical properties of CCMT 12 04 04-KM (uncoated carbide)….118

Table 6.4: Parameters for the constitutive material law ……………………………...……..120

xi

xi

LIST OF FIGURES

Figure 1.1: Pillars of sustainable machining ……………………………………..………...…7

Figure 2.2: Financial costs and environmental impacts for using conventional cutting

fluid …………………………………………………………………………………..8

Figure 2.3: Applications of ADI in the marketplace …………………………..………..……9

Figure 2.4: Titanium components of an airplane jet engine (turbofan) ……………..……10

Figure 2.5: Heat generation during the machining process ………………..……………..15

Figure 3.1:CNMG120416MR (ISO) cutting tool insert ……………..………………………35

Figure 3.2: CNMT 12 04 08-KM H13A (ISO) cutting tool insert ……………………….…..35

Figure 3.3: CNC-lathe machine 15HP ST-10 HAAS ……………………………………….37

Figure 3.4: Experimental setup ……………………………………………………...……….38

Figure 3.5: Schematic experimental setup for machining ADI under different cooling

Strategies …………………………………………………………………………38

Figure 3.6: Cutting forces data evaluation system: (a) Kistler 9257b dynamometer,

and (b) Kistler 5070 charge amplifier …………………………………………40

Figure 3.7: (a) Tool maker microscope brand Mitutoyo (b) Standard measurement

of the maximum flank wear (VBmax) on the tool flank face …………...….….41

Figure 3.8: Scanning Electron Microscope (SEM) ………………………...……………….42

Figure 3.9: SEM images for a coated carbide tool (CNMG120416MR) after machining

ADI at V=120 m/min and f=0.3 mm/rev under dry machining ……...………...43

Figure 3.10: Mitutoyo surface roughness tester SJ 201P ……………………...…………..44

Figure 3.11: Power sight ………………………………………………..…………………….45

Figure 3.12: Experimental schematic ………………………………………………..……...47

xii

xii

Figure 3.13: Experimental setup (a), and temperature measuring arrangements (b),

(c) and (d) ………………………………………………………...……………...47

Figure 3.14: Experimental data acquisition flow diagram ………………………..………..49

Figure 3.15: Keyence VH-Z250R microscope ………………………..…………………….50

Figure 3.16: Thesis methodologies ………………………………………...………………..51

Figure 4.1: Experimental setup schematic …………………………………………...……..53

Figure 4.2: Cutting condition effects on tool life …………………………………………….56

Figure 4.3: Cutting condition effects on machined surface quality ………………...……..56

Figure 4.4: The interaction effect between the lubrication strategy and feed rate for

the average surface roughness results ………………………..……………....61

Figure 4.5: The interaction effect between the lubrication strategy and feed rate for the

tool life results …………………………………………………………………….63

Figure 4.6: ADI-dry cutting model for surface roughness results ……………………..….65

Figure 4.7: ADI-dry cutting model for tool life results ………………………..…………….66

Figure 4.8: ADI-flood cutting model for tool life results ………………………..………….66

Figure 4.9: ADI-flood cutting model for surface roughness results ……….……….………67

Figure 4.10: ADI-MQL cutting model for surface roughness results ………………….…..67

Figure 4.11: ADI-MQL cutting model for tool life results ………………………….….….…68

Figure 4.12: ADI-Nanofluid cutting model for tool life results ……………………….….…68

Figure 4.13: ADI-Nanofluid cutting model for surface roughness results ………..………68

Figure 4.14: EDS image for diffusion on the tool at dry machining f=0.2 mm/rev and

V=120 m/min ……………………...………………………..…………………....70

Figure 4.15: SEM images of inserts after machining at V=120 m/min and f=0.2 mm/rev

under different coolant strategies ………………………………………..…....71

Figure 4.16: SEM images of inserts after machining at V=120 m/min and f=0.3 mm/rev

xiii

xiii

under different coolant strategies …………………………………..………....73

Figure 4.17: Thermal conductivity results (Al2O3 nanofluid vs. vegetable oil) ………......76

Figure 4.18: Tool wear results ……………………………………………………………......78

Figure 4.19: Flank wear results at a feed rate of 0.2 mm/rev …………………….…….…79

Figure 4.20: Flank wear results at a feed rate of 0.3 mm/rev ………………………………79

Figure 4.21: The MQL-nanofluid mechanism schematic …………………………..……...80

Figure 4.22: Scanning Electron Microscope (SEM) images at a cutting speed of

180 m/min and a feed rate 0.2 mm/rev using MQL-Al2O3 nanofluid (a)

and classical MQL (b) …………………………………………..……………....82

Figure 4.23: The main tool wear modes when machining ADI at a cutting speed of

120 m/min and a feed rate of 0.3 mm/rev using (a) MQL-nanofluid, and

(b) pure MQL …………………………………………………………………..82

Figure 4.24: Images of (a) an AQUASONIC-50HT device used to achieve the

dispersion, and (b) a regular and a nanofluid-MQL …………………..……..91

Figure 4.25: The experimental schematic ………………………………...……………..…92

Figure 4.26: The main effects for the maximum flank wear (VB) ……………….………..95

Figure 4.27: Nanofluid-MQL mechanism schematic …………………………..…………..96

Figure 4.28: The main effects plot for surface roughness (Ra) ……………………...……97

Figure 4.29: The main effects plot for power consumption ……………………..…………98

Figure 4.30: Coefficient of friction and induced wear versus nanoparticle content...…….99

Figure 4.31: SEM and EDS pictures of the tool after using regular MQL ………………..100

Figure 4.32: SEM and EDS pictures of the tool after using MQL-Nanofluid 2 wt. % ……101

Figure 4.33: SEM and EDS pictures of the tool after using MQL-Nanofluid 4 wt. % ……101

Figure 4.34: The nanofluid-MQL mechanisms (a) rolling effect and (b)

ploughing effect ……………………………………………………………….103

xiv

xiv

Figure 6.1: Methodology schematic of the thermal behavior prediction for

machining ADI …………………...…………………………………...………...115

Figure 6.2: Experimental setup view (a); the installation and temperature measuring

arrangements (b), (c) & (d) ………………………………...…………………..117

Figure 6.3: Meshing and boundary conditions definition …………………………..……..121

Figure 6.4: (a) Finite element simulated cutting temperature (b) Temperature values

across the rake face (oC) …………………………………………..…………122

Figure 6.5: The heat flux contours ……………………………………………..…………...122

Figure 6.6: MQL versus other cooling methods in terms of heat transfer

Mechanisms .................................................................................................123

Figure 6.7: Geometry setup for the CFD analysis …………………………………… …..124

Figure 6.8: Meshing setup for the CFD analysis ………………………………...………...125

Figure 6.9: Velocity contours (a), and heat transfer coefficient distribution (b) …………126

Figure 6.10: ANSYS Fluent/ANSYS transient thermal coupled link …………..…………127

Figure 6.11: Comparison between the experimental and the simulation temperatures

under dry machining conditions …………………………..…………………128

Figure 6.12: Comparison between the experimental and simulation temperatures

under MQL conditions ……………………………………..…………………129

xv

xv

NOMENCLATURES

ADI Austempered ductile iron

Al2O3 Aluminum oxide nanoparticles

ALE Arbitrary Lagranian Eulerian

ANOVA Analysis of variance

BUL Built-up-edge layer

CAMQL Chilled air minimum quantity lubrication

CFD Computational fluid dynamic

CNC Computer numerical control

CNTs Carbon nanotubes

CO2 Carbon dioxide

c Material specific heat

DOE Design of experiments

DPM Discrete phase model

EDS Energy dispersive X-ray spectroscopy

EHS Environmental health and safety

EPA Environmental protection agency

f Feed rate

FEA Finite element analysis

FEM Finite element model

HSS High speed steel

KN Kilo newton

Kp Thermal conductivity of nanoparticle

Kf Thermal conductivity of base fluid

Km Thermal conductivity of resultant

GPa Gage Pascal

J Joule

L24OA 24 orthogonal array

L9OA 9 orthogonal array

MPa Mage Pascal

xvi

xvi

MQL Minimum quantity lubrication

MQCL Minimum quantity cooling lubrication

MWCNT Multi walled carbon nanotubes

MWF Metalworking fluid

NDM Near-dry machining

NFMQL Nanofluid minimum quantity lubrication

NIOSH National institute for occupational safety and health

OSI Overall sustainable index

OStc Operator personal health, express the toxic effect

OShtc Operator personal health, high temperature surface exposure

Pn Operator personal health, noise effect

PVD Physical vapor deposition

RSM Response surface methodology

Ra Surface roughness

SDS Sodium dodecyl sulfate

SEM Scanning electron microscope

T Temperature

TRIP Transformation induced plasticity

SI Sustainable index

SiC Nano-silicon carbide

v Cutting speed

Vf Nanoparticles volume fraction

VB Maximum flank tool wear

Wt. % Weight percentage

WSI Weighted sustainable index

ρ Material density

η Inelastic heat fraction

σ Flow stress

εpl Plastic strain

ε Strain rate

ε Strain

xvii

xvii

γ Rake angle

θ Clearance angle

m Frictional coefficient

T Temperature rate

χ Ductile softening term

K1 Material constant

K Material constant

εG Limit strain

εG Limit strain rate

T0 Base temperature

Tm Melting temperature

n Hardening exponent

β Material constant of the temperature function

ψ(T) Thermal softening term

γ Rake angle

θ Clearance angle

xviii

xviii

LIST OF APPENDICES

Appendix I ………………………………………………………………………………...144

Appendix II ………………………………………………………………………………..149

1

Chapter 1: Introduction

1.1 Preamble

In metal cutting, the cutting fluid and design machining parameters impact the machining outputs

such as cutting forces, surface quality, cutting temperature, and friction coefficient. During

machining, heat is generated in three cutting regions: the primary shear region, the secondary

deformation region, and the tertiary region (tool flank-workpiece contact). Approximately 2/3 to

3/4 of the heat is generated by friction force as the chips slide along the cutting tool face, and

through the plastic deformation of the metal. The heat that is generated becomes a more significant

issue when machining difficult-to-cut materials such as Ti-6Al-4V titanium alloy and austempered

ductile iron (ADI). Cooling and lubrication during the machining process can be provided by

cutting fluids. Using a cutting fluid decreases the tool wear and extends the tool life. In the

manufacturing industry, the use of conventional flood coolant is a common cooling strategy.

However, due to its relatively high cost and the negative implications to the operator`s health and

the environment, sustainable cooling approaches have become more desirable. One of these

approaches is Minimum Quantity Lubrication (MQL), which has been determined to be an

effective near dry machining method. Its cooling capacity, however, is relatively low compared to

a conventional flood coolant. This study aims to improve the MQL thermal performance in the

cutting zone with respect to cutting forces, power consumption, tool life, and surface roughness

during the turning of T1-6Al-4V alloy and austempered ductile iron (ADI).

To the best of the author's knowledge there is a gap in the available literature regarding FE models

of ADI Machining due to the unavailability of its constitutive equation.

2

2

1.2 Thesis Scope

The application of sustainable cooling strategies in metal cutting processes has always been of

attentiveness. In metal cutting, the cutting fluids consumption and the power (energy) requirements

are linked to the environmental effects of the cutting practice. Difficult-to-cut materials such as

austempered ductile iron (ADI) and Ti-6Al-4V titanium alloy are preferred over other metals such

as aluminium and steel alloys. This is due to their unique properties, which include a high strength

to weight ratio, fatigue resistance, and excellent corrosion resistance. ADI has become widely used

in the automotive industry, and titanium alloys are used in biomedical and aerospace

implementations due to their resistance to crack growth, resilience to creep, and fatigue behaviour.

Second chapter presents a literature review of the study performed on the sustainability features of

the machining of ADI and Ti-6Al-4V alloy with a focus on investigating the near dry and the

environmental friendly cooling approach. The literature showed that research has focused on the

coolant approaches, namely flood cooling, cryogenic, minimum quantity lubrication (MQL),

minimum quantity cooling lubrication (MQCL), and nanofluid-MQL. The literature also presented

a research gap in that the constitutive equation for ADI is not available by default because of its

contradictory nature. That is why there have been no studies in which an FE model of machining

simulations has been performed for ADI. Also, insufficient research has been attempted to

incorporate finite element models (FEM) and computational fluid dynamic (CFD) modelling in

cutting operations to investigate the interactions of MQL technique during the machining of ADI.

This research envisages covering the gap in the available literature on the use of nanofluid-MQL

as a potential environmentally sustainable cooling approach for machining difficult-to-cut

materials, such as ADI and Ti-6Al-4V titanium alloy. In manufacturing processes, efficient

machining can significantly decrease the costs and the environmental impacts of the product. The

3

3

current study also intends to address the potential applications of FEM and CFD simulations to

facilitate the use of a sustainable cooling approach (MQL) in the cutting of ADI.

1.3 Thesis Objectives

Cooling strategies have a controlling effect on the performance of the cutting process. There has

been limited research on how to enhance the efficiency of coolant strategies and optimize the

parameters. To that end, an experimental study and a numerical model of new coolant strategies

when machining difficult-to-cut materials would increase the potential for more economic,

efficient and environmental machining processes. This research thesis has the following aims:

1. The terms “environmental" and "sustainability” are liberally used to refer to efficient power

consumption, the use of a minimal volume of coolant (litre/min in flood to mill-litre/hr in

MQL), and the use of vegetable oils. MQLs have a niche of applications that would prove

to be beneficial both economically and environmentally. However, the open literature still

requires more research and experimental data on the effects of using different cooling

approaches on the process performance during the machining of difficult-to-cut materials.

In particular, MQL and MQL-nanofluid hybrid strategies must be addressed, which is one

of the objectives of this thesis.

2. This work assesses the performance of an MQL combined with nanofluids, i.e. a nanofluid-

MQL. The MQL method has an inefficient cooling capacity compared to standard flood

coolant, particularly when machining difficult-to-cut materials such as ADI and Ti-6Al-4V

alloy. Therefore, numerous researchers have attempted to improve the MQL’s cooling

capacity. In this work, nanoparticles (Al2O3) are added to the MQL base oil fluid in order

to enhance its cooling and lubricating properties.

4

4

3. Wear progression and mechanisms when machining difficult-to-cut materials (e.g. ADI

grade 2 and Ti-6Al-4V titanium alloy) are identified. Understanding the tribological

behaviour and characteristics of the MQL media and nano-additives during the machining

of difficult-to-cut materials is crucial to optimizing the process, and needs further

investigation.

4. Numerical models are being developed to better simulate the machining environment. One

of the biggest challenges in the metal cutting process is to predict the cutting temperature

either experimentally or numerically. High temperature in the cutting zone has a

controlling impact on tool wear and the resulting surface integrity of the machined surface.

Many researchers have investigated different experimental and numerical approaches to

accurately predict the cutting temperature. An accurate prediction of the cutting

temperature is one of the greatest factors that control the accuracy of the predicted cutting

forces and tool wear. Developing a model that accounts for both the thermal and the

mechanical sides of the ADI machining process has yet to be realized. To the best of the

researcher’s knowledge, the available literature has rarely addressed the simulation of heat

transfer analysis in the cutting zone for cutting ADI grade 2. However, this thesis develops

a model that utilizes both finite elements (FEA) and computational fluid dynamics (CFD)

to simulate the influence of an MQL cooling media on the machining environment of ADI

grade 2. The results of this work were promising, and prove that this concept is possible.

The work sheds some light onto the challenges that face such a model. This work is the

foundation upon which an accurate prediction of the temperature distribution in the cutting

zone will be achieved, as it is one of the objectives of the modelling component of this

research thesis.

5

5

1.4 Thesis Layout

The research study is divided into six chapters and follows a sandwich style, in which Chapter 4

represents the published articles. The thesis outline is as follows:

Chapter 1 introduces the scope and objectives of the thesis, and outlines the overall structure of

this research.

Chapter 2 reviews the existing state of difficult-to-cut materials in the manufacturing industry,

the sustainable coolant strategies used, and the gaps in the available research on the expanding use

of alternative coolant strategies. This leads into a discussion of how minimum quantity lubrication

performs under different machining criteria. The review continues by discussing the issues of

MQL, with alternative options to improve its performance by adding nanoparticles.

Chapter 3 details the experimental setup adapted to conduct the intended study. The machining

experiments are conducted on Austempered Ductile Iron (ADI) grade 2 and Ti-6Al-4V titanium

alloy.

Chapter 4 represents and discusses the findings and conclusions obtained from the published

papers.

Chapter 5 presents the assessment study for machining Ti-6Al-4V alloy.

Chapter 6 discusses the numerical outputs obtained from the finite element model (FEM) and the

computational fluid dynamics model (CFD), as well as the experimental temperature.

Chapter 7 provides the conclusions and the recommended future work to further improve the

practicality and performance of the CFD mode

6

6

Chapter 2: Literature Review

2. 1 Introduction

For decades, machining processes have been playing a crucial role in the global economy.

However, due to a combination of high competition in the industrial marketplace and

environmental concerns, manufacturing industries have recently been focusing on minimizing

machining costs, increasing the quality of the product, and reducing their ecological footprint.

Sustainable manufacturing is a novel technique for producing product in an efficient manner.

Sustainability has many descriptions and various pillars. Figure 2.1 presents the three major pillars

of sustainable manufacturing: economy, society, and environment [1]. In metal removal processes,

the dry condition is considered to be environmentally friendly machining because no cutting fluid

is used. However, dry machining often results in problems by cause of the high amount of heat in

the cutting zone, especially for difficult-to-cut materials.

In order to improve tool life and machined surface quality during the machining process, cutting

fluids are utilized extensively to reduce the heat generated at the cutting zone [2]. However, the

excessive use of conventional cutting fluids in flood cooling and aerosols has been found to be the

source of greenhouse gas, which negatively affects the operator’s health and environs [3]. Disposal

of coolants contaminates the environment because these substances contain machined particles

and hazardous chemicals [4]. Coolant disposal, supply, storage, and maintenance processes

increase the total production costs (Figure 2.2) [5]. Thus, concerns about environmental issues,

unfavourable impacts on the operators’ health, and increased total production costs, have

highlighted the importance of decreasing or eliminating the cutting fluids application. Some

government regulations have been established; for example, the National Institute for

7

7

Occupational Safety and Health (NIOSH) in North America has established regulations that

recommend occupational exposure to metalworking fluid (MWF) aerosols be limited to 0.5 mg/m3

instead of 5 mg/m3 [6]. Also, the US Environmental Protection Agency (EPA) has provided

monitoring instruments for the industry to develop sustainable machining processes [7].

Figure 2.1 Pillars of sustainable machining. [1]

8

8

Figure 2.2 Financial costs and environmental impacts for using conventional cutting fluid.

As mentioned previously, the use of cooling strategies is important during the metal removal

process to reduce the heat throughout the work material, thereby decreasing the tool wear rate and

enhancing the machining performance. This need for a cooling strategy becomes more significant

when machining difficult-to-cut materials, such as ADI and Ti-6Al-4V titanium alloy, due to their

poor thermal conductivity. The excessive heat generated during machining negatively influences

the tool life, surface quality, and power consumption.

Recently, the use of ADI has been increasing rapidly in many engineering applications, such as

the defense, oil and gas, automotive, mining, and agricultural industries (Figure 2.3), due to the

unique and promising characteristics of ADI. These include a high strength to weight ratio, high

wear and corrosion resistance, high yield stress, and high toughness [8-10]. For example, ADI is

9

9

currently used to fabricate heavy equipment parts that require a long service life, such as

suspensions and crankshafts; these parts must meet exacting specifications in terms of fatigue

strength [11]. The outstanding properties of ADI can be attributed to its unique ausferrite

microstructure, which results from a heat treatment operation [12]. From the metallurgy

perspective, ADI is classified into different grades based on its mechanical properties. The

classification of ADI grades depends on the composition of the alloying elements, heat treatment

temperature, and heating and quenching times [12-14]. ADI Grade 2 has become the most

commonly used material in the manufacturing of vehicle parts due its remarkable properties.

However, since the first use of ADI in engineering applications, many challenges have been

experienced in attempting to machine this material successfully. ADI has poor machinability due

to transformation induced plasticity (TRIP) and the formation of martensite during its machining

process [15]. The difficulties that arise when cutting ADI limit its practical applications.

Figure 2.3 Applications of ADI in the marketplace. [16]

10

10

Figure 2.4 Titanium components of an airplane jet engine (turbofan). [17]

Titanium alloys have served as effective replacements for steel and for aluminum alloys in many

engineering applications due to their wear resistance, high strength at high temperatures, and

resistance to chemical degradation. Titanium alloys, when used in an airplane jet engine, save

weight and/or space, and increase the engine efficiency as a result of their ability to work

effectively at high temperatures (Figure 2.4) [18]. The Ti-6Al-4V titanium alloy is commonly used

in gas turbine, marine, biomedical and sports equipment applications [19]. However, other

properties, such as high chemical reactivity and poor thermal conductivity, negatively affect the

machinability of Ti-6Al-4V and accelerate tool failure. Therefore, machining such difficult-to-cut

materials results in a shorten tool life, a long machining cycle, and high generated heat in the

cutting zone.

11

11

2.2 Literature Review

Machining of difficult-to-cut materials is difficult, and the appearance of severe adhesive and

abrasive wear on a tool's cutting edge during the machining process shortens the tool's life. As a

result, research has investigated the factors that influence the machinability of ADI and titanium

alloys. Akdemir et al. [20] investigated the effect of depth of cut and cutting speed on the

machinability of ADI. The study evaluated ADI machinability by measuring surface roughness,

cutting forces, and flank wear. Experimental data and a response surface methodology (RSM)

technique were utilized to expose the influence of machining factors on ADI machinability. The

investigation found that the depth of cut has an insignificant influence on cutting forces. The study

also indicated that the cutting speed was a significant parameter that affected tool life (flank wear),

tangential cutting force, and surface roughness.

Parhad et al. [21] investigated the impact of depth of cut on ADI machinability in terms of

machined surface quality. It was observed that increasing the cutting speed improved the machined

surface quality. The surface roughness increased as the cutting speed decreased. This study also

concluded that the best machined surface quality was obtained when ADI was turned at a speed

between 150-200 m/min, a feed rate of 0.1 mm/rev, and a depth of cut of 2 mm.

Hegde et al. [22] investigated the influences of heat treatment and cutting parameters on a

manganese ADI alloy under dry condition. It was found that the cutting speed and austempering

temperature have the greatest effects on tool wear, whereas the feed rate has the most significant

effect on the surface roughness. Meena et al. [23] also drilled ADI under dry machining to study

the influence of cutting parameters on tool wear, cutting forces, chip morphology, and machined

surface integrity. The findings showed that a combination of a higher speed and a lower feed rate

resulted in higher cutting forces. Polishetty [24] also found that during the machining of ADI,

12

12

cutting parameters controlled the mode of tool failure, such as built-up-edge, tool wear, chipping,

and tool damage, while the surface roughness was significantly affected by the chip morphology

and tool wear.

Arft et al. [25] evaluated the machining aspects of ADI. This study found that the cast process of

ADI has a major impact on its machining operation; hence, an appropriate tool selection should be

taken into an account when machining ADI. It was also observed that from a metallurgy

perspective, increasing the percentage of element alloy, such as molybdenum, negatively affected

the tool life. Polishetty et al. [26] revealed that during the machining of ADI, the high rate of plastic

deformation, the high heat generated in the cutting zone, or combination of both are responsible

for the resulting strain-induced transformation which influences the machinability of ADI.

In another study, Meena et al. [27] investigated the effect of the austempering time on the

microstructure transformation, the mechanical properties, and the strain-hardening behavior of

ADI. It was reported that the austempering process time period affected both the strength and the

elongation of ADI. Austempering the material for shorter time periods resulted in lower strength

values and higher elongation. However, a longer time period for austempering increased the

strength value and reduced elongation. Sinlah et al. [28] also revealed that the mechanical

properties and microstructure of ADI significantly influenced tool wear and cutting forces.

The machinability of titanium alloys is affected by several factors, such as the properties of the

alloy, the cutting design factors, the geometry and properties of the cutting tools, and the cooling

approaches [29]. As mentioned previously, properties such as low thermal conductivity and high

chemical reactivity negatively affect the machinability of titanium. Hartung et al. [30] pointed out

that the cutting speed rapidly influenced the crater tool wear when machining titanium. This type

13

13

of tool wear was due to both the low thermal conductivity of titanium and the chemical reaction

between the titanium and the cutting tool material. Maekawa et al. [31] also found that crater wear

increased with an increase in cutting speed because the increased speed resulted in a higher cutting

temperature.

Tugrul et al. [32] reported that a better tool life was obtained when machining titanium by

combining a lower speed and a lower feed rate. Ramana et al. [33] also reported that the cutting

speed had a major effect on the reduction in the tool wear rate, while the feed rate had a major

effect on the surface roughness. This idea was supported by Ojolo et al. [34] when machining

titanium under dry machining using three different cutting tools (HSS, tungsten carbide, and

DNMG carbide) to examine the effect of cutting parameters on tool life. This study revealed that

the tool life was most affected by the cutting speed, followed by the feed rate, and the depth of cut.

The authors concluded that machining titanium at a low cutting speed with a suitable feed rate

should extend tool life.

Narutaki et al. [35] provided experimental results which indicated that coated carbide tools showed

a better surface roughness when compared to other cutting tools when machining a titanium alloy.

That was supported by Liu et al. [36] when machining a Ti-6Al-4V titanium alloy at different

cutting parameters using MQL vegetable oil. The analysis indicated that the feed rate was the main

factor influencing cutting forces and surface roughness. These studies concluded that cutting speed

significantly affects the wear rate and tool life, while surface roughness is mainly influenced by

the feed rate.

Narasimhulu et al. [37] machined titanium Ti-6Al-4V with a PVD coated tool at various cutting

parameters using dry machining. This study also found that the feed rate and the cutting speed

14

14

were the most significant factors that can influence the surface roughness. Ying-lin et al. [38]

reported that when machining titanium, the elevated heat in the cutting zone can lower the quality

of the finished surface due to chips melting and adhering to both the tool face and the machined

surface.

The dissipation of the generated heat from the cutting zone is limited in ADI and Ti-6Al-4V

titanium alloy due to the low thermal conductivity of these materials, resulting in poor

machinability. The heat is generated by the plastic deformation (primary zone), sliding chips effect

(secondary zone), and the friction between the tool surface and workpiece surface (tertiary zone)

as shown in Figure 2.5) [39, 40]. Optimizing the cutting parameters is difficult when machining

ADI and Ti-6Al-4V titanium alloy since the heat generated in the cutting zone is not efficiently

reduced by removing the chips, as is the case when machining other materials [41]. Instead, the

heat settles in the tool tip, thereby accelerate the cutting tool failure. The high temperatures of the

cutting tool, and the heat generated in the primary shear zone at the tool–chip interface, make the

machinability of these materials challenging. Shaw et al. [42] reported that during the machining

process, the majority of the supplied energy (approximately 80%) was used for material

deformation in the primary and secondary zones, while the rest of energy was wasted because of

sliding friction happening at the rake and flank faces. Therefore, many studies have attempted to

improve the performance of coolant strategies in order to reduce the influence of the high

temperature generated at the cutting zone when machining difficult-to-cut materials. The coolant

not only reduces the heat but also decreases the friction between the cutting tool and the workpiece

material [43]. In the material removal process, employing an effective cutting fluid is important

for increasing tool life, creating a high material removal rate (MRR), and cooling and lubricating

the cutting zone.

15

15

Figure 2.5 Heat generation during the machining process.

The effectiveness of the machining process often depends on the use of cutting fluid to minimize

the generated cutting forces and reduce the cutting temperature. One wet machining strategy uses

a standard flood coolant as a cooling technique, based on the assumption that a large quantity of

coolant will enhance the heat removal capacity during the machining process. The idea behind this

assumption is to flush a huge amount of coolant and lubricant over the cutting zone to lower the

generated cutting temperature. Although flood coolant has a noticeable ability to reduce the heat

in the cutting zone, its negative health and environmental impacts and the cost of the amount of

cutting fluid used, make it an unsustainable cooling strategy. In terms of sustainable and

economical machining, combining the choice of proper cutting parameters and the cooling

approach is crucial for new manufacturing. In machining Ti-6Al-4V and ADI, the main challenges

for the cutting fluids are minimizing the frictional effect and reducing the cutting temperature

while having low environmental effect. Many research works have thus proposed sustainable

coolant strategies to maintain a reasonable tool life and cost when compared to those achieved by

flood or any other form of wet cutting.

16

16

Dry condition is another method of machining; it is environmentally friendly because no coolant

is used during the machining process. Also, it does not contaminate chips and workpieces, and

thereby reduces the chips’ disposal cost and limits the development of respiratory and skin diseases

for the operator(s). However, dry machining often results in problems due to excessive heat at the

tool-workpiece and the tool-chip interfaces [44]. This becomes more significant when cutting

difficult-to-cut materials, such as ADI and Ti-6Al-4V titanium alloys, in which the amount of heat

generated significantly, affects tool life even when using a coated protective layer [45, 46]. Dry

machining therefore has many challenges and it has been successfully applied to only a small

number of materials.

Many researchers have conducted experiments in order to reduce the high heat generated during

the machining of difficult-to-cut materials, and to improve their machinability [47-51]. Massive

benefits in cooling and lubricating methods have been discussed in the available literature, such as

minimum quantity lubrication (MQL), compressed air cooling, cryogenic cooling, dry machining,

and nanofluid lubrication [52-55]. Accordingly, the machining of difficult-to-cut materials

requires to be systematically investigated while considering operator health and environmental

impacts to achieve the most reasonable machining strategy. Several researchers have investigated

novel formulations of cutting fluid, and have applied alternatives to the conventional flood

technique for cooling strategies.

An interesting technology, which is increasingly being employed in metal removal processes, is

Minimum Quantity Lubrication (MQL). In the implementation of cutting fluids, the MQL

technique is considered to be a sustainable strategy and a good alternative cooling strategy to

conventional cutting fluids. The MQL method using vegetable oil has provided the most promising

environmental solution compared to cooling and lubrication methods such as compressed air

17

17

cooling and dry machining [56]. MQL is a sustainable and cleaner cooling/lubrication technique

due to the use of small amounts of vegetable oil during the machining process, which is then

disposed of in an environmentally friendly manner. In addition, MQL has the potential to balance

between dry machining and flood techniques [57]. MQL can also be identified as near-dry

machining (NDM) or micro-lubrication [57, 58].

Klocke et al. [59] documented MQL as a lubricating and cooling method that used less cutting

fluid: about three to four orders of magnitude lower than a flood coolant at a flow rate of 50 to 500

ml/h. In comparison, Tschatsch and Reichelt [60] have reported an MQL of 50 ml/h and 2000

ml/h. This is still very low in comparison with regular flood cooling, which is approximately

1200x103 ml/h [61]. The ability of the lubricant to penetrate into the chip/tool interface enables

the MQL to reduce the cutting temperature and friction, thereby extending tool life. These factors

will demand that researchers focus on investigating the machining performance of MQLs,

particularly during the machining of difficult-to-cut materials such as titanium and ADI.

Many studies have reported that MQLs improved machining performance in terms of machined

surface quality and tool wear compared to dry machining [48, 62, 63], as well as reduced the costs

of both machine cleaning and the disposal process [64]. Ramana et al. [33] performed a study of

titanium Ti-6Al-4V alloy at different cutting parameters under MQL, flood, and dry machining to

investigate the influences of cutting parameters and coolant strategies on machinability. This study

indicated that the MQL approach provided a better performance when compared to flood cooling

and dry machining in terms of machined surface quality. Revankar et al. [65] also indicated that

the MQL method provided the best surface roughness in comparison to dry and flood machining.

Deiab et al. [66] evaluated the performance of MQL, dry machining, and flood cooling in

determining surface roughness, tool wear, and energy consumption during the machining of Ti-

18

18

6Al-4V titanium alloy. It was found that the vegetable oil based MQL provided the best results

and was considered to be a sustainable alternative coolant compared to the dry condition and flood

coolant.

An experimental study was conducted by Dhar et al. [67] to investigate the effect of cooling

conditions on flank wear and surface roughness during the machining of AISI 4340 steel. It was

found that the MQL condition significantly decreased surface roughness and tool wear, and

reduced the resultant cutting temperature. Kaynak et al. [56] conducted a study to compare the

machining performance of cryogenic to dry and MQL during the machining of Inconel 718. The

study found that cryogenic and MQL techniques improved the machinability of Inconel 718 in

terms of surface quality, tool wear, and cutting temperature. The study also revealed that the MQL

mode significantly lowered the cutting forces when compared to the dry machining and cryogenic

techniques at low cutting speeds. It was concluded that the MQL approach dropped the cutting

temperature by approximately 20% when compared to dry machining.

Chetan et al. [68] designed a mathematical model to predict the specific cutting energy when

cutting a nickel-based alloy using MQL approach. The authors reported that an increase in the

pressure and flow rate of the MQL mist reduced the specific cutting energy in which the MQL

shortened the contact length in the chip-tool interface, as well as minimized the shear flow stress

above the rake face of the cutting tool. Khan et al. [69] investigated the effects of flood coolant,

dry condition, and MQL during finish turning of CP-Ti Grade 2. The study found that MQL

machining provided a lower tool wear rate in comparison with the wet and dry conditions. Also,

the effectiveness of a MQL over the machining process was determined by Boswell et al. [57].

This review study found that improvements were observed from using MQL lubrication to

determine various machining aspects, such as tool wear, surface roughness, cutting force, and

19

19

cutting temperature. It was also concluded that the MQL technique has the ability to reduce the

environmental impacts that were caused by conventional flood cooling. Ginting et al. [70] pointed

out that the MQL lubrication was less costly when compared to the conventional flood because it

required less power and achieved a shorter machining time. Several investigations have revealed

that the MQL technique is a viable alternative to conventional cutting fluids under similar

performance parameters, and is suitable for green machining [71-73].

Although the MQL approach appears to exhibit good lubricating features, it has some problems

such as clogging of chips/debris and an inefficient cooling capacity compared to the standard flood

coolant [74-76]. The base MQL is therefore not an efficient alternative to a standard flood coolant

particularly when cutting ADI and titanium alloy. Therefore, numerous researchers have attempted

to improve the MQL’s cooling capacity. One of these methods is Minimum Quantity Cooling

Lubricate (MQCL), which mixes MQL mist with cooled compressed air. Weinert et al. [64]

demonstrated that MQCL is a safe and efficient cooling and lubrication approach, which could be

an effective alternative to the classical MQL technique. In another study, Raza et al. [53] showed

that the MQL and MQCL techniques improved the tool life and surface roughness when compared

to the base synthetic coolant and dry machining. When reviewing the available literature, Revuru

et al. [77] found that when machining titanium alloy, MQCL had a better cooling capacity, and

provided promising results compared to the classic MQL. Zhang et al. [78] machined Inconel 718,

and reported that the MQCL technique increased the tool life and reduced the cutting forces

compared to dry machining. It was also claimed that the vegetable oil based MQCL technique met

the standards for environmentally friendly machining and could be an effective alternative coolant

to MQL and dry machining. Furthermore, Su et al. [79] conducted comparative experiments to

study the effects of dry machining, MQL, and Chilled Air Minimum Quantity Lubrication

20

20

(CAMQL) on the machining performance of Inconel 718 and the milling of AISI D2 cold work

steel. The study found that the use of CAMQL resulted in reasonable enhancements to the surface

finish and tool life. As well, it was found that there was a considerable improvement in chip

formation during the machining of Inconel 718 when compared to the dry and MQL methods.

Boswell et al. [80] conducted milling experiments for an aluminum alloy using different cooling

methods including dry, flood, cooled air, MQL and CAMQL. The findings indicated that the

lowest cutting force was provided by MQL, followed by CAMQL. It was also concluded that the

MQL and CAMQL techniques were demonstrated a good cutting performance. The study also

found that both the MQL and CAMQL methods provided a machined surface quality equivalent

to that achieved by a standard flood coolant.

Although the above methods showed an improvement in the machining results, and a comparable

surface quality with respect to flood coolant, the cooled air did not effectively enhance the

lubrication performance and thermal conductivity of the MQL technique. It may also have required

expensive equipment for operation and have necessitated strict safety measures. In most cases the

base cutting fluid of the MQL system is water or oil. Instead of applying cooled compressed air to

enhance the cooling performance of an MQL, nanoparticles could be dispersed into a base cutting

oil, creating NanoFluid-MQL (NFMQL). This new coolant has been proposed to improve the

lubrication and heat capacity removal ability of the MQL approach during machining operations.

Nanoparticles are typically metals or metallic oxides. Various types of nanoparticles with excellent

properties have been used to improve the thermal conductivity and lubricity of an MQL coolant,

such as aluminum oxide (Al2O3), carbon nanotubes (CNTs), TiO2, MoS2, C60, CuO, and diamond

[81]. Nanoparticles have remarkable properties, such as high thermal conductivity and convection

coefficients, which could improve the cooling and lubricating properties of an MQL. Furthermore,

21

21

by dispersing nanoparticles into the base fluid, the viscosity and thermal conductivity of an MQL

is enhanced [82-84]. The appropriate distribution of nanoparticles into the base cutting oil has been

beneficial to the machining process as it helps to decrease the induced friction at the tool-chip

interface [85]. In the aforementioned studies, a nanofluid-MQL has been proposed to improve the

cooling and lubricating properties of the base MQL and has shown promising results for

machinability. In this regard, in current research study, a nanofluid-MQL has been proposed as an

effective cooling strategy for reducing the frictional heat as well as environmental and health

hazards in the machining difficult-to-cut materials due to their promising tribological properties,

high thermal conductivity, and anti-bacterial features [86, 87].

Other studies have concentrated on the grinding process. Huang et al. [88] investigated the

influence of adding multi-walled carbon nanotube (MWCNT) nano-additives to an MQL during

the grinding of NAK80 mold steel. The study found that nano-lubrication enhanced the grinding

performance in terms of grinding temperature, surface roughness, and grinding forces compared

to dry machining and pure MQL. Pashmforoush et al. [89] used a water-based copper nanofluid

during the grinding of Inconel 738 super alloy, and found that the surface roughness was improved

by 62.16% and 36.36% when compared to the dry grinding and standard fluid methods

respectively. Molaie et al. [90] performed experiments which investigated the impacts of the

vibration assisted grinding process combined with MQL nanofluids based on MoS2 nanoparticles.

The experiments determined that the use of a nanofluid-MQL significantly reduced the grinding

tangential force. It was also shown that combining vibration assisted with nanofluid-MQL

produced a noticeable improvement in the surface roughness. As well, it enhanced the grinding

performance while minimizing the costs by reducing the amount of lubricants used.

22

22

Setti et al. [91] conducted a grinding process over Ti-6Al-4V using a nanofluid based on an MQL

with Al2O3 and CuO nano-additives. This study indicated that the implementation of nanofluids

forms tribofilms in the machining zone. This reduces the coefficient of friction, increases the

effective cooling ability, and results good chip formation. It was also noted that adding Al2O3 and

CuO nano-additives to water or base oil enhanced their thermal behavior. It has been stated in the

literature that nanofluids make an important contribution towards meeting the heat dissipation task

when machining hard-to-cut materials since nanoparticles provide a high practical thermal

conductivity value when compared to the base fluid [92].

Other studies in the literature have indicated that nanofluids make an important contribution to

milling operations. Yuan et al. [93] carried out end milling experiments on 7050 aluminum alloy

using an MQL nanofluid to explore the influence of three types of nanoparticles: nano-copper

(Cu), nano-silicon carbide (SiC), and nano-diamond (diamond). The results showed that using

canola oil-based diamond nanofluids improved the cutting force by 10.71%, and using natural 77

oil-based diamond nanofluids improved the surface quality by 14.92% compared to the dry

condition. Nguyen et al. [94] evaluated the influence of MQL with nano-platelets (xGnP) on

machinability characteristics during the milling of Ti-6Al-4V alloy. In this study, the nanofluid-

MQL reduced the flank wear and increased the tool life with respect to both pure MQL and dry

machining. Kim et al. [95] demonstrated that adding diamond nanoparticles to an MQL coolant

during the micro end milling of a Ti-6Al-4V alloy could be effective at minimizing the surface

roughness and milling forces.

Sahu et al. [96] conducted turning cutting tests on Ti-6Al-4V titanium alloy using three different

cooling approaches: dry, conventional flood, and nanofluid-MQL. It was found that the nanofluid

23

23

decreased the tool wear by 34%, reduced cutting forces by 28%, and lowered the surface roughness

by 7% compared to a standard flood coolant at the same cutting speed of 150 m/min.

Su et al. [97] evaluated the performance of MQL with graphite-LB2000 and graphite-PriEco6000

nanoparticles during the machining of AISI 1045 medium carbon steel. The results demonstrated

that the use of a graphite oil-based nanofluid-MQL helped to lessen the cutting temperature and

cutting forces. Eltaggaz et al. [98] performed an investigation on the effect of Al2O3 nanoparticles

with MQL in the turning of ADI. The study revealed that the nanofluid-MQL considerably reduced

tool flank wear when compared to pure MQL. This was due to the high thermal conductivity and

unique tribological aspects of the Al2O3 nanofluid. It was also concluded that the use of nanofluid

can be an effective cooling and lubricating approach for machining ADI in a sustainable manner.

Other studies have focused on the drilling process. Huang et al. [81] reported that the use of

nanofluid-MQL with 2 wt. % for the micro-drilling of 7075-T6 aluminum alloy reduced the cutting

force, torque, and burr wear. This study also indicated that a significant improvement in the quality

of the drilled holes was observed due to a considerable reduction in the cutting temperature during

the drilling process. Nam et al. [99] performed micro-drilling on a titanium alloy employing

vegetable oil based MQL with nano-diamond particles (0.4 wt. %). In this research, the nanofluid-

MQL reduced drilling torque and thrust force, particularly at a low feed rate of 10 mm/min. It was

also reported that nanofluid-MQL effectively limited chip adhesion to the cutting tool and

minimized the burr of the drilled holes.

Other studies have focused on specific aspects related to machining and the addition of

nanoparticles. Several have studied the machining of both Ti-6Al-4V [29, 101] and Inconel 718

[102,103] using a multi-walled-carbon nanotube with the MQL. In these studies, better results

were achieved when compared to the base MQL in terms of power consumption, flank wear, and

24

24

machined surface quality. Additionally, another study revealed that nanofluids have effective

capabilities to extract the induced heat [104]. Different attempts [47, 105] have been made to study

and solve the excessive heat generation problem during cutting processes to achieve better

machinability, especially for difficult-to-cut materials. Another attempt [106] have focused on

modeling various machining characteristics, such as residual stresses, machined surface hardness,

and friction behavior.

Most research has agreed that nano-fluids provide higher thermal conductivity when compared to

the base fluid; however, determining the appropriate concentration of the nanoparticles is an

immense challenge. The addition of excessive nanoparticles may result in nanoparticle

agglomeration, which reduces the lubrication properties, thereby negatively influencing the tool

life and surface quality. Lee et al. [86] stated that adding nanoparticles to the base lubricants

improved the tribological properties of the nanofluid; however, the tribological properties of

nanoparticles depend on their physical properties such as shape and size. This study also found

that by adding appropriate amounts of nanoparticles, the machining performance of the nanofluids

was effectively improved. In this regard, several studies have been conducted to investigate the

effect of nanoparticle concentrations on the machining performance of nanofluids. Sharma et al.

[107] reported that an increase in the nanoparticle concentration improved the thermal conductivity

and viscosity of the nanofluid. In another study, Sharma et al. [108] found that an increase in the

particle volume fraction effectively improved the thermal conductivity of the nanofluid. Zhang et

al. [109] performed grinding tests on 45 steel using a jet MQL nanofluid (soybean oil + MoS2)

with different mass fraction concentrations (2, 4, 6, 8, and 10 %) of MoS2 nanoparticles. It was

observed that, when increasing the mass fraction nanoparticle concentration from 2% to 6%, there

was a noticeable reduction in the specific grinding energy, grinding forces, and friction coefficient.

25

25

When nanoparticle concentrations were increased above 6%, the grinding forces, the specific

grinding energy, and the friction coefficient also increased. This increase can be attributed to an

agglomeration of nanoparticles. The nanoparticles were then unable to lubricate the grinding zone,

thus negatively affecting the surface finish. It was concluded that a 6% mass fraction concentration

of nanoparticles provided the lowest specific grinding energy, tangential grinding forces, and

friction coefficient. In fact, adding the appropriate amount of nanoparticles into the base oil

promises to be advantageous for the cutting process as nanoparticles help to decrease the

coefficient of friction in the cutting zone [110].

Vasu et al. [111] conducted turning machining on an Inconel 600 alloy using Al2O3 nanoparticle

based MQL and regular MQL vegetable oil. It was found that the MQL with a 6% nanoparticle

concentration considerably improved the tool wear, surface roughness, cutting temperature, and

cutting forces when compared to both MQL with a 4% volume fraction and regular MQL vegetable

oil. In addition, Wang et al. [112] conducted grinding experiments using nanofluid-MQL with

different concentrations of Al2O3 nanoparticles (0, 0.5, 1, 1.5, 2, 2.5, 3, 3.5, and 4 vol. %). It was

revealed that the nanofluid concentration had a significant effect on the grinding performance. It

was also found that a nanofluid concentration of 2.0 vol. % provided the best tribological

performance, and that a further increase in the nanoparticle concentration did not lower the tool

wear and friction.

Many researchers have commented on the limited number of investigations regarding the

effectiveness of cutting fluids for difficult-to-cut materials. The effectiveness of the cutting fluid

is determined by how well it decreases the cutting forces, tool wear, and surface roughness. The

larger challenge is determining the optimal sustainable coolant strategy for machining difficult-to-

cut materials at an economical cost [113]. This work proposes to develop a better understanding

26

26

of the influence of cutting parameters and cooling strategies on the machinability of both a titanium

Ti-6Al-4V alloy and ADI Grade 2 in order to reduce the energy consumption, minimize the

production costs, and improve the machinability. As well, the current work intends to determine

the optimum sustainable cooling approach to machining these metals while considering the

operator (s) health and environmental impact. In addition, this study proposes to fill the gap in the

available literature regarding the optimal cutting conditions for machining these materials.

27

27

Table 2.1: Summary of available research on MQL and MQCL coolants.

Reference(s) Lubrication Cutting

fluid

Salient

features

Conclusion

Ramana et al.[33], and Yusuf Kaynak [56]Boswell et al. [57], Revankar et al. [65], Deiab et al. [66], Dhar et al.[67] , and Khan et al. [69],

MQL

Vegetable oil

Studied MQL at different cutting parameters and cooling strategies

MQL found to be more effective than dry machining, flood, and cryogenic. MQL showed better results in tool wear, surface roughness, power consumption compared to dry and flood

Chetan et al. [68]

MQL

Vegetable oil (sunflower oil)

Designed mathematical model for predicting the specific cutting energy during machining process under MQL

The increase in pressure and flow rate of MQL reduced the friction at chip-tool interface hence helps to reduce the specific cutting energy. High feed rate increases the sliding speed over the tool rake face and there by reduces the coefficient friction.

Ginting et al. [70]

MQL

Vegetable oil

Compared the sustainability and machining cost between the MQL and conventional flood

MQL less cost in which save cost of a huge amount of cutting fluid, carbon cost, and save the disposal process. MQL can be replaced the flood cooling because it’s more sustainable strategy.

Raza et al. [64] Revuru et al. [77]

MQL & MQCL

Vegetable oil & Vegetable oil +

cold air

Investigated the effect of 6 cooling strategies on tool wear

MQL and MQCL improved the tool life and machined surface quality compared to other cooling strategies

Zhang et al. [78]

MQCL

Biodegradable vegetable oil +

cold air

Study the relationship between tool wear propagation and cutting forces during milling Inconel 718

The use of MQCL increased the tool life and reduced the cutting forces. MQCL method can be an effective alternative coolant to pure MQL method.

Boswell et al. [79]

MQCL

Vegetable oil +

cold air

Study the effect of different cooling methods on machining results

MQL and MQCL are provided reasonable surface quality and better cutting performance compare to dry, flood, and cooled air.

Su et al. [80]

MQL & CAMQL

Cooled air -20 oC + UNILUB

2032

Compared the machining performance of CAMQL to cooled air, pure MQL, and dry machining

The CAMQL approach showed a better tool life and chip formation results during machining Inconel 718 compared to dry condition and classical MQL. CAMQL can be an environment friendliness cooling approach.

28

28

Table 2.2: (a) Summary of available research on nanofluid coolants.

Reference(s) Machining

process

Nano-

additives

Salient

features

Conclusion

Huang et al. [88]

Grinding

Multi-walled carbon

nanotube (MWCNTs)

Investigated the influence of adding MWCNTs on grinding outputs

Nanofluid showed improvement in tool life, grinding temperature, and grinding forces compared to pure MQL and dry machining.

Pashmforoush et al. [89]

Grinding

Water-based copper

Investigated the influence of nanofluid, standard fluid and dry condition on surface roughness of grinding Inconel 718

Water-based copper nanofluid improved the surface roughness compared to dry (62.16%) and flood (36.36%)

Molai et al. [90]

Grinding

MoS2

Investigated the impact of nanofluid on grinding tangential force

The use of nanofluid enhanced the machining performance as well as provided opportunity to reduce the cost

Setti et al. [92]

Grinding

Al2O3 and

CuO

Study the influence of using nanofluid on machinability of Ti-6Al-4V

Nanofluid formed tribofilms in machining zone in which minimized the friction and enhanced the cooling ability of MQL coolant.

Yuan et al. [93]

Milling

Nano-copper

(Cu) Nano-silicon carbide (SiC)

Nano-diamond

Exam the effect of three different type of nano-additives

Copper and diamond nanofluid showed the best performance in reduction of cutting forces and surface roughness while silicon nanofluid illustrated effective surface quality. It was concluded that the nano-additive with big size negatively affected machined surface quality.

Nguyen et al. [94]

Milling

Nano-platelet (xGnP)

Characterized the machinability of milling Ti-6Al-4V alloy under different cooling methods

The nanofluid MQL method reduced the flank wear rate resulting increase in cutting tool life compared to dry machining and classical MQL approach.

Kim et al. [95]

Micro-Milling

Diamond nanoparticles

Evaluated the performance of nanofluid MQL during micro end milling process

Adding diamond nanoparticles to MQL enhance the milling performance of MQL coolant in which reduced the milling forces and improved the machine surface quality.

29

29

Table 2.2: (b) Summary of available research on nanofluid coolants.

Reference(s) Machining

process

Nano-

additives

Salient

features

Conclusion

Sahu et al. [96] Hegab et al. [100]

Turning

Multi-walled carbon

nanotube (MWCNTs)

Compared the performance of nanofluid-MQL using MWCNTs to dry and flood

Nanofluid-MQL provided lower value of tool wear, cutting forces, the power consumption, and surface roughness compared to other cooling method under the same cutting condition.

Yu Su et al. [97]

Turning

Graphite-LB2000 and

Graphite-PriEco6000

Evaluated the performance of graphite nano-additives

The Nanofluid graphite contributed to reduce both cutting force and cutting temperature.

Eltaggaz et al. [98]

Turning

Aluminum oxide

(Al2O3)

Study ADI machinability under using pure MQL and nanofluid MQL

Adding Al2O3 enhanced the thermal conductivity of nanofluid compared to pure MQL due to high thermal conductivity and good tribological characteristics of nano-additives.

Huang et al. [81] (2 wt. %) Nam et al. [100] (4 wt. %)

Micro-drilling

nano-diamond

Effect of naofluid-MQL method in mico-drilling

Nanofluid-MQL decreased the cutting force, torque, and burr wear. Also, the use of nanofluid significantly reduced the cutting temperature and improved the quality of drilled holes.

Hegab et al.[101] [102]

Turning

Multi-walled carbon

nanotube (MWCNTs)

Comparison pure MQL and naofluid-MQL for machining Inconel 718

Nanofluid-MQL showed better improvement in tool life, power consumption, and machined surface quality

Lee et al. [86]

Grinding

Nano-diamond

With 0.2 wt % and 0.8

wt %

Discussed the tribological properties of nanoparticles

They reported that nanoparticles enhance the tribological properties of the base lubricant, thereby improve its grinding performance.

Sharma et al. [108]

Turning

Al2O3-MoS2 hybrid

nanoparticle

Investigated the effect of increase the nanoparticle concentrations

Increase the nanopaticale concentration increases the thermal conductivity and viscosity of the nanofluid.

Zhang et al. [109]

Grinding

MoS2 nanoparticle

s

Investigated different nanoparticle mass fraction concentrations

Increasing mass fraction from 2% to 6% improved the grinding machining; however, increased nanoparticle concentration above 6% negatively affected the machinability due to agglomeration phenomena.

Vasu et al. [11]

Turning

Aluminum oxide

(Al2O3)

Two volume fraction concentrations 4% and 6%

It was found that using 6% volume fraction showed better improvements in tool wear, surface roughness, cutting temperature, and in cutting forces reduction compare to 4%.

30

30

Chapter 3: Research Methodology

3. 1 Introduction

The experimental approaches and pieces of equipment that were used in the data acquisition of the

experimental study are presented in this chapter. The experimental tests were conducted to study

the cutting performance of ADI and Ti-6Al-4V titanium alloy utilizing a sustainable nanofluid-

MQL vegetable oil cooling strategy.

This chapter presents the details of the experimental techniques and methods used during the

current study. This chapter also provides information about the design of the experiments, the

cutting tool and workpiece materials used, and the experimental setups and the devices employed

to record and determine the cutting parameters and output results.

3.2 Design of Experiments (DOE)

Any material machining process is controlled by a number of factors [114]. These operation

parameters have a governing effect on the machining process. Thus, an investigation of the how

to control these parameters is important to make the cutting process more efficient. A systematic

method known as design of experiments (DOE) is used to investigate the effect of the governing

parameters on the output of the machining operations [114,115]. Using DOE, test matrices were

developed to evaluate cutting process outputs, such as machined surface roughness (Ra) and

cutting tool wear. DOE is also used to determine power/energy consumption when machining

difficult-to-cut materials, e.g., austempered ductile iron (ADI) and titanium alloys (Ti-6Al-4V). It

is also an approach used to explore and understand the cause and effects of the relationships in a

certain operation. The design of experiments technique consists of the following features:

31

31

Planning stage: This phase deals selecting activities related to the problem statement, the research

objectives and aims, the machining of the test matrix with different levels of variables, and the

different experimental setups.

Procedure stage: This stage involves conducting the experiments and collecting the output

parameters in order to evaluate effectiveness of the cutting operation.

Results and analysis: The findings and conclusions from the procedure stage are then plotted by

analyzing the data from the machining experiments. In order to validate and confirm the

conclusion, further machining tests are often required.

3.3 Planning of Experiments

Machining tests were carried out using different cooling strategies, on a CNC HAAS machine.

These strategies included the dry machining, the conventional flood, a vegetable oil- based MQL,

and a vegetable oil-based nanofluid MQL. In the current study, the steps of the DOE were as

follows:

1. Problem statement: this study intends to study the machinability of two difficult-to-cut

materials (ADI and Ti-6Al-4V alloy), to investigate the effect of a sustainable cooling

strategy (vegetable oil-based nanofluid-MQL), to improve the FEM numerical model to

predict the generated heat on the machining tool, and to investigate the ability of the CFD

simulation to predict the heat removal ability of the MQL technique.

2. Aims and objectives: the purpose of this study to investigate the potential of using a

sustainable coolant (a nanofluid-MQL) as an alternative to the conventional flood method

when machining ADI and Ti-6Al-4V alloy. Another goal is to develop a model that would

utilize both finite element model (FEM) to simulate cutting heat distribution on the cutting

32

32

tool and computational fluid dynamics (CFD) to simulate the influence of the MQL coolant

on the machining environment.

3. Identification of the effect of processing parameters: this study considers input cutting

variables, included cutting speed, feed rate, depth of cut, and cooling strategies for

machining, as well as cutting outputs included surface roughness, cutting forces, wear

mechanisms, tool wear, power consumption, and cutting temperature. The nanoparticle

concentration is an additional parameter for the nanofluid-MQL cooling approach.

4. Selecting of machining conditions: this study intends to choose both suitable and

recommended cutting variables in experiment matrix.

5. Improvement of numerical models: this research intends to improve the FEM and CFD

models to predict the effecting parameters such as cutting energy, cutting forces, chip

formation, and cutting tool tip temperature.

6. Implementation of experiments: experimental setups are used to collect sets of experimental

data.

7. Analysis of results: an analysis of the collected data is performed utilizing the proper

analysis methods and interpreting the findings. Also, some experimental data are utilized to

validate the numerical models.

3.3.1 Cutting Tool and Workpiece Materials

Machinability refers to the ease of machining a material. The term is relative, and various indices

such as machined surface quality, tool life, and power consumption can be used to rate it. Because

of the number of possible cutting process variables and their ranges, evaluating machinability often

entails a large number of experiments that are time consuming and expensive. When machining

any material, its properties should be considered. In this research, the workpiece materials used in

33

33

the turning trials were two difficult-to-cut materials: ADI grade 2 and Ti-6Al-4V titanium alloy.

Stock of these materials was available in cylindrical form.

3.3.1.1 Austempered Ductile Iron (ADI)

Tables 3.1 and 3.2 illustrate the chemical composition (wt. %) of ADI grade 2 and its mechanical

properties respectively. Mechanical properties of ADI can be attributed to the high carbon

austenite in its ausferritic matrix (ferrite and carbon-enriched austenite) formed after the

austempering treatment process for casting ADI [11]. Additionally, some researchers have

reported that alloying elements added during casting, austenitizing, and austempering temperatures

contribute to the variations in ADI’s mechanical properties [116-119].

Table 3.1: Percent weight chemical compositions of ADI grade 2 (wt %). [120]

Element Wt. % Element Wt. % Element Wt. %

C 2.8 Mn 0.4 Mo -

Si 2.4 Mg 0.03

Cu 0.6 Cr 0.02

Table 3.2: ASTM A897/897M-02 mechanical properties of ADI grade 2. [121]

Property Value Property Value

Tensile Strength 1050 (MPa) Impact Energy 80 (J)

Yield Strength 700 (MPa) Typical Hardness 340 (BHN)

Young’s Modulus 157.9 (GPa) Elongation (%) 7

34

34

3.3.1.2 Ti-6Al-4V Titanium Alloy

Ti-6Al-4V titanium alloy was used as a workpiece material in this research. The chemical

composition and mechanical properties of this difficult-to-cut material are shown in Tables 3.3

and 3.4 respectively.

Table 3.3: Chemical compositions of Ti-6Al-4V (wt %). [29]

Element Wt. % Element Wt. % Element Wt. %

Ti Balance C 0.1 N 0.05

Al 5.5-6.7 Fe 0.3 H 0.0125

V 3.5-4.5 O 0.2

Table 3.4: Ti-6Al-4V titanium alloy mechanical properties (at room temperature). [115]

Property Value Property Value

Tensile Strength 993 (MPa) Poisson´s ratio 0.34

Yield Strength 830 (MPa) Hardness (HRC) 36

Young’s Modulus 114 (GPa) Elongation (%) 14

3.3.1.3 Turning Cutting Tool Inserts

The machining tests for turning ADI grade 2 and Ti-6Al-4V alloy were performed using coated

and uncoated carbide inserts. The turning tool inserts were chosen based on Sandvik Coromant´s

recommendations [122]. The specifications of the turning cutting inserts are described in the

following sections.

35

35

A) - Coated Carbide CNMG120416MR (ISO)

This cutting insert has a substrate consisting of an HC with good resistance for plastic

deformation, and its nose radius is 1.588 mm (see Figure 3.1). It has a thin physical vapour

deposition (PVD) consisting of a (Ti,Al)N+(Al,Cr)2O3 coating with a good adhesion, toughness,

and wear resistance. This insert is an excellent heat resistant super alloy. It also has a roughing

geometry with a stronger edge line for demanding applications [122].

Figure 3.1 CNMG120416MR (ISO) cutting tool insert. [122]

B) - Uncoated Carbide CNMT 12 04 08-KM H13A (ISO)

This uncoated carbide cutting insert has a nose radius of 0.8mm. It was developed by Sandvik

Coromant (see Figure 3.2). This insert is appropriate for medium to rough turning of alloy cast

irons and malleable irons. It has both a good thermal and mechanical resistance, and offers

outstanding edge security [122].

Figure 3.2 CNMT 12 04 08-KM H13A (ISO) cutting tool insert. [122]

36

36

3.4 Experimental Setups

All machining tests were performed on the 15HP ST-10 HAAS CNC-lathe machine shown in

Figure 3.3. Each experiment was repeated twice to minimize experimental errors. A Kistler multi-

channel piezoelectric dynamometer connected to the data acquisition system was employed to

measure the cutting forces during the cutting processes. The energy consumption was recorded

during the machining process using a Power-Sight device. As well, a Mitutoyo roughness tester

SJ 201P was used to monitor and determine the machined surface roughness produced during the

machining tests. In order to measure the wear on the flank face of the cutting tool, a Mitutoyo tool

maker was employed. A scanning electron microscope (SEM) was then used to investigate and

analyze the major wear mechanisms. A high-resolution Keyence VH-Z250R microscope was used

to analyze the comparative chip formation. The MQL system, in the form of an air–oil mixture,

was supplied by the stand-alone booster system (Eco-Lubric) as an external cooling delivery

system. It had a base fluid flow rate of 40 ml/h and an air pressure of 0.5 MPa. The vegetable oil

ECOLUBRIC E200 was used as a base cutting fluid. The type of oil used (rapeseed) is

environmentally harmless, suitable for industrial applications, and has a biodegradability of 90%

in 28 days [66]. The nozzle was directed at the cutting tool edge at a 45o angle.

To prepare the nano-cutting fluid, aluminum oxide gamma nanoparticles (Al2O3) were used as

nano-additives. Al2O3 nanoparticles were selected due to their superior tribological properties,

their high thermal conductivity, and antitoxic aspects. The nanoparticles (Al2O3), which were 18

nm in diameter, had 93% purity, and a 138 m2/g specific surface area, were employed for nano-

cutting fluid preparation. An ultrasonic device (AQUASONIC-50HT) was used to disperse the

nanoparticles into the cutting fluid over a period of 3 hours at 60◦C. For stirring step, a magnetic

37

37

stirrer was used for 30 minutes to ensure full dispersion of the nano-additives into the base cutting

fluid.

The use of a physical or chemical treatment, such as surfactants, is recommended in order to obtain

the powerful forces necessary to ensure sufficient dispersion of the clustered nano-additives, and

thus to achieve the desired enhancements to the thermal conductivity and viscosity of the resultant

nanofluid [124-126]. The main reason for using the surfactant is to improve the stability of the

resultant nanofluid and make it more hydrophilic. The surfactant also increases the surface charges

of the nano-additives; therefore, the repulsive forces between the nanoparticles are increased.

Sodium dodecyl sulfate (a surfactant) of 0.2 mg was therefore used to prepare the nanofluid.

Figure 3.3 CNC-lathe machine 15HP ST-10 HAAS.

In this research, different experimental setups were used to investigate machining characteristics

during the machining of ADI and Ti-6Al-4V titanium alloy.

38

38

Figure 3.4 Experimental setup.

Figure 3.5 Schematic experimental setup for machining ADI under different cooling strategies.

39

39

Table 3.5 shows the specifications of the ECULUBRIC E200L vegetable oil used as a base fluid

in the MQL system in this study. This vegetable oil is mostly based on rapeseed oil, and is produced

by ACCU-Svenska AB, Sweden.

Table 3.5: Specifications of the base vegetable oil utilized in the MQL system. [127]

Property Description

Chemical description Pure vegetable-based lubricant without any chemical modifications.

Health risk Environmentally friendly; no risk to operator-(s) health

Flash point 325 Co

Ignition point 365 Co

Density 0.92 g/cm3 (at 20 Co)

Viscosity 70cP (at 20 Co)

Partition coefficient < 3%

3.5 Measurement of Machining Variables

This section discusses the equipment that was used to obtain detailed measurements of the

machining variables.

3.5.1 Cutting Forces

During the machining experiments for ADI and Ti-6Al-4V alloy, the cutting forces were recorded.

A Piezoelectric Kistler dynamometer (Model: 9257b) [128] was used to measure the cutting forces

in the X (Fx), Y (Fy), and Z (Fz) directions. Also, a Kistler 5070 charge amplifier with four

channels was used to obtain the input signals from the force measurement. The Dynoware software

package on the computer was used to calculate and present the output signals from the charge

amplifier. The specifications of the Kistler dynamometer are shown in Table 3.6.

40

40

Figure 3.6 Cutting forces data evaluation system: (a) Kistler 9257b dynamometer and (b) Kistler

5070charge amplifier.

Table 3.6: Specifications of Kistler dynamometer. [128]

Model 9257b (Dynamometer) & 5070 (Charge amplifier)

Calibration Range (KN)

Fx direction 0-7 (KN)

Fy direction 0-7 (KN)

Fz direction 0-7 (KN)

Sensitivity (pc/N)

Fx direction -7.750

Fy direction -7.750

Fz direction -7.750

3.5.2 Tool Wear and Wear Mechanism

The flank wear appears on the cutting tool due to the friction between the tool flank face and the

machined surface. In the current research, a Mitutoyo tool maker microscope was utilized to

determine and investigate the amount of tool flank wear as seen in Figure 3.7(a). In addition, Figure

41

41

3.7(b) presents the standard measurement of the tool flank wear (VBmax) throughout the machining

tests. Information regarding the Mitutoyo tool maker microscope is presented in Table 3.7.

Figure 3.7 (a) Tool maker microscope brand Mitutoyo (b) Standard measurement of the max

flank wear (VB max) on the tool flank face [129].

Table 3.7: Mitutoyo tool maker microscope information. [130]

Model Mitutoyo TM 510

Effective area 150x92 mm

X Y range 100x50 mm

Max. height of workpiece 107 mm

Max weight of workpiece 5 kg

Max. magnification 30X

Lighting Green filter, adjustable light intensity

Transmitted Source of light: Tungsten bulb (2W, 24V)

Reflection of lighting Adjustable light intensity

42

42

After each machining test, the cutting insert edges were marked with a dot a few millimeters from

the cutting edge with a different coloured paint, and each insert was put in a small plastic bag.

Each plastic bag was labeled with the cutting parameters such as feed rate, cutting speed, depth of

cut, cooling strategy, test number, workpiece material, and cutting material. To investigate the

wear mechanism, a scanning electron microscope (SEM) was used (Figure 3.8). The SEM images

were analyzed in order to gain a deeper understanding of tool wear mechanisms. The analysis of

the SEM images helped to predict tool failure and to determine the type of tool wear. Figure 3.9

illustrates the SEM images of flank wear (a) and crater wear(b) on the cutting tool insert used to

machine ADI at a speed of 120 m/min and a feed rate of 0.3 mm/rev under dry machining (left)

and flood coolant (right).

Figure 3.8 Scanning Electron Microscope (SEM).

43

43

Figure 3.9 SEM images for a coated carbide tool (CNMG120416MR) after machining ADI at

V=120 m/min and f=0.3 mm/rev under dry machining.

3.5.3 Machined Surface Roughness (Ra)

In this work, a Mitutoyo surface roughness tester SJ 201P was used to measure the surface

roughness of the machined surface. For all of the machining trials, surface roughness readings

were determined three times and the average values were recorded in order to decrease the

experimental errors. Figure 3.10 illustrates the Mitutoyo surface roughness tester SJ 201P that was

used for the machining experiments on the ADI and Ti-6Al-4V alloy. The general specifications

of this instrument are shown in Table 3.8.

44

44

Figure 3.10 Mitutoyo surface roughness tester SJ 201P.

Table 3.8: Specifications of Mitutoyo surface roughness tester SJ 201P. [130]

Model Mitutoyo SJ 201P

Drive speed 0.25 mm/sec, 0.5 mm/sec, and 0.75 mm/sec

Detecting method Differential inductance

Length of evaluation 12.5 mm

Force measuring 4 mN

Stylus tip radius 5 µm

Length sampling 0.25 mm, 0.8 mm, and 2.5 mm

Roughness standard JIS, DIN, ISO, ANSI

3.5.4 Power Consumption

To determine the power (energy) consumption during each machining test, a Power Sight logger

PS 2500 was selected to collect the power consumption data to be processed by the computer

(Figure 3.11) [131]. The power was computed through the three voltage and current input jacks on

the power logger. The value of the current was wired directly to the CNC lathe power line and the

45

45

voltage value was wired into the main power line jack for an input voltage value of 195-210 Volts.

The specifications of this instrument are presented in Table 3.9.

Figure 3.11 Power sight.

Table 3.9: PS 2500 power data logger specifications. [115,131]

Model Power logger PS 2500

Range of Operation

Between 0 and 50 oC and 70% humidity

Rate of Measurement For each current and voltage analyses 2

cycles/sec (utilizes 130 samples/cycle at

60 Hz). The total measurements were

updated once/ sec.

Measurement of Frequency DC range: 45-66 Hz, 360-440 Hz

Accuracy: ± 0.5 %

3.6 Numerical Methodology

This section provides information about the numerical techniques used for predicting the cutting

temperature in the tool during the machining of ADI grade 2 and for evaluating the heat removal

capability of the MQL cooling during this research. Details and information about the finite

element model (FEM) and the computational fluid dynamic (CFD) are also provided in this

section.

46

46

3.6.1 Design of Experiment

The experiments were conducted using ADI grade 2 samples under dry condition. The properties

of the samples were obtained as per the ASTM A897 standard specifications (Tables 3.1 and 3.2).

The cutting tool was made of tungsten carbide without coating and was classified as H13A grade

materials, which are known for their strong wear resistance and toughness. The thermo-mechanical

properties of the cutting tool insert are shown in Table 3.10. The aim of the experimental analysis

was to tabulate the recorded temperature generated at 2 mm from the tool tip; the cutting forces,

and the power consumption under dry conditions using different machining parameters. The

machining parameters that varied were the cutting speed and the feed rate. The depth of cut for all

tests was held constant at 1mm, and each tool insert was used for one cutting pass. This was done

to reduce the effects of rubbing friction arising from a worn tool nose, which can cause temperature

readings that do not align with the modeled temperatures. Each run was conducted three times to

evaluate the replicability of the results. The measured dry cutting temperatures were used to

validate the friction model and verify the accuracy of the finite element analysis. Figure 3.12

presents the experimental schematic.

Table 3.10 Cutting tool (CCMT 12 04 04-KM) thermo-mechanical properties. [123]

Property Value

Modulus of elasticity 670 GPa

Poisson’s ratio 0.25

Thermal conductivity 46 W/m.K

Specific heat 203 J/Kg.K

47

47

Figure 3.12 Experimental schematic.

3.6.2 Experimental Setup

The experimental setup included several data acquisition components to collect the machining

variables.

Figure 3.13 Experimental setup (a), and temperature measuring arrangement (b), (c), and (d).

48

48

3.6.3 Thermocouple and Temperature Arrangement

The cutting temperature was captured using a metal K-type thermocouple TJ 36 embedded in the

tool holder and connected to an Omega HH374 data logger, as shown in Figure 3.13 (b). The Omega

HH374 data logger [132] was employed to record the cutting temperature at the nodal junction.

Table 3.11 presents the specifications of the Omega 4 channel data logger. Figure 3.13 (d) shows a

through hole approximately 1.8 mm in diameter was drilled in the tool holder in order to embed the

thermocouple. In order for the measurement to be closer to the cutting tool tip, a hole approximately

1.7 mm in diameter was also made in the uncoated carbide insert up to a depth of 2 mm, as shown

in Figure 3.13 (c). A metal K-type thermocouple 1.6 mm in diameter was fitted through the tool

holder and seat, and then embedded into the insert, as shown in Figure 3.13 (b-d). As seen in Figure

3.13 (b), the temperature measurement was arranged to measure the nodal temperature of a point

situated approximately 2 mm away from the insert tip. To validate the FEM model, the nodal

temperatures determined at a point 2mm distant from the tool tip during the machining of ADI

grade 2 under dry conditions were considered. Figure 3.14 illustrates the method of data acquisition

for the primary variables. In addition, a Keyence VH-Z250R microscope was used to analyze the

chip formation, as shown in Figure 3.15.

49

49

Figure 3.14 Experimental data acquisition flow diagram.

Table 3.11 Specifications of Omega HH 374 data logger. [132 ]

Property Value

Dimensions 197 mm (L) – 65 mm (W) – 36 mm (H)

Weight Approximately 285 g

Measurement range -200 to 1372 oC (-328 to 2501 oF)

Resolution 0.1 oC/ oF < 1000 -, 1 oC/ oF – 1000 -

Accuracy ± (0.1 % of reading + 0.7 oC) or ± (0.1 % of reading + 1.4 oF)

Temperature coefficient 0.01 % of reading + 0.05 oC per oC (< 18 oC or > 28 oC)

Sample rate 1 times per second

Operation temperature -10 to 50 oC (14 to 122 oF)

Operation humidity 10 to 90% RH

Standard accessories Instruction manual, 9V battery, Windows software, carrying case, USB cable, K type bead probe x 2 pcs

Optional accessories AC adapter

50

50

Figure 3.15 Keyence VH-Z250R microscope.

51

51

Figure 3.16 Thesis methodologies.

52

52

Chapter 4: Results and Discussion

This chapter provides the results from research work that has been published on machining of ADI

grade2 and Ti-6Al-4V titanium alloy. Most of the findings are from appended published papers.

4.1 The Effects of Using Different Coolant strategies on Tool Life and Surface

Roughness when Machining ADI

4.1.1 Introduction

Austempered ductile iron alloys (ADI) are an interesting class of materials because of their unique

microstructure and mechanical properties. Accordingly, there is a growing interest in using ADI

in several applications due to its mechanical properties. However, as with difficult-to-cut materials,

the machinability rating of ADI is low and there is still a need to understand the impact of the

cutting process parameters. Machinability of a material depends not only on its properties and

microstructure, but also on the proper selection and control of process variables. The current

section is focused on performing a machinability analysis of ADI in order to understand the effect

of the cutting process parameters on the machined surface quality and tool life. It also studies the

effect of different coolant strategies. Thus, the motivation of this study is to develop a better

understanding of the influence of cutting parameters and cooling strategy to be able to machine

ADI directly with acceptable tool life and cycle time. The design of experiments technique and

response surface methodology is employed to analyze and model the measured responses. In

addition, the cutting tool wear mechanisms are identified and discussed.

Austempered ductile iron (ADI) grade 2 was used as the workpiece and coated carbide tool

(CNMG120416MR) with 1.6 mm nose radius was used for the cutting tests. Four different coolant

strategies were utilized in this study, namely, dry machining, flood cooling, MQL with vegetable

53

53

oil, and MQL with nanofluid. Al2O3 nanoparticles were dispersed into MQL 4 wt. % as suggested

in the literature [98, 133-136].

Turning experiments were conducted at a depth of cut (DOC) of 0.5 mm and length of cut of 50

mm, while different levels of cutting speeds and feeds were utilized as shown in Table 4.1. The 24

cutting experiments were conducted using L24 mixed orthogonal array based on the design of

experiments technique. Each run was performed twice to evaluate the results replicability. The

cooling strategies were evaluated at different levels of high and low cutting speed and feed, which

chosen based on the ranges of recommended cutting parameters by the insert manufacturer. Figure

4.1 shows the experimental setup schematic.

Table 4.1: Assignment of levels to cutting process variables.

Symbol Level 1 Level 2 Level 3 Level 4

Cutting process

variables

Feed rate (mm/rev) A 0.2 0.3 - -

Cutting speed (m/min) B 120 180 240

Lubrication strategy C Dry Flood MQL Nanofluid

Figure 4.1: Experimental setup schematic.

54

54

4.1.2 Results and discussions

The findings of this study were discussed in this section. A maximum flank tool wear of (VB =

0.5 mm) was used as the tool life criteria. The tool life and average surface roughness results are

shown in Table 4.2. The test matrix was developed with the aim of relating the influence of input

parameters to all measured responses.

4.1.2.1 Tool Life and Surface Roughness

The best performance of the cutting tool in terms of tool life was recorded at cutting speed of 120

m/min, feed rate of 0.2 mm/ rev using MQL nanofluid technique as shown Figure 4.2. The 4%

nanofluid coolant showed the highest tool life in comparison to the other employed techniques.

The nanofluid-MQL shows improvement percentages in tool life of 19.4, 4.1, and 30.5% in

comparison with tests performed at same cutting conditions using MQL, flood, and dry techniques,

respectively. Flood coolant shows better results only at cutting speed of 120 m/min and feed rate

of 0.3 mm/rev; however, it is still not an environmentally friendly technique while nanofluid-MQL

is a sustainable technique as it uses a small amount of cutting fluid (environmentally friendly

system).

Several studies have shown that the use of MQL results in a better surface than dry machining [10,

137, 138] which is relatively consistent with that shown in Figure 4.3. The MQL also reduces the

cutting temperature, cutting forces, tool wear, and friction coefficient in comparison with dry

machining. However, the average surface roughness results suggest that the nanofluid appears to

be the best choice when operating at a low feed rate; however, the flood coolant appears to provide

relatively better results at higher feed rate values.

55

55

Table 4.2: L24OA experimental result.

Run

#

Lubrication

technique

Feed rate

(mm/rev)

Cutting speed

(m/min)

Tool life

(min)

Average surface

roughness (μm)

1 Dry 0.2 120 2.72 0.952

2 Dry 0.2 180 1.98 0.895

3 Dry 0.2 240 1.2 0.857

4 Dry 0.3 120 2.29 1.614

5 Dry 0.3 180 1.48 1.634

6 Dry 0.3 240 0.88 1.769

7 Flood 0.2 120 3.41 1.257

8 Flood 0.2 180 2.31 1.016

9 Flood 0.2 240 2.19 1.761

10 Flood 0.3 120 3.45 1.468

11 Flood 0.3 180 2.3 1.126

12 Flood 0.3 240 1.04 1.053

13 MQL 0.2 120 3.01 0.695

14 MQL 0.2 180 2.27 1.005

15 MQL 0.2 240 1.7 0.735

16 MQL 0.3 120 2.27 1.775

17 MQL 0.3 180 1.86 1.564

18 MQL 0.3 240 1.4 1.69

19 Nanofluid 0.2 120 3.55 0.85

20 Nanofluid 0.2 180 2.68 0.88

21 Nanofluid 0.2 240 2.3 0.625

22 Nanofluid 0.3 120 2.85 1.467

23 Nanofluid 0.3 180 2.42 1.98

24 Nanofluid 0.3 240 1.2 1.55

56

56

Figure 4.2: Cutting conditions effects on tool life.

Figure 4.3: Cutting condition effects on machined surface quality.

57

57

Additionally, from Figures 4.2 and 4.3, it can be seen that the shorter tool life and the coarser

surface finish were obtained when machining ADI at the highest levels of cutting speed and feed

(240 m/min and 0.3 mm/rev) for all employed cooling strategies. Thus, compared to the tests

performed using lowest levels of cutting speed and feed rate, the tool life values are decreased by

65, 64, 52, and 63% for dry, flood, MQL, and nanofluid, respectively. Consequently, the surface

roughness values are increased by 46, 58, and 51% for dry machining, MQL, and nanofluid,

respectively under using the highest levels for both cutting speed and feed rate.

Generally, the appropriate chip breaking features and potential low machining temperatures make

cast iron materials proper for machining processes, which are conducted under dry condition.

However, Klocke et al. [139] pointed out that in many cases the use of cooling strategy was

significant in machining ADI and was associated with some substantial benefits which agree with

the results shown in Figure 4. 2. The flood coolant has an advantage in terms of chip clearing as

the high liquid pressure at the cutting zone helps to move the chips away of the cutting tool, thereby

avoiding the recut chips phenomenon and eliminating a crater wear which could damage the

machined surface and dull the cutting tool faster.

Many researchers [140-142] have reported that adding nanoparticles into a base fluid offers a

reasonable reduction in friction and greatly enhances its wettability. Different mechanisms were

discovered to clarify the enhancement in lubricity of nanofluids, including effect of ball bearing,

tribo-film formation, mending, and polishing effects [143]. In this work, the nanofluid

enhancements can be attributed to the thermal conductivity improvements of the MQL due to the

proper dispersion of Al2O3 nanoparticles into the base cutting fluid which enhances the resultant

thermal conductivity value in comparison with the base lubricants because of the remarkable

58

58

properties of the added Al2O3 nanoparticles (e.g., wear resistance, and excellent thermal

conductivity).

The resultant nanofluid was atomized in the MQL system through the applied compressed air,

forming a fine mist surrounded with a base cutting fluid film [144]. This mist has the ability to

penetrate into the chip-tool interfere zone and forming a tribo-film which helps in absorbing the

generated heat and reducing the induced friction. The mentioned characteristics enormously

improve the cooling and lubrication properties and retain the cutting tool hardness for a longer

time as have been confirmed by Khandekar et al. [143]. In addition to the previous point, using

MQL-nanofluid reduced the friction more effectively since the dispersed nanoparticles serve as

spacers, eliminating the contact between tool and workpiece [145]. Thus, the MQL-nanofluid

offers promising results in terms of tool life improvement (ranging between 15 and 21%) in

comparison with tests done using classical MQL. Furthermore, an important point to be considered

is that in most machining levels, when 4% gamma-Al2O3 nanoparticles were dispersed into

vegetable oil, the surface roughness values were decreased.

This improvement was at a value higher than that of pure MQL, except in the case of low-feed-

low-speed level where the levels of improvement for both cases are the same as shown in Figure

4.3. The findings show that the best surface roughness was 0.625 μm for MQL + 4% gamma-Al2O3

nanoparticles at cutting speed of 240 m/min and feed rate of 0.2 mm/rev. The nanofluid-MQL

showed improvement in surface quality of 15, 58, and 27.1% in comparison with tests performed

at same cutting conditions using MQL, flood, and dry techniques, respectively. The main goal of

using nanoparticles was to enhance a heat transfer coefficient and thermal conductivity of the base

fluid as well as lubricate the cutting zone during machining process. Because the nanoparticles

have a nano-metric sizes and spherical shapes, they can be sprayed into the cutting zone, and then

59

59

be inset into fractured grooves and pores in the cutting zone, as well as between the chips and

cutting tool. Then the nanoparticles lubricate the cutting zone which can largely reduce the friction

force, and then the cutting forces. The better surface obtained by using nanofluid was probably

due to the more effective lubrication and cooling of the tool/workpiece interface and wettability of

work material.

4.1.2.2 ANOVA Analysis

Table 4.3 displays the ANOVA results for the average surface roughness. It showed that the

interaction effect between the lubrication strategy and feed rate was the only significant variable

at 99% confidence level. Additionally, the interaction effect between the cutting speed and feed

rate, and the feed rate effect separately were not significant above 90%; however, they still have

acceptable statistical summation about 1.9 and 1.95, respectively (Table 4.4). The lubrication

technique only did not show a significant effect on the resultant surface roughness; however, it

still has a promising combined effect with the feed rate as their interaction provided the highest

statistical summation. Plot of the interaction effect between the lubrication strategy and cutting

feed rate for the measured average surface roughness data showed the optimum level, which

provided the lowest summation of average surface roughness as shown in Figure 4.4. It was

observed that using MQL with nanofluid and cutting feed rate of 0.2 mm/rev provided the optimal

average surface roughness performance (i.e., A1.C4).

60

60

Table 4.3: ANOVA results for average surface roughness (Ra).

Sources Statistical

summation

Degree of

freedom

Variance F

calculated

Results

Lubrication

technique

0.099 3 Pooled

factor

Feed rate (f) 1.909 1 Pooled

factor

Cutting

speed (v)

0.013 2 Pooled

factor

Lubrication*

v

0.250 6 Pooled

factor

Lubrication* f 2.981 3 0.99 200 Significant at 99%

confidence level

v*f 1.951 2 Pooled

factor

Error 0.099 20 0.00495 Pooled

factor

Total 1.909 23 Pooled

factor

61

61

Figure 4.4: The interaction effect between the lubrication strategy and feed rate for the average

surface roughness results.

Similarly, Table 4.4 shows ANOVA results for tool life. It was found that the interaction effect

between the lubrication strategy and cutting speed was the only significant variables at 95%

confidence level. Similar to the surface roughness ANOVA results, the lubrication technique only

did not show a significant effect on the resultant tool life; however, it still has a promising

combined effect with the feed rate as their interaction provided the highest statistical summation.

The plot of the interaction effect between the lubrication strategy and cutting speed for the

measured tool life shows the optimum level which provided the highest summation of tool life as

shown in Figure 4.5. It was found that cutting speed of 120 m/min with flood or MQL with

nanofluid provided the optimal tool life results (i.e., B1.C2 or B1.C4).

62

62

Table 4.4: ANOVA results for tool life results.

Sources Statistical

summation

Degree of

freedom

Variance F

calculated

Results

Lubrication

technique

2.164 3 Pooled factor

Feed rate (f) 1.44 1 Pooled factor

Cutting

speed (v)

8.48 2 Pooled factor

Lubrication*

v

10.96 6 1.82 14 Significant at

99% confidence

level

Lubrication* f 3.69 3 Pooled factor

v*f 10.10 2 Pooled factor

Error 2.21 20 0.13 Pooled factor

Total 13.18 23 Pooled factor

63

63

Figure 4.5: The interaction effect between the lubrication strategy and feed rate for the tool life

results.

4.1.2.3 Response Surface Methodology (RSM)

Response surface methodology (RSM) was implemented in order to build mathematical models

for each lubrication strategy in terms of cutting speed and feed rate to express the predicted average

surface roughness and tool life. Therefore, analyzing the influence of the studied independent

variables can be demonstrated using this modeling technique as the mathematical model relates

the process responses to facilitate the optimization of the process [146]. RSM technique has been

employed in various machining processes (e.g., milling, turning, and drilling) to study the process

variables effects on specific machining quality characteristics (e.g., average surface roughness,

cutting forces, tool life) [147]. The predicted mathematical models for the average surface

roughness using dry, flood, MQL, and MQL with nanofluid are provided in equations 4.1–4.4,

64

64

respectively. Similarly, the tool life models were developed as shown in equations 4.5–4.8,

respectively. Also, the coefficients of correlations for the all studied cases were determined and

results were listed as in Table 4.5. In order to validate the proposed mathematical models, average

model accuracy is calculated for each case as shown in equation 4.9.

Ra) dry = 7.71 f + 0.000025 v – 0.6856 …………………………………………..................................... (4.1)

Ra) flood = 12.945 f – 0.0118 v – 0.0765 fv + 0.000087v2 + 0.704 … … … … … … … … (4.2)

Ra) MQL = 10.521 f + 0.00849 v – 0.0104 fv – 0.0000168 v2– 1.858 … … … … … . . … (4.3)

Ra) nanofluid = 4.18 f – 0.0078 v + 0.0256 fv + 0.28 ………………...(4.4)

Tool life) dry = – 4.166 f – 0.0121 v + 4.9975 ……………………………….…(4.5)

Tool life) flood = 14.11 f – 0.012 v – 0.0991 fv + 0.00006 v2 + 3.455 … … … … … … … . . . . (4.6)

Tool life) MQL = – 11.43 F − 0.021 v + 0.0366 fv + 0.0000083 v2 + 6.828 … … … . … . (4.7)

Tool life) nanofluid = – 0.086 f – 0.0037 v – 0.033 fv + 4.89 … … … … … … … . . … … . … . (4.8)

Average Model accuracy = ∑ 1 −

Abs(experimental value − predicted value)experimental value

ni=1

n … … . (4.9)

65

65

Table 4.5: The coefficient of correlations results.

Coolant

strategy

Tool life results at feed rate of

0.2 mm/rev

Average surface results at

cutting speed of 180 m/min

Dry

-0.94 0.99

Flood

-0.91 0.91

MQL

-0.86 0.97

Nanofluid -0.84 0.92

The mathematical models developed under dry machining technique for tool life, and average

surface roughness showed average model accuracy about 97.28 and 95.97%, respectively. For

MQL strategy, the proposed tool life mathematical model offered 98.78% average accuracy; while

the average surface roughness model showed about 90.72%. In terms of flood technique results,

the tool life model offered 94.33% average accuracy, and 93.21% average model accuracy was

obtained in terms of average surface roughness results. The average accuracy of the nanofluid

models were 90.3 and 89.2% for average surface roughness and tool life, respectively. 3-D surface

plots for the dry and flood mathematical models are presented as shown in Figures 4.6, 4.7, 4.8,

and 4.9 while Figures 4.10, 4.11, 4.12, and 4.13 show the MQL and nanofluid 3-D surface models.

Figure 4.6: ADI-dry cutting model for surface roughness results.

66

66

Figure 4.7: ADI-dry cutting model for tool life results.

Figure 4.8: ADI-flood cutting model for tool life results.

67

67

Figure 4.9: ADI-flood cutting model for surface roughness results.

Figure 4.10: ADI-MQL cutting model for surface roughness results.

68

68

Figure 4.11: ADI-MQL cutting model for tool life results.

Figure 4.12: ADI-Nanofluid cutting model for tool life results.

Figure 4.13: ADI-Nanofluid cutting model for surface roughness results.

69

69

4.1.2.4 Wear Mechanisms Analysis

Understanding the tool wear mechanisms is important for optimization of the machining process.

From a microstructure perspective, the austenite phase in ADI is thermally stable due to its rich

carbon content. It can undergo a strain-induced transformation when the ADI is locally subjected

to stress; producing sphere particles of hard martensite which enhances tool wear [119]. All SEM

images of worn phases were captured at the end of the cutting tests.

The scanning electron microscope (SEM) images visibly show that the flank wear is the dominated

tool wear. Based on the literature [148, 149] the flank wear is essentially caused by abrasion, where

the hard particles of the workpiece rub against the tool surface. Furthermore, during chip

deformation, particles are pulled out from the contact region of the cutting tool and are taken away

by the flowing chips. This act is responsible for tool flank wear and results in shorter tool life. This

research evaluated the tool wear at the cutting speed of 120 m/ min and feed of 0.2 and 0.3 mm/rev

for different coolant methods. The results showed that the dominant tool wear observation was the

abrasion wear. It was also found that carbon and magnesium particles stacked on the flank face

are most likely due to diffusion wear as can be seen from the energy dispersive X-ray spectroscopy

(EDS) image in Figure 4.14.

70

70

Figure 4.14: EDS image for diffusion on the tool at dry machining f=0.2 mm/rev and V=120

m/min.

Generally, in all coolant strategies, a few spots of work material adhesion were observed on the

insert flank face, which can be attributed to a high cutting temperature at interface zone and the

chemical reaction of workpiece materials. Figure 4.15(a) presents the cutting edge under dry

machining at a cutting speed of 120 m/ min and feed of 0.2 mm/rev. It can be observed that

adhesion wear with a relatively large abrasion wear appeared along the flank face (red circle). In

the case of flood coolant strategy (Figure 4.15 (b)), according to SEM observations, less extent of

abrasion occurred. An amount of chipping (yellow) was also marked on the cutting edge. As

presented previously in Figure 4.2, when using the flood strategy, the cutting tool has longer life

than when using dry condition; however, it appears more severely damaged than dry in SEM

images due to the oxidation, external black layer, resulting from flood coolant Figures 4.15(b) and

4.16 (b).

71

71

Figure 4.15: SEM images of inserts after machining at V=120 m/min and f=0.2 mm/rev under

different coolant strategies.

The most important SEM observation of the flooding strategy is the appearance of thermal damage.

As a chemical reaction wear, thermal damage occurs when chemical compound are formed by

reaction of the cutting tool material with the workpiece material at high temperatures in the

presence of oxygen [150]. As shown in Figure 4.15(b), oxidation wear, which appears as a dark

area, was marked on the extremity of the tool edge (chip/tool contact region). An oxide layer can

act as a barrier against wear at the tool surface; however, it can also accelerate the wear rate.

Usually, during the machining process, the air is available in the chip/tool interface area. The

presence of air, water (in flood coolant), and high temperature was most likely producing the

oxidation wear. Moreover, as seen in Figure 4.15(c), the vegetable oil MQL yielded relatively

72

72

small abrasion, which seen within the red circle. This could be attributed to the lubricant conditions

(i.e., forming a fine mist surrounded with a base cutting fluid film) where the vegetable oil reduces

the rubbing effect between the hard particles and tool surface, generating less friction and less heat.

The main purpose of dispersed gamma-Al2O3 nanoparticles into MQL vegetable oil is to enhance

the thermal conductivity and heat transfer coefficient of the coolant. Therefore, in the case of

nanofluid coolant (Figure 4.15(d)), less abrasion wear was observed, which is labeled by a red

circle. These results indicate that dispersion of gamma- Al2O3 nanoparticles into vegetable oil

improved the performance of the MQL coolant. Furthermore, Figure 4.16 (a) shows the insert

cutting edge with abrasive and adhesive wear after dry machining at cutting speed of 120 m/min

and feed of 0.3 mm/rev.

A clear chipping was observed in the case of flood coolant as described in Figure 4.16 (b), which

could cause a tool failure. The oxidation wear was also observed in the flood strategy at a feed rate

of 0.3 mm/rev. From Figure 4.16 (b), it can be seen that a large thermal damage was located very

close to the cutting tool edge. As mentioned previously, the presence of oxygen and high

temperatures could cause this type of wear mechanism [156]. In the case of MQL, a small chipping

was detected as seen in Figure 4.16 (c). Meanwhile, the nanofluid cooling yielded less abrasion

wears as shown in Figure 4.16 (d). Ultimately, when machining ADI at cutting speed of 120 m/min

and feed of 0.3 mm/ rev, the extent of flank wear was slightly increased (9–13%) compared to feed

of 0.2 mm/rev, and the amount of chipping in the flood method was visibly observed. Increasing

the feed rate value mainly leaded to increase the induced generated heat, thus resulting into quick

thermal softening and rapid cutting edge wear.

73

73

Figure 4.16: SEM images of inserts after machining at V=120 m/min and f=0.3 mm/rev under

different coolant strategies.

4.1.3 Conclusion

Machining of ADI grade 2 under four cooling strategies (i.e., flood, dry, MQL, and MQL with

nanofluid) was investigated. Three levels of cutting speeds were used: 120, 180, and 240 m/min,

and two feeds 0.2 and 0.3 mm/rev. A constant depth of cut of 0.5 mm and length of cut of 50 mm

were used for each pass. L24 mixed orthogonal array based on the design of experiments

techniques was employed. MQL with nanofluid shows promising results for both average surface

roughness and tool life results.

ANOVA is implemented to study the cutting process variables effects on the measured responses.

The interaction effect between the feed rate and cooling strategy is the only significant variable,

which affects the average surface roughness. The tool life ANOVA results show that interaction

74

74

effect between the cutting speed and cooling strategy is the only significant variable. The RSM

technique was employed to develop mathematical models for tool life and average surface for each

cooling strategy and acceptable models accuracies are observed. Regarding the tool wear

mechanisms analysis, the SEM examination revealed that abrasive and adhesive wear mechanisms

are dominant modes of tool wear when machining ADI at different cutting parameters, which can

be attributed to the hardness and chemical reaction of ADI. Due to a presence of air and elevated

temperature at the cutting zone, oxidation wear is noticed when using the flood coolant. Among

the four coolant strategies, the nanofluid provided better results at different cutting parameters in

terms of tool life and surface quality. The performance of MQL vegetable oil is improved after

dispersing 4% gamma- Al2O3 nanoparticles, hence, increasing its thermal conductivity is observed.

Also, it can be concluded that using nanofluids can reduce the induced friction at the cutting zone,

and consequently that could lead to a significant reduction in the cutting forces. The enhancement

in lubricity of nanofluids can be attributed to effect of ball bearing, tribo-film formation, mending,

and polishing effects. The better surface obtained by using nanofluid is probably due to the more

effective lubrication and cooling of the tool/workpiece interface and wettability of work material.

75

75

4.2 Hybrid Nano-Fluid-MQL Strategy for Machining ADI

4.2.1 Introduction

Property such as low thermal conductivity undesirably affects ADI machinability and accelerates

cutting tool failure. Additionally, other issues associated with cutting ADI are the high cutting

temperature, high pressure and dynamic loads, and tendency of chip to adhere to cutting tool face.

To overcome such issues, a proper coolant should be applied. However, flood coolant has

sufficient effects in reducing the generated cutting heat; further alternatives are still required to

decease its environmental and health impacts. Minimum quantity lubrication (MQL) serves as the

best alternative to flood cooling from an environmental perspective as it minimizes the amount of

cutting fluid; however, its heat capacity is lower than the traditional flood coolant. To improve the

cooling and lubricating efficiency of MQL, aluminum oxide (Al2O3) gamma nanoparticles were

used in and its effect on the tool wear behavior during cutting of ADI was investigated.

Maxwell equation for thermal conductivity [151] ,equation 4.10, was applied to determine the

thermal conductivity of resultant nano-fluid, where Kp is the thermal conductivity of Al2O3

nanoparticle, Kf is the thermal conductivity of base fluid (vegetable oil), Km is the thermal

conductivity of resultant, and Vf is the nanoparticles volume fraction. As in Figure 4.17, the

thermal conductivity of the resultant nano-fluid was increased by 9% compared to the base fluid

thermal conductivity.

Km = (Kp+2 Kf+2 (Kp−Kf) Vf

Kp+2 Kf− (Kp−Kf) Vf) Kf ……………. (4.10)

76

76

In the current study, three design parameters were studied (i.e. cutting feed, speed, and added nano-

additives percentage). Table 4.6 shows the design parameters, their corresponding levels, and tests

performed. The MQL method has become the best alternative coolant in different machining

processes because of the environmental concerns and excessive usage of cutting fluid in flood

coolant method. In order to reduce any environmental concern of applying MQL nano-additives,

a standard procedure was implemented. During the experimentation stage, certain safety rules were

followed to keep a specific health and safety level in the machine-shop to avoid any health hazard

to the machine operator [152].

Figure 4.17: Thermal conductivity results (Al2O3 nanofluid vs. vegetable oil).

4.2.2 Results and Discussions

4.2.2.1 Tool Wear

In this study, maximum flank wear (VB) of 0.4 mm was used as the tool life criteria. Figure 4.18

presents the tool wear progress versus the total cutting length. This work revealed that the Al2O3-

nanofluid provided better results compared to regular MQL. On one hand, the lowest flank wear

77

77

(i.e. the longest tool life) was obtained at a cutting speed of 120 m/min, a feed rate of 0.2 mm/rev,

and in the MQL-nanofluid condition (test 7). Furthermore, a 23.5% improvement was observed

in tool life in test 7 compared to test 1, which was performed at the same cutting speed and feed

rate but using regular MQL. On the other hand, the highest flank wear was detected in test 3,

performed at a speed of 240 m/min and a feed rate of 0.2 mm/rev using the regular MQL. On

average, the MQL-nanofluid technique resulted in an improvement of 26.4% in tool life

compared to the regular MQL.

Table 4.6: The studied parameter levels and the plan of experiments.

Test #

Lubrication technique Feed rate (mm/rev) Cutting

speed (m/min)

1 MQL 0.2 120

2 MQL 0.2 180

3 MQL 0.2 240

4 MQL 0.3 120

5 MQL 0.3 180

6 MQL 0.3 240

7 MQL-nanofluid 0.2 120

8 MQL-nanofluid 0.2 180

9 MQL-nanofluid 0.2 240

10 MQL-nanofluid 0.3 120

11 MQL-nanofluid 0.3 180

12 MQL-nanofluid 0.3 240

78

78

Figure 4.18: Tool wear results.

Figure 4.19 shows a comparison between the use of pure MQL and nanofluid- MQL at different

cutting speeds with the same feed rate of 0.2 mm/rev. Another comparison between these strategies

at a feed rate of 0.3 mm/rev was shown in Figure 4.20. Based on the results shown in Figure 4.19

and 4.20, the MQL-nanofluid showed a significant reduction in flank wear, especially at the lower

feed rate (0.2 mm/rev). These reductions can be attributed to the ability of the nanofluid mist to

penetrate easily into the tool-workpiece interface and hence to increase film thickness.

Consequently, a higher film thickness prevents dissipation of the generated heat into the workpiece

and cutting tool material.

79

79

Figure 4.19: Flank wear results at a feed rate of 0.2 mm/rev.

Figure 4.20: Flank wear results at a feed rate of 0.3 mm/rev.

4.2.2.2 Nano-particle Effects

In addition, the rolling action phenomenon [29, 100] of nanoparticles reduces the coefficient of

friction between the cutting tool and workpiece. According to results, using MQL-nanofluid at a

feed rate of 0.2 mm/rev minimized the flank wear by 23.5, 26.5, and 26.4% at a cutting speed of

80

80

120, 180, and 240 m/min, respectively, compared to classical MQL. At a feed rate of 0.3 mm/rev,

using MQL-nanofluid reduced the tool wear by 12.4, 3.62, and 2.8% at a cutting speed of 120,

180, and 240 m/min, respectively, compared to the classical MQL. It has been demonstrated that

the Al2O3 nanoparticles in MQL vegetable oil are atomized in the MQL device with the use of

compressed air. Consequently, a fine mist is created, which is composed of nanoparticles and

vegetable oil (Figure 4.21). This mist is able to penetrate into the tool-workpiece interface zone

and forms a tribofilm. This tribofilm offers a significant function in decreasing the generated

machining heat, as well as minimizing the coefficient of friction at the tool-workpiece interface

zone. Accordingly, these improvements significantly enhance the cooling and lubrication

functions, and maintain tool hardness for a longer period. Therefore, in terms of tool wear

behavior, the MQL-nano-fluid showed better performance compared to the pure MQL as

confirmed in another study [153].

Figure 4.21: The MQL-nanofluid mechanism schematic [153].

81

81

4.2.2.3 Scanning Electron Microscope (SEM) Investigation

However, more reductions in tool wear were observed when using the lower feed rate level (i.e.

0.2 mm/rev). This reduction in tool wear was due to the formation of tribo-film at the tool-

workpiece interface. When a lower feed rate was employed, more nanoparticles can penetrate into

the tool-workpiece interface, and hence, the film thickness could increase. Consequently, higher

film thickness prevents the dissipation of generated heat into both the workpiece and tool materials,

as a result, the cutting process performance enhanced. In addition, the tool wear modes were

investigated for both test 7 and test 3 because they showed the lowest and highest tool wear. The

MQL-nanofluid technique showed less abrasive wear (test 7, Figure 4.22 (a)) as the nanofluid

reduced the rubbing between the cutting tool and work material as a result of tribo-film formation.

The tribo-film helps to reduce the induced friction and, therefore, decreases the generated tensile

stresses. As shown in Figure 4.22 (b), abrasion and adhesion were the dominant tool wear

mechanisms in classical MQL (test 3). These tool wear mechanisms are due to the higher relative

generated heat and friction at the cutting zone. Another set of images was captured for the cutting

tool at a cutting speed of 120 m/min and a feed rate of 0.3 mm/rev using regular MQL and MQL-

nano-fluid shown in Figure 4.23. Noticeable flank wear can be seen in Figure 4.23 (b) (regular

MQL), which is mainly attributed to a hybrid mechanism that includes abrasive and adhesive wear.

82

82

Figure 4.22: Scanning Electron Microscope (SEM) images at a cutting speed of 180 m/min and a

feed rate 0.2 mm/rev using; MQL-Al2O3 nano-fluid (a) and classical MQL (b).

Figure 4.23: The main tool wear modes when machining ADI at a cutting speed of 120 m/min

and a feed rate of 0.3 mm/rev using: (a) MQL-nano-fluid, (b) pure MQL.

83

83

In general, for both MQL and nano-fluid techniques, the adhesion wear was the most significant

tool wear mechanism; however, it can be noted that using MQL-Al2O3-nano-fluid decreased the

flank wear severity in comparison with the classic MQL method (Figure 4.23 (a)). Also, as shown

in Figure 4.23 (b), the crater wear formation was reduced using the nano-fluid coolant.

Furthermore, a cleaner rake face and more a stable cutting edge were obtained on cutting tool when

applying nanofluid coolant. Away from the crater wear investigations, similar tool wear modes

were noticed when machining with pure MQL and MQL-nano-fluid; however, different wear rates

and wear levels were observed. Also, the molten chips and a built-up-edge layer (BUL) were

detected when machining with pure MQL (Figure 4.23 (7)) because of the high contact between

the workpiece and cutting tool under sufficient pressure and temperature. The improvement in

cutting tool conditions when using nano-fluid coolant is mainly due to the enhancement in thermal

conductivity of the nano-fluid coolant which decreases the generated cutting temperature and

reduces the tool wear severity accordingly. Many studies [154–160] revealed that the nano-fluids

have shown effective results on the cutting performance characteristics through different cutting

operations such as milling, turning, and grinding. The suspended nano-additves in the base fluid

produce impressive enhancement in the thermal conductivity and convective heat transfer

coefficient, and consequently a substantial huge amount of generated heat could be carried out

from the cutting zone. Furthermore, the tool wear rate was reduced by using nanofluid-MQL due

to the frictional behavior enhancement as the nano-additives work as spacers in the tool-workpeice

interface area and reduce the induced coefficient of friction. Accordingly, the cutting forces,

friction, tool wear rate, and surface roughness could be minimized.

84

84

4.2.2.4 Nanofluid-MQL Mechanism

In order to gain further improvements on cutting performance using MQL nano-fluids, an

understanding of the mechanisms is highly important. The MQL nano-fluid mechanism can be

summarized as follows (Figure 4.21):

1. By applying a source of compressed air, the nano-cutting fluid is atomized into the

MQL nozzle, resulting in a very fine mist.

2. Thus, the droplets of the nano-cutting-fluid are formed on the workpiece and cutting

tool surfaces and a tribo-film also formed which significantly enhances the

tribological characteristics and reduces the induced friction.

3. The increasing in nano-additive concentration increases the number of nano-

additives at the tool-workpiece interface; these nano-additives perform a vital role as

spacers, reducing the contact between the tool and workpiece.

4. Due to the high compression with increasing the nano-additives concertation, the

nano-additive shape is changed and the shearing is more intense. Some of the nano-

additives in cutting zone are ejected away from tool-workpiece interface by other

additives being sprayed from the nozzle.

5. Therefore, due to the extreme pressure in the naofluid-MQL and existence of a gap

at tool–workpiece interface; nano-additives provide high contact resistance which

helps in forming a chemical reaction film on the workpiece surface. The increase in

nano-additive concentration increase the thickness of this thin protective film on the

machined surface as has been discussed in another study [29, 161].

To sum up, using Al2O3 nano-fluids with MQL showed significant improvements during cutting

ADI. It should be stated that MQL-nano-fluids technique helps in enhancing the interface bonding

85

85

between the tool and workpiece, which generally improves the machinability performance.

Another function of applying MQL-nano-fluids is to dissipate the generated heat, and therefore

the tool wear behavior could be enhanced. Also, the resultant nano-mist works as spacers to reduce

the induced coefficient of friction. Based on the previous considerations, MQL-nano-fluid

achieves a balance between accomplishing a sustainable environment, and enhancing the

machining performance.

Finally, Nanofluid- MQL approach has been successfully introduced as an effective interactive

approach for cooling and lubrication when machining ADI. Applying such an interactive approach

showed promising results in terms of improving the tool wear behavior with maintaining a

sustainable machining environment.

4.2.3 Conclusion

In this study, MQL-nanofluid technique offered significant improvement in tool wear behavior

during cutting of ADI in comparison with the classical MQL. The combination of MQL-nanofluid

at a cutting speed of 120 m/min and a feed rate of 0.2 mm/rev showed the best tool life, while tests

at a higher speed of 240 m/min and a feed rate of 0.2 mm/rev using classical MQL provided the

worst tool wear value. Due to the improved convection, wettability, and conduction characteristics

of the proposed nanofluid, the tool wear was significantly reduced because of the tribo-film

formation along the tool-workpiece interface zone. Also, the nano-fluid mist served as spacers to

limit the induced rubbing between the tool and workpiece, and hence the resultant friction was

reduced. In addition, more improvements in the tool wear behavior were noticed at the lower feed

rate level (0.2 mm/rev). The nano-fluid mist appeared to easily settle on the tool and workpiece

surfaces when the lower feed rate was employed, and therefore the tribo-film thickness increased

and the lubrication effect significantly improved (i.e. protective film mechanism).Moreover,

86

86

similar tool wear modes (i.e. flank and crater wears) were observed when machining ADI using

pure MQL and nanofluid-MQL; however, using nanofluid-MQL, lower wear rates and wear levels

were occurred due to the significant heat and tribological characteristics of the resultant nano-fluid.

Thus, MQL-nano-fluid approach has been successfully introduced as an effective interactive

approach for cooling and lubrication when machining ADI. Finally, more investigations need to

be performed in order to understand the effects of the nano-additives size and concentration on the

coefficient friction and the induced tool wear.

87

87

4.3 - Application of MQL with Nano-Fluids during Machining of Ti-6Al-4V

Alloy: Experimental Investigation and Analysis

4.3.1 Introduction

The cutting fluids are extensive used in almost all kind of manufacturing industries. Apart of work-

tool interface cooling they also perform the lubrication regime. Therefore, they account for

generating good working conditions at the tool work interface. It also leads to the enhancement of

several properties and process parameters like surface topology or cutting forces. On the contrary,

these fluids have some detrimental effects on both human health as well as on the environment.

The majority of such fluids contains potentially harmful chemicals. The recycling costs of such

fluids are very high and their disposal also poses several issues. Operators handling such chemicals

over a longer period of time are bound to suffer from skin or lung related diseases. The past two

decades have witnessed a collective research all over the globe for developing dry or near dry

machining processes to bypass the hazardous chemicals and their use. Due to its simple application

and eco-friendly features, MQL has emerged as one of the best solutions to tackles this challenge.

It has been widely used for most of the machining operations ranging from turning to grinding.

The results obtained after investigating these process clearly indicate the benefits of the MQL

process while machining some different materials such as steel, aluminum, Inconel, and titanium.

Cutting fluids are mainly used to reduce the cutting temperature and friction, and to wash away

chips from the cutting zone during machining operation. When cutting difficult-to-cut materials,

the high amount of heat generated significantly affects the tool life. Therefore, much research has

been attempted to produce cutting fluids with superior tribological and thermal properties linked

with the operator health and environment impacts. Minimum quantity lubrication (MQL) is an

environmentally friendly coolant strategy, but it has leak in cooling capacity.

88

88

Many applications in the field of automotive and aeronautical industries utilize the difficult-to-cut

materials due to their effective properties which are namely high corrosion resistance and better

strength-to-weight ratio [162]. On the other hand, there are characteristics including low thermal

conductivity and high hardness, which have negative effect of the machined surface roughness and

tool life. An understanding of the influence of different machining design variables (e.g., cutting

velocity, tool geometry, coolant strategy, feed rate, depth of cut), has a significant act in improving

machinability, productivity, and reducing the total machining cost. Several studies have been

conducted with an aim to specify the optimal machining variables and select the better coolant

strategy as machining of hard-to-cut materials is still facing different problems particularly,

ceramics, titanium alloys, polymers, nickel-based alloys, composite based materials, and hardened

steels [163-167]. However, the unique characteristics of nickel based alloys, such as, high thermal

fatigue resistance, high erosion resistance, and high melting temperature; the disposal of the

generated heat during the machining is not suitable due to the low heat conductivity of these alloys

[168, 169]. The increase of the machining generated heat over the critical limit, which is 650○C,

leads to negative effects on the machined workpiece and the tool [170]. In general, improper tool

wear behavior and surface integrity are correlated with the machining of nickel alloys [171, 172].

The usage of cutting fluids during machining is mainly to take out the cutting generated heat.

Despite the ability of flood cooling to eliminate machinability problems, there are governmental

restrictions to use it to reduce its crucial impacts on the health and environment. Alternatively, the

use of minimum quantity lubricant (MQL) reduces the cutting fluid quantity during machining.

During machining difficult-to-cut materials, MQL is not a proper alternative of the flood coolant.

To enhance the wettability and fluid thermal aspects, MQL nano-fluid approach has been used [83,

173]. The nano-additives such as aluminum oxide (Al2O3) gamma nano-particles have superior

89

89

cooling and lubrication properties. The effect of the nano-fluids technology on the performance of

machining processes has been studied in several articles. For example, the nano-diamond particles

were dispersed in paraffin and vegetable oils under MQL during micro drilling of aluminum 6061

[178]; aluminum oxide-based nano-fluid has been used in micro-grinding of tool steel [179];

MQL-nano-fluid with chilly CO2 gas was implemented during micro-end-milling of Ti-6Al-4V

[180]. It should be stated that MQL-nano-fluid showed promising results in enhancing the

machining performance in all above-mentioned studies. Furthermore, multi-walled carbon

nanotubes based nano-fluids with MQL showed better results in surface roughness [181] and flank

wear [182] compared to classical MQL and Al2O3 based nano-fluids during machining Ti-6Al-4V.

In addition, many attempts have focused on employing the MQL-nanofluid technique when

machining especially when machining Inconel 718 [183, 184], and austempered ductile iron (ADI)

[98, 153], and promising results have obtained in terms of tool wear, chip morphology, cutting

power, and surface quality. Despite of the previous attempts, there is a research gap in establishing

a multi-objectives model to optimize these machining outputs

In this study, to enhance the thermal conductivity and viscosity of the MQL approach,

nanoparticles were dispersed into the base fluid. Furthermore, the ANOVA analysis was

performed to investigate the effects of cutting speeds, feed rates, and nanoparticle concentrations

during the machining of a titanium Ti-Al6-V4 alloy. Aluminum Oxide (Al2O3) nanoparticles were

selected to use as nano-additives at different weight fraction concentrations (0, 2, and 4%). The

main objective of this work is to study the nanoparticle concentration effects on surface roughness,

tool life, and power consumption when machining Ti-Al6-4V alloy. Besides, a clear tribological

mechanism is presented and discussed to physically justify the nanoparticle concentration effects

on the machining performance.

90

90

4.3.2 Experimentations and Methodology

The machining tests were conducted for Ti-Al6-V4 alloy using the vegetable oil based MQL-

nanofluid coolant. The length of each cutting pass was 50 mm with a depth of cut of 0.5 mm.

Through the use of an MQL system, an Eco-Lubric booster system supplied an air-oil mist with 0.5

MPa air pressure and 40 ml/h nominal oil flow rate, as employed in previous studies [71, 98, 153].

In this section, the Al2O3 nanoparticles with a 22 nm average diameter, 134 m2/g specific surface

areas, 92% purity were used as a nano-additives due to its great tribological property and anti-toxic

aspect (see Figure 4.24). However, the dispersion of nanoparticles into the oil base fluid is

challenging due to microscopic forces that applied on nanoparticles. These forces such as gravity

force, Van der Waals dispersion forces, and density difference promote sedimentation and

nanoparticle agglomeration [185]. The dispersion of nanoparticles into the base oil is crucial

because it affects the viscosity and thermal conductivity of the resultant nanofluid-MQL. In this

regard, recommended to use chemical or physical treatment, such as surfactants, to promote a

sufficient dispersion of nanoparticles [186]. Sodium Dodecyl Sulfate (SDS) was utilized as a

surfactant (0.2 gm). Surfactants are believed to make the nanoparticle performance more

hydrophilic, and to increase the surface charges of the nanoparticles, thereby increasing the

repulsive forces between the nanoparticles [187].

In order to evaluate the agglomerate of nanoparticles into the MQL-nanofluid, Zeta potential for

stability analysis was also conducted. The Zetasizer nano-device was employed to estimate the zeta

potential absolute values for nanofluids with two Al2O3 concentrations. The larger the value of the

Zeta potential absolute, the better the dispersion of Al2O3 nanoparticles and a modeled level of zeta

potential has been noticed. Flank wear and surface roughness data were obtained subsequent to each

machining pass, using a surface roughness tester and an optical microscope respectively. Power

91

91

sight manager device was utilized throughout the cutting operation to assist in recording the power

consumption. Figure 4.25 presents the flow chart of the experimental setup.

The cutting fluid usage when applied as a standard flood coolant can be an environmental concern;

however, when employing the MQL approach, a certain amount of base oil was utilized producing

very fine mist and significantly reducing the volume of coolant used. Safety measures

recommended when using nanoparticles where followed to eliminate any health and environment

concerns associated with utilizing nano-additives. In terms of the disposal process, the nano-fluids

were carefully filtered after used and hazardous waste disposal steps were followed as

recommended by the Environmental Health and Safety (EHS) department [188].

(a) (b)

Figure 4.24: Images of (a) an AQUASONIC-50HT device used to achieve the dispersion and (b) a

regular and nanofluid-MQL.

92

92

Figure 4.25: The experimental schematic.

Through machining tests using an MQL nanofluid with different concentrations of volume fraction

(0, 2, and 4 wt %), various cutting speeds (120, 170, and 220 m/min) and feed rates (0.1, 0.15, and

0.2 mm/rev) were selected as design variables for machining Ti-6Al-4V titanium alloy. Each test

was repeated twice to obtain more reliable data. In this study, three design parameters were selected

with three levels each. Table 4.24 illustrates the independent process factors studied and the

corresponding levels assigned for the machining of titanium alloy workpiece.

The orthogonal array L9 Taguchi method (L9OA) was applied during the cutting experiments.

Experiments and the analysis of the variances (ANOVA) were conducted for each of the design

parameters and labeled as A (cutting speed), B (feed rate), and C (nanoparticle concentration), as

seen in Table 4.7. MINTAB 17 software was used to perform the ANOVA tests, which were used

to explore the influences of design factors on tool wear, machined surface quality and power

consumption. Using the ANOVA, the importance of cutting parameters and nanoparticle

concentrations was investigated with respect to the studied responses to discover the optimum

combination of cutting parameters. All ANOVA analyses were conducted with a confidence level

of 95% (α = 0.5).

93

93

Table 4.7: L9OA with respect to the studied design variables.

Experiment

No.

A:

Speed (m/min)

B:

Feed rate (mm/rev)

C:

Nanoparticles (wt. %)

1 120 0.1 0

2 120 0.15 2

3 120 0.2 4

4 170 0.1 2

5 170 0.15 4

6 170 0.2 0

7 220 0.1 4

8 220 0.15 0

9 220 0.2 2

4.3.3 Results and Discussion

The current work records the findings with regards to maximum flank wear in the tool (VB), surface

roughness (Ra), and power consumption. Table 4.8 describes the results for the machining quality

characteristics that were studied. The flank wear varied from 0.162 to 0.561 mm, and the surface

roughness varied from 0.512 to 2.81µm. The main effects for VB and Ra are shown in Figures 4.26

and 4.28 respectively.

In terms of tool wear, the lowest maximum flank wear (VB) was observed at a speed of 170 m/min,

a feed rate of 0.2 mm/rev, and a nanoparticle concentration of 4 wt%. The speed significantly

affected the tool wear due to high heat generated in the cutting zone. However, in Figure 4.26, it

can be seen that the nanoparticle concentration was also found to affect the tool wear. Table 4.9

shows a 42.65 % contribution with regards to the tool wear.

94

94

Table 4.8: The machining parameter levels and the output measurements.

Experiment No.

A B C Surface

roughness (µm)

Max flank wear (mm)

Power consumption

(W. h)

1 1 1 1 0.891 0.236 2325

2 1 2 2 0.852 0.165 2041

3 1 3 3 1.692 0.162 1959

4 2 1 2 0.512 0.192 2079

5 2 2 3 1.421 0.153 2106

6 2 3 1 2.812 0.563 2454

7 3 1 3 0.563 0.180 2084

8 3 2 1 1.890 0.561 2634

9 3 3 2 0.951 0.242 2236

This effect on tool wear was likely due to the good thermal conductivity of the nanoparticle being

used and its ability to penetrate into the chip/workpiece zone. This improved the machining

performance, the viscous property, and the heat removal ability of the nanofluid. The nanofluid is

atomized into the nozzle of the MQL system and forms as a very fine mist. Accordingly, the nano-

lubricant droplets are formed on the tool and machined surface and a tribo-film is also layered (see

Figure 4.27), which significantly improves the tribological characteristics of the nano-lubricant

and decreases the induced friction in the cutting zone. The improvement in tool wear could also

be affected by the rolling action of the nanoparticles, which changed the wear behavior from

sliding friction to rolling friction.

95

95

Table 4.9: Analysis of variance for maximum flank wear.

Sources Degrees

of freedom

Sum of

squares

Mean of squares

Contribution percentage%

Speed 2 0.031672 0.015836 27.25

Feed 2 0.022406 0.011203 19.27

Nanoparticles wt. % 2 0.049576 0.024788 42.65

Error 2 0.012566 0.006283 10.81

Total 8 0.116221 100

Figure 4.26: The main effects for the maximum flank wear (VB).

120 170 240 0.1 0.15 0.2 0% 2% 4%

96

96

Figure 4.27: Nanofluid-MQL mechanism schematic.

Based on the analysis, the lowest surface roughness was noted at a feed rate of 0.1 mm/rev, a

cutting speed of 170 m/min, and a nanoparticle concentration of 2 wt. %. Feed rate and

nanoparticles wt. % were found to have the greatest impact on surface roughness quality (see

Figure 4.28), contributing 46.55% and 40.75% respectively (see Table 4.10). It is known that the

feed rate has a significant effect on the surface roughness in machining processes. In addition, the

impact of an increase in nanoparticle concentration can be attributed to the increase in

nanoparticles, which can penetrate the chip and workpiece surface and settle in any micro-grooves

or slits. This is known as the surface protective effect of nano-mist.

Table 4.10: Analysis of variance for surface roughness.

Sources Degrees

of freedom

Sum of

squares

Mean of squares

Contribution percentage%

Speed 2 0.373 0.186 8.45

Feed 2 2.0551 1.027 46.55

Nanoparticles wt. % 2 1.799 0.899 40.75

Error 2 0.187 0.093 4.25

Total 8 4.414 100

97

97

Figure 4.28: The main effects plot for surface roughness (Ra).

Regarding the power consumption analysis (see Table 4.11), the lowest power consumption was

obtained at a feed rate of 0.2 mm/rev, a cutting speed of 170 m/min, and a nanoparticle

concentration of 4 wt. %. Feed rate and nanoparticle wt. % were found to have the highest impact

on the measured power consumption; however, the speed still had an acceptable statistical

contribution of about 7.8%, as can be seen in Figure 4.29 and Table 4.11. It was noted that

increasing the nanoparticle concentrations played a vital role in decreasing the induced friction as

the nanoparticles act as spacers at the tool-workpiece contact, and therefore significantly affect the

measured power consumption. However, Ghaednia et al. [189] reported that the increase in

nanoparticles concentration would increase the particles abrasive induced wear as shown in Figure

4.30. Thus, the flank wear rate would grow and thereby would influence the machined surface

finish. In order to fully understand the MQL nanofluid effects on cutting performance, the next

section will investigate the mechanisms related to the nanoparticle effects.

98

98

Table 4.11: Analysis of variance for power consumption.

Sources

Degrees

of

freedom

Sum of

squares

Mean of

squares

Contribution

percentage%

Speed 2 5378.667 2689.33 7.8

Feed 2 49252.67 24626.33 71.3

Nanoparticles wt. % 2 12780.67 6390.33 18.5

Error 2 1664 832 2.4

Total 8 69076 100

Figure 4.29: The main effects plot for power consumption.

99

99

Figure 4.30: Coefficient of friction and induced wear versus nanoparticles content. [189]

4.3.4 Tool Wear Mechanisms

Analysing the tool wear progression is crucial for optimization of the metal cut operation. All

scanning electron microscope (SEM) pictures of damaged tools were taken at the ending of the

machining trials. The images clearly demonstrate that the flank wear is dominated wear phase.

Also, according to Energy Dispersive X-ray Spectroscopy (EDS) images, particles of magnesium

and carbon adhered to the tool rake face most properly due to diffusion wear.

Generally, in all cases (MQL, MQL-Nanofluid 2%, and MQL-Nanofluid 4%) a spots of abrasion

were detected on the flank face due to a rubbing between the cutting tool and workpiece at tool-

workpiece interface. Figure 4.31 shows the condition of the cutting tool after using regular MQL

(without nanoparticles). The results show that an extreme abrasion wear with relatively small

adhesion on flank face tool wear. The flank wear is mostly due to the abrasion wear, since the hard

particles in the material rub against the tool surface. Additionally, during chip deformation, the

100

100

flowing chips take off particles from the contact zone of the tool. This action is responsible for

causing flank wear and effecting the tool life. In addition, due to the high temperature in the cutting

zone and the chemical reaction of work materials, spots of work material adhesion were observed.

Figures 4.32 and 4.33 present the condition of the cutting tool after using MQL-Nanofluid. It can

be seen that a comparatively small abrasion spots observed along the tool flank face using MQL-

Nanofluid 4% compared to using MQL-Nanofluid 2%. Small cater wear was also appeared on the

rake face when increase the nanoparticles concentration into MQL-Nanofluid coolant. It can

attribute that to the tribological aspect of the adding nanoparticle where activated the rolling

mechanism which lead to reduce the rubbing action between tool and workpiece. It should be

stated that increase the density of nanoparticles in contact interface decreased the induced friction

as the nanoparticles act as spacers at the tool-workpiece contact and therefore significantly affect

the tool life.

Figure 4.31: SEM and EDS pictures of the tool after using regular MQL.

101

101

Figure 4.32: SEM and EDS pictures of the tool after using MQL-Nanofluid 2 wt. %.

Figure 4.33: SEM and EDS pictures of the tool after using MQL-Nanofluid 4 wt. %.

102

102

4.3.5 Nanofluid- MQL Mechanisms

The sticking friction zone is known as a seizure zone, which is impenetrable and surrounded by

slipping zone. Under conditions of seizure, tool and work materials become essentially one piece

of metal at the tool-workpiece interface (full sticking friction). According to Trent [190], the

cutting fluid can access the slipping zone, where the apparent contact area in 1/1000 of the real

contact area, while in the seizure zone there is as intrinsic contact. However, in the seizure

condition the the adhered workpiece surface layer often remains attached to the cutting tool edge.

The experimental observations indicated that the nanofluid-MQL method showed lower flank and

crater tool wears compared to the pure MQL. It can contibutted that to the nanoparticle improved

the degree of sliding and reduced the friction force between the tool and machined surface as well

as reduced the possibility of seizure condition. The MQL nanofluid mechanism is shown in Figure

4.34. These nanofluid droplets are applied to the tool-workpiece interface. Two main effects can

be observed while applying the MQL nanofluids: rolling and ploughing mechanisms. These effects

are related to the nanoparticle concentration. The rolling effects result in less contact between the

workpiece and cutting tool. However, increasing the nanoparticles in the cutting zone may limit

the rolling mechanism due to the high compression that results with an increase in the nanoparticle

concentration. This results in a different mechanism known as the ploughing mechanism. In the

ploughing mechanism, the nanoparticles are pushed away from the tool-workpiece zone by other

particles atomized through the MQL nozzle.

The increase in nanoparticle concentration forms a thicker protective film on the tool and machined

surfaces. The nanoparticle concentration is therefore considered a critical design variable to be

optimized when using MQL nanofluid technology, as controlling this variable would have a

significant influence on the induced mechanisms (i.e., rolling or ploughing). Selecting the proper

103

103

nanoparticle concentration would thus help to achieve better frictional behavior while avoiding the

drastic influence of the ploughing mechanism.

Figure 4.34The MQL-nanofluids mechanisms (a) rolling effect and (b) ploughing effect.

Ultimately, the MQL-nanofluid approach is an effective interactive technique for lubricating and

cooling can be used as a sustainable cooling strategy during machining Ti-6Al-4V titanium alloy.

4.3.6 Conclusion

In this work, MQL nanofluid technology was employed when machining Ti-6Al-4V. Cutting

speed, feed rate, and different nanoparticle concentrations (wt. %) was selected as the design

variables to be studied. Three machining outputs were investigated in this study: machined surface

quality; tool life; and power consumption. ANOVA tests were conducted to analyze the design

104

104

variable effects on three machining outputs. According to ANOVA, the feed rate and the

nanoparticle concentration were the most significant parameters in determining the surface

roughness at contribution rates of 46.55% and 40.75% respectively. The ANOVA analysis also

showed that the nanofluid at 2 and 4 wt. % concentration of Al2O3 provided better results than the

regular MQL in determining surface roughness. Tool wear was also found to be affected by

nanoparticle concentration. The analysis indicated that the nanoparticle concentration influenced

the flank wear at a contribution rate of 42.65 %. The nanoparticles which were used enhanced the

thermal conductivity and lubrication properties of the MQL based cutting fluid. They also

minimized the coefficient of friction between the tool-workpiece as well as reduced the possibility

of seizure condition. In terms of the power consumption, the lowest power consumption was

noticed at a feed rate of 0.2 mm/rev, a cutting speed of 170 m/min, and nanoparticle concentration

of 4 wt. %. Furthermore, a detailed mechanism for MQL nanofluid was discussed. Two main

effects (i.e., ploughing and rolling) were presented and investigated. In addition, the relationship

between these effects and the nanoparticle concentrations were highlighted. Applying MQL

nanofluid showed better results in terms of improving the tool performance, surface quality, and

power consumption when compared to the classical MQL. In the future, more work could be done

to determine the appropriate nanoparticle concentration needed to achieve better frictional

behavior while avoiding the drastic influence of the ploughing mechanism.

105

105

Chapter 5: Sustainability Assessment of Machining Difficult-to-Cut

Materials

5.1 Introduction

Recently, when considering manufacturing processes, a sustainability assessment methodology is

a crucial issue. Life Cycle Assessment (LCA) methodology can be utilized to assess the

manufacturing operations. However, it is essential to study the concept of sustainable

manufacturing at different levels, including the process, the product, and the system levels, as the

connections between these levels provide the intended sustainability. Joshi et al. [191] replaced

the 3R method (reduce, reuse, and recycle) of sustainable manufacturing with the new sustainable

manufacturing 6R method (reduce, reuse, recover, redesign, remanufacture, and recycle).

Hypothetically, this accomplishes a closed loop and a multiple life-cycle standard. Although this

new approach (6R) includes a wider range of sustainability characteristics, it does not contain an

independent tool to optimize the process. Thus, an effective assessment tool which may provide

sustainable and optimal solutions is necessary. In this work, a previous assessment algorithm

[101] was used to assess the cutting processes for Ti-6Al-4V titanium alloy to estimate the levels

of sustainable design variables, while considering sustainable indicators as well as the machining

responses (i.e. flank wear (VB) and surface roughness (Ra)). This algorithm concentrates on the

three major aspects of sustainability, namely economic, environmental, and societal. This

algorithm includes five main steps:

defining the machining outputs and sustainable indicators

determining the sustainable factors (i.e. the interaction effect between the machining

outputs and the sustainable indicators)

determining the sustainable index (SI) (i.e. the normalization step)

106

106

determining the weighted sustainable index (WSI ) (i.e. the weighting step)

determining the overall sustainable index for each run (OSI)

Three sustainable metrics were selected in this work, namely energy consumption, environmental

impacts, and the operator's personal health and safety. The carbon dioxide emissions indicator was

used to express the environmental impact. In terms of the operator's personal health and safety,

three indicators were employed: OStc, OShtc, and Pn, which are the toxic effect, the exposure to high

surface temperatures, and the noise effect, respectively.

In this work, four case studies are presented and discussed. These cases are different in terms of

the weighting factors assigned to the sustainable indicators. Equal weighting factors are given for

the machining outputs. The assigned weighting factors for each case are given in Tables 5.1 to 5.4.

Table 5.1: Case 1 Equal weighting factors for all sustainable metrics and indicators.

Case#

1

Machining

Characteristics

Metric

Sustainable Metrics

VB Ra Energy

Consumption

Environmental

Impact

Operator's Personal Health and

Safety

0.5 0.5

0.333 0.333 0.333

Energy CO2 OStc OShtc Pn

1 1 0.333 0.333 0.333

107

107

Table 5.2: Case 2 Different weighting factors for all sustainable metrics and equal weighting for

indicators.

Case#

2

Machining

Characteristics

Metric

Sustainable Metrics

VB Ra Energy

Consumption

Environmental

Impact

Personal Health and Operational

Safety

0.5 0.5

0.25 0.5 0.25

Energy CO2 OStc OShtc Pn

1 1 0.333 0.333 0.333

Table 5.3: Case 3 Different weighting factors for all sustainable metrics and equal weighting for

indicators.

Case#

3

Machining

Characteristics

Metric

Sustainable Metrics

VB Ra Energy

Consumption

Environmental

Impact

Personal Health and

Operational Safety

0.5 0.5

0.5 0.3 0.2

Energy CO2 OStc OShtc Pn

1 1 0.333 0.333 0.333

108

108

Table 5.4: Case 4 Different weighting factors for all sustainable metrics and equal weighting for

indicators.

Case#

4

Machining

Characteristics

Metric

Sustainable Metrics

VB Ra Energy

Consumption

Environmental

Impact

Personal Health and

Operational Safety

0.5 0.5

0.3 0.3 0.4

Energy CO2 OStc OShtc Pn

1 1 0.333 0.333 0.333

5.2 Assessment Results

The algorithm mentioned previously was used to find the levels of sustainable design variables

from the machining results of Ti-6Al-4V titanium alloy. These results are presented in Table 4.8.

The assessment results for Case Study 1 are presented in Tables 5.5a and b. The results of the other

three case studies are appended (Appendix I). Run # 4 provided the optimal and the sustainable

process parameter levels, which were performed at a cutting speed of 180 m/min, a feed rate of

0.1 mm/rev, and a nanoparticle concentration of 2 wt.%. The estimated optimal levels using the

assessment algorithm were in agreement with the flank wear and the energy consumption results.

Furthermore, the algorithm offered the same optimal levels in terms of the surface roughness

(Table 5.6). The current assessment technique considers different machining outputs and

sustainable indicators and serves as a multi-objective optimization solver which achieves a balance

between the machining performances and the sustainability aspects.

109

109

In addition, in all case studies, the nanoparticle concentration of 2 wt. % was the optimal

sustainable value; however, 4 wt. % of nanoparticles offered a better machining performance as

discussed in Chapter 4. Some previous studies have claimed that the interaction between the

ploughing and the rolling mechanisms has a physical relationship with the nanoparticle

concentration [98, 100]. Thus, it can be concluded that the nanoparticle concentration effects need

to be further emphasized in order to understand the physical effect on the machining quality

performance as the nanoparticle affects the tribological and heat transfer characteristics when

machining with nano-cutting fluids.

Table 5.5 (a): Case (1) Sustainable Factors and Sustainable indexes (A: cutting speed, B: feed rate,

and C: nanoparticle concentration wt. %)

110

110

Table 5.5 (b): Case (1): Weighted Sustainable index.

Table 5.6: Assessment algorithm outputs versus optimal machining experimental outputs.

Experimental test responses

Assessment algorithm outputs

optimal machining experimental outputs

Max flank wear (mm) 0.19 0.153

Surface roughness (µm) 0.512 0.512

Energy consumption

(kWh)

2.079 1.959

5.3 Conclusion

In this chapter, the sustainability effects were found to be optimized during the cutting process of

Ti-6Al-4V when using nano-cutting fluids. Achieving a sustainable environment for machining

processes is important, and therefore, the integration of machining outputs and sustainable

indicators is a promising step towards this goal. In this work, three sustainable metrics were

selected, namely energy consumption, environmental impact, and the operator's personal health

and safety. In addition, different weighting factors were employed to find the level of sustainable

process parameters. The results showed that the optimal combination of sustainable process

parameters was a cutting speed of 180 m/min, a feed rate of 0.1 mm/rev, and a nanoparticle

111

111

concentration of 2 wt. % (Run # 4). There was good agreement between the assessment algorithm

outputs and the optimal machining experimental outputs for flank wear, surface roughness, and

energy consumption. The current model has the ability to be a multi-objective optimization solver

because it can take into consideration both machining characteristics and environmental impacts.

Finally, more focus should be given to understanding the physical effects of nanoparticle

concentrations on the tribological and heat transfer characteristics during machining with nano-

cutting fluids.

112

112

Chapter 6: Modelling and Numerical Simulation

An Integrated Numerical Model for Using Minimum Quantity Lubrication

(MQL) When Machining Austempered Ductile Iron (ADI)

6.1 Introduction

The cutting temperature and the thermal behaviour of the cutting process are the key factors

affecting the tool wear mechanism and the tool life. Cutting fluids are required to reduce the high

heat generated during machining process, particularly when machining difficult-to-cut materials

such as austempered ductile iron (ADI) and Ti-6Al-4V [192, 193]. Flood coolant is an effective

approach to eliminate the high generated heat when machining ADI. However, the use of

conventional flood coolant is not a sustainable cooling approach because it costly and negatively

impacts the operator’s health and the environment. The cost of the cooling lubricant ranges from

7% to 17% of the total machining costs [194] while the tool cost is only between 2 to 4% [195].

Previous researches have focused on proposing a sustainable cooling approach when machining

difficult-to-cut materials. Minimum quantity lubrication (MQL) is one of the environmentally

friendly cooling approaches, which has been demonstrated to be an effective near dry machining

technique. Instead of costly experimental research, several researchers have claimed that

computational fluid dynamics CFD analysis coupled with FEM, or virtual prototyping, would be

a powerful method for predicting the temperature of various cutting fluids [196]. Virtual

prototyping remains a cheaper, more environmentally friendly, and faster approach to testing the

effectiveness of the coolant strategy, while CFD analysis is a computational resolution of the

Navier-Stokes equations in the fluid domain [197]. Researchers have tried to model the air-oil

mixture flow using a discrete phase model (DPM). For example, Balan et al. [198] investigated

the oil droplet characteristics as a function of inlet parameters. They found that the droplet size

decreased as the flow rate and atomization pressure increased. They reported that for better

113

113

lubrication and wettability, medium-sized mist droplets with a diameter between 6-16.3 µm should

be used. The solid-mist interaction results were obtained through experiment setups only, using

different flow configurations.

ADI remains one of the hard-to-cut materials, and its machinability has interested several

researchers. Klöpper [199] has dedicated his research efforts to identifying the machining

characteristics of ADI. This study investigated the relationship between the cutting parameters and

different machining outputs. Arft and Klocke [200] analyzed the effect of tool geometry on cutting

forces. An optimized geometry was designed by means of FE analysis. However, few researchers

have investigated the MQL application when machining ADI. For example, Eltaggaz et al. [98,

153] investigated the relationship between design cutting parameters and machining aspects (tool

life, surface roughness, and power consumption) of ADI using dry, flood, MQL, and nanofluid-

MQL cooling conditions. It was found that adding nanoparticles to the MQL base oil (nanofluid-

MQL) extended the tool life and improved the machined surface quality. Previous studies were

mainly focused on using MQL-nano-fluid approach while turning alloys such as Inconel 718 [100]

and Ti-6Al-4V [29].

To the best of the author’s knowledge, there is a gap in the open literature on the ADI turning

process regarding numerical modelling of the thermal interaction. Thus, this research focuses on

simulating the cutting heat distribution on the cutting tool during the machining of ADI. It also

develops a model to predict the cutting tool tip temperature and to evaluate the MQL's cooling

ability. As the constitutive equation for ADI is not available by default because of its contradictory

nature, there is no study in which an FE model of machining simulation is performed for ADI.

114

114

In this study, a 2D commercial ABAQUS explicit code was used with the user defined subroutine

VUHARD in order to determine the maximum temperature when cutting ADI under MQL and dry

conditions. The ABAQUS subroutine "VUHARD" was therefore coded for this purpose: the

subroutine's role is to calculate the equivalent stress in each iteration according to the stress flow

formula given by the Broddmann Equation [201]. The subroutine was coded in FORTRAN and

then added to the Abaqus model to solve the problem. Afterwards, the resultant temperature was

mapped into an ANSYS fluent CFD/CHT (coupled heat transfer) to investigate the convective heat

transfer occurring on the rake face since this area is the most exposed to the coolant flow. The

CFD model compared the MQL coolant strategy with dry conditions to predict the heat distribution

on the cutting tool and to evaluate the MQL thermal efficiency. Experimental tests, conducted

using a CNC Lathe on Grade 2 (ADI), validated the numerical model’s results for both dry and

MQL approaches. Thus, the novelty of this work is mainly based on the interactive design

approach as this work offers a new integrated numerical model (FEM & CFD) to simulate the

thermal and heat transfer behavior when machining ADI.

6.2 Experimental Setup

The methodology schematic is shown in Figure 6.1. The FEM of the cutting operation was

developed using the recommended cutting parameters [202]. The temperature of the cutting tool,

which was invoked in the CFD model, was generated by an FEM model and set as a thermal

boundary condition. An accurate instantaneous estimation of the heat coefficient was obtained and

mapped into a transient thermal analysis to observe the temperature evolution as a function of

distance and cutting time. The major difference between the MQL technique and the conventional

fluid coolants is the added conduction that occurs at the exposed surface. Hence, the oil film that

is formed acts as a heat sink. The results from the model were compared to the experiments.

115

115

Figure 6.1: Methodology schematic of thermal behavior prediction for machining ADI.

The experimental setup of the temperature measuring arrangement is shown in Figure 6.2. Cutting

parameters were selected in accordance with the literature [203]. Table 6.1 illustrates the

machining parameters which were applied during the machining process in this study.

Table 6.1: Machining parameters and equipment specifications.

Machining Parameters Value

Cutting speed (m/min) 140

Feed rate (mm/rev) 0.1

Depth of cut (mm) 0.5

Cutting environments Dry and MQL approach

A metal K-type thermocouple, which was 1.6 mm in diameter, was fitted through the tool holder

and seat, and embedded into the insert as shown in Figure 6.2 (b, c & d). The thermocouple was

connected to an HH 374 Omega data logger. This data logger was employed to save the

116

116

temperature values that were obtained during each experiment. Due to the constant contact

between the cutting tool and the workpiece, the thermocouple probe could not be set at the tip. As

seen in Figure 6.2 (b), the thermocouple probe was situated approximately 2 mm away from the

cutting tool tip, where it measured the temperature. The temperature was determined at this point

2 mm from tool tip during the machining of ADI grade 2 under both dry conditions and an MQL

mist. The tooltip temperature can be extrapolated from the numerical simulation after validation.

The fluid supply chain ends with a convergent nozzle ordered for the purpose. The nozzle

orientation was set manually in order to ensure the maximum wettability of the contact zone. The

outlet nozzle diameter was 2 mm, and it was confirmed that no pressure loss occurred in the fluid

channel.

117

117

Figure 6.2: Experimental setup view (a); the installation and arrangements for the cutting

temperature measurements (b), (c) & (d).

The machining experiment was executed using a sample of ADI grade 2 as a workpiece. Table 6.2

presents the thermo-mechanical properties of ADI grade 2 as stated in the ASTM A897 standard

specifications. Table 6.3 shows the properties of the uncoated tungsten carbide cutting tool. It

consists of H13A grade materials which are known for their strong wear resistance and toughness.

118

118

Table 6.2: Thermo-mechanical properties of the ADI grade 2. [201, 202]

Property Value Property Value

Modulus of elasticity (GPa) 157.9 Thermal conductivity (w/m.k) 21.8

Poisson’s ratio 0.25 Specific heat (J/Kg.k) 461

Thermal expansion coefficient

(mm/mm/°C×10-6)

14.3 Modulus of elasticity (GPa) 157.9

Table 6.3: Thermo-mechanical properties of CCMT 12 04 04-KM (uncoated carbide). [203]

Property Value

Modulus of elasticity (GPa) 670

Poisson’s ratio 0.25

Thermal conductivity (w/m.k) 46

Specific heat (J/Kg.k) 203

119

119

6.3 Finite Element Modeling

6.3.1 ADI Model for Orthogonal Turning Process

In order to obtain the temperature distribution throughout the workpiece and the cutting tool, a

finite element model of the turning operation was simulated using a 2D commercial explicit

dynamic code. ABAQUS Explicit 14.6 was used as it is robust and reliable. Using an Explicit

dynamic solver is highly recommended for extreme nonlinear material behaviour and phenomena

involving high deformations and strains [204]. The geometry was sketched with the real

dimensions of the cutting tool, and the features of the assembly were meshed using CPERT4

elements which allowed the temperature to be calculated as a nodal variable. The flow stress

behaviour of the ADI was modelled using the Brodmann formulation [201] with the appropriate

material constants as advised in the literature [199, 202]. This material model is not available by

default; therefore the user defined subroutine VUHARD was used in each iteration for flow stress

calculations. An accurate flow stress calculation is important for a representative thermal behavior

since the temperature increase is due to plastic deformation, as shown in Equation (6.1):

ρ. c(T). T = η. σ. εpl … … … … … … … … … … … … … … … … … … … (6.1)

Where ρ is the material density, c is the material specific heat, T is the temperature rate, and η is

the inelastic heat fraction which is set to 0.9 for this study, while σ and εpl are the flow stress and

the plastic strain respectively. Equation (6.2) presents the constitutive equation, which describes

the flow stress as function of the strain, strain rate and temperature:

σ(ε, ε, T) = (K1 + K. εn + η. ε). ψ(T). χ(ε, ε, T) … … … … … … … … … … … . . (6.2)

120

120

Where (T) describes the thermal softening that occurs due to the high temperature values reached

during cutting as shown in Equation (6.3). In addition, the χ function is added to take into account

the ductile softening behavior as shown in Equation (6.4). The design parameters used for this

study are provided in Table 6.4.

(T) = e−β.(T−T0)

Tm … … … … … … … … … … … … … … . (6.3)

χ(ε, ε, T) = 1 − e−εG.εG.T

ε.ε.TG … … … … … … … … … … … … … … (6.4)

Table 6.4 Parameters for constitutive material law. [199]

Parameter Value Parameter Value Parameter Value

K1 55 [MPa] K 2000.6 [MPa] n 0.1006

η ]2Ns/m[ 0.0292 mT 1800 [K] β 3 [MPa/K]

Gε 33000 εG ]1-s[2000 GT 100 [K]

A fully coupled temperature displacement explicit solver was chosen with a local Arbitrary

Lagranian Eulerian (ALE) formulation in order to prevent distortion errors; enhanced hourglass

control was also activated to prevent any hourglass effect. The mesh was finer in the cutting zone.

The model results have been ensured to be independent of mesh size through performing several

comparisons with other trials with smaller mesh size. The total number of elements was 300, 000.

No intermediate sacrifice layer was therefore defined between the chip and the non-machined

workpiece zone. Boundary conditions were neatly introduced to replicate the cutting conditions,

as shown in Figure 6.3. The machined part of the workpiece was sloped to prevent distortion. The

simulation was carried out as a 2D representation; the depth of the cut was set as a shell thickness

121

121

of 0.5 mm. The cutting tool was characterized by a rake angle of γ=0°, while the clearance angle

was θ=17°. In order to replicate the complex contact status of the tool-chip interface, a shear

friction criteria with a frictional coefficient of m=0.3 was invoked. This study did not involve a

chip separation criterion since the ALE re-meshing rule was activated.

Figure 6.3: Meshing and boundary conditions definition.

6.3.2 FEM Results

The temperature and heat flux contours were obtained as shown in Figure 6.4. The temperature

plots show the temperature evolution on the tooltip as a function of the cutting time (Figure 6.4-a)

and temperature values across the constrained part of the rake face (Figure 6.4-b). The maximum

temperature reached at the cited cutting conditions was approximately 629°C. The chip sliding

effect over the rake face significantly increased the temperature. The heat flux contours were also

presented in Figure 6.5.

122

122

Figure 6.4: (a) Finite element simulated cutting temperature, (b) Temperature values across the

.C)orake face (

Figure 6.5: The heat flux contours.

(mm)

123

123

6.4 CFD Modeling

6.4.1 FEM Results

During the cutting process, heat transfer cannot be negligible when temperature gradients take

place. A CFD model was created to solve the momentum, continuity and energy equations in both

the solid and the fluid domains (i.e. a conjugate heat transfer model). This section investigates the

occurrence of heat transfer between the cutting tool and the environment. When using pressured

air cooling and conventional flood cooling, heat convection occurs on the rake face of the cutting

tool because this face is mostly exposed to the cooling flow. However, in the current case (MQL),

a thin layer of the base cutting fluid was observed. This enabled the addition of the conduction of

the tool's dissipated heat as shown in Figure 6.6.

Figure 6.6: MQL versus other cooling methods in terms of heat transfer mechanisms.

6.4.1.1 Geometry for CFD Model: the geometry was designed for the CNMG433-NR IC5005

cutting insert. In order to delimit the fluid domain, an enclosure was built around the cutting

tool. It is important to share the topology between the fluid and the solid domains to avoid the

124

124

automatic generation of bad faces in the interface. The nozzle was designed to conform to the

nozzle geometry that was used in the experimentation phase. The whole schematic was an

accurate imitation of the real experimental conditions shown in Figure 6.7.

Figure 6.7: Geometry setup for the CFD analysis.

6.4.1.2 Meshing: The CFD model was meshed using tetrahedral elements with an automatic

refinement in the borders and the contact interface. As shown in Figure 6.8, the mesh around

the heat wall contact of the solid bodies with the fluid domain had to be very fine to accurately

model the energy transfer interactions caused by the injected coolant strategy. There are six

sections of the model that should incorporate a named selection prior to importing the mesh

into the setup. The initial named selection is the Heat-Wall of the model, of which the required

face can be seen selected. The next named selection was the Flank face. The remaining solid

domain walls were assigned to the named selection Insert-Wall. It is not important to select the

contact face between the heat tip and the insert, because this would cause a conflict in the mesh

interfaces when they are imported to Fluent. The outlet face of the enclosure walls was assigned

125

125

the named selection outlet. The remaining walls of the enclosure were selected and titled

Enclosure. In addition, the results of the CFD model have been ensured to be independent of

mesh size through performing several comparisons with other trials with smaller mesh size.

Figure 6.8: Meshing setup for the CFD analysis.

6.4.1.3 Modeling approach: A density based solver with steady state energy and momentum

equation resolutions was used to control the time step to avoid divergence in the solution. The

energy equation was enabled after 10 iterations of the momentum equations. The turbulence model

used in this case was the SST (Shear Stress Transport) model which is known for its accuracy in

near wall conditions (i.e. the fluid solid interface). The material properties were introduced

depending on the used cutting fluid.

6.4.2 CFD Results under Dry and MQL Conditions

The CFD model was built in order to obtain the velocity, the pressure, the temperature contours,

and the convective heat coefficient, which all varied according to the fluid that was being

modelled. As is shown in (Figure 6.9-b), the heat transfer coefficient varies along the rake face

126

126

and reaches its maximum point right on the tooltip. This is introduced as an imported load to the

transient thermal model to determine the temperature evolution at one specific point. Figure (6.9-

a) shows the velocity contours of the pressured air and the convective heat coefficient right on the

rake face of the cutting tool. Two models, dry machining and MQL, were built to evaluate the heat

exchange. The loads are transferred from Fluent to the transient thermal solver within the ANSYS

workbench workspace. The thermal solver was used to add a thin layer of oil when investigating

the MQL method. The most important part was to establish the link between the ANSYS Fluent

and the ANSYS transient thermal, as shown in Figure 6.10.

Figure 6.9: Velocity contours (a), and heat transfer coefficient distribution (b).

127

127

Figure 6.10: ANSYS Fluent/ANSYS transient thermal coupled link.

Regarding the dry machining results, A FEM model was developed to predict the cutting tool tip

temperature when machining ADI using dry conditions. The temperature recorded at a distance of

2 mm away from the tool nose was 109.7oC (approximately 2mm inside the rake face). To verify

the FEM model results, dry machining was conducted on the ADI workpiece. The thermocouple

was placed 2 mm away from the cutting tool tip. Figure 6.11 illustrates the results obtained from

the thermocouple data logger, which are plotted as temperature (oC) vs time (s). The dry cutting

temperature distribution continually rises due to the lack of coolant. The maximum temperature

during machining was 116oC. Due to the location of the thermocouples, the temperature was

recorded over ten seconds period while cutting was being performed. The simulated temperature

values were in good agreement with the experimental results with an error of 5.74%.

128

128

Figure 6.11: Comparison between the experimental and the simulation temperatures under dry

machining conditions.

In terms of MQL results, the modelled temperature at a distance of 2 mm from the nose was

55.01oC at ten seconds into the machining process. The recorded temperature from the

thermocouple data logger was plotted as temperature (oC) vs time (s), as shown in Figure 6.12.

The maximum temperature during the experimental machining time was 49.1oC. This indicated

that there was a 10.74% margin of error between the results from the experimental and the

numerical models. The error in the case of the MQL approach was higher relative to that of the

dry condition.

129

129

Figure 6.12: Comparison between the experimental and simulation temperatures under MQL

conditions.

This is still a preliminary attempt to simulate the effect of MQL on the resultant temperature when

machining ADI; however, the main contribution of the current FEM analysis was the development

of a constitutive equation for the machining of ADI grade 2. It should be stated that the current

model does not consider the tribological effects of MQL. In addition, the oil layer was assumed to

be uniformly distributed, which is not accurate enough to simulate the effect of the resultant mist

from the MQL system. Furthermore, possible methods of implementing this work with the aid of

the developed computational fluid dynamics model would be to convert the model into a transient

simulation activating the wall film boundary conditions on the walls of the cutting tool. Thus,

further investigations need to be performed in order to achieve better accuracy of the integrated

numerical model (FEM & CFD).

130

130

6.5 Conclusion

The current work focuses on simulating the cutting heat distribution on the cutting tool during

machining of ADI by developing an integrated numerical model (finite element & computational

fluid dynamics). The investigations within this work were focused on the cutting temperature (i.e.

2 mm away from the tool nose for both dry cutting and MQL cooling). In terms of dry cutting, the

simulated temperature values were in good agreement with the experimental results with an error

of 5.74%. Regarding the MQL, there was a 10.74% margin of error between the results from the

experimental and the integrated numerical model. Overall, the temperatures obtained using the

model for the dry condition was in good agreement with the experimentally obtained temperatures.

However, the CFD model error in the MQL simulation was relatively high when compared to the

dry test. This is still a preliminary attempt to simulate the effect of MQL on the resultant

temperature when machining ADI; however, the main contribution of the current FEM analysis

was the development of a constitutive equation for the machining of ADI grade 2. Finally, further

modifications can be added to the current integrated model to enhance its accuracy such as;

considering the tribological effects of MQL, and developing a transient CFD model.

131

131

Chapter 7: Conclusions and Future Work

This chapter provides the conclusions drawn from the current research and presents the plans for

future work in this area. This thesis focused on investigating the effects of cutting design variables

and cooling strategies on machining characteristics during the machining of both a Ti-6Al-4V

alloy and ADI. The current research also attempted to enhance the cooling ability, viscosity, and

thermal conductivity of the sustainable cooling strategy (nanofluid-MQL) by adding Al2O3 nano-

additives. This was achieved by performing turning tests with appropriate cutting tools and the

recommended range of cutting parameters. In addition to experimental activities, finite element

modeling (FEM) was used to predict the tool tip temperature and the distribution of the temperature

in the cutting zone when machining ADI under dry conditions. Computational fluid dynamics

(CFD) was also used to evaluate the MQL cooling performance during the turning of ADI. The

main conclusions from this work can be summarized as follows:

The overall cutting performance benefits from applying a nanofluid-MQL as this is a

sustainable cooling strategy with a good performance. The use of a nanofluid-MQL

reduced the cutting temperature and the friction in the cutting zone, and consequently

reduced the tool wear, the power consumption, and the cutting forces. Adding Al2O3

nanoparticles to the vegetable-oil based MQL markedly enhanced the lubricity and the

thermal conductivity of the nanofluid-MQL, and improved its machining performance by

enhancing its viscosity and thermal conductivity.

The nanofluid-MQL technique offered a significant improvement in tool life during the

machining experiments in comparison with the classical MQL. The combination of

nanofluid-MQL 4 wt. % at a cutting speed of 120 m/min and a feed rate of 0.2 mm/rev

showed the best tool life during the machining of ADI.

132

132

Nano-additives placed in a vegetable-oil MQL cooling system improved the convection,

wettability, and conduction characteristics of the proposed nanofluid-MQL; therefore, the

tool wear was significantly reduced because of the tribo-film formation along the tool-

workpiece interface zone. Also, the nanofluid-MQL mist droplets served as spacers to limit

the rubbing induced between the cutting tool and the workpiece, thereby reducing the

coefficient of friction in the cutting zone. Moreover, using a nanofluid-MQL decreased the

rate of tool wear due to the enhanced tribological characteristics contributed by the nano-

additives. Thus, this research has successfully introduced nanofluid-MQL as an effective

interactive approach for cooling and lubrication when machining difficult-to-cut materials

such as ADI and Ti-6Al-4V titanium alloy.

It was also found that the nanoparticle concentration plays a vital role in the tribological

characteristics of the nanofluid-MQL.

The FEM and CFD models used respectively for predicting the cutting temperature distribution

under dry conditions and for evaluating the cooling performance of the MQL approach during the

machining of ADI were found to be in good agreement with the experimental results. The main

contribution of the current FEM analysis was the development of a constitutive equation for the

machining of ADI grade 2. Further modifications can be added to the current integrated model to

enhance its accuracy such as; considering the tribological effects of MQL, and developing a

transient CFD model.

133

133

7.1 Contributions

The current research has contributed to the present knowledge as follows:

• This study addresses the research gap in the literature which relates to the investigation of

nanofluid cooling strategies during the machining of ADI grade 2 and Ti-6Al-4V titanium

alloy using nanoparticles.

• The current study provided a clear understanding of the role of MQL-nano-cutting fluids,

their cooling and lubricating mechanisms, and the ploughing and rolling mechanisms.

• The knowledge gained during this study regarding machining operations, cutting tool

performance, tool wear behavior,and power (energy) consumption can be a solid

foundation to further approaches for improving the machinability of these materials under

a sustainable cooling strategy (e.g. nanofluid- MQL).

• The main contribution of the preliminary numerical phase was the development of the

constitutive equation for the machining of ADI grade 2, which is not readily available in

the open literature.

• The current research developed the knowledge and tools needed to address both

sustainability and technical challenges associated with the machining of difficult-to-cut

materials in terms of tool wear, and surface quality.

7.2 Future Work

Future research should focus on:

• Including more design variables such as different MQL air pressures, MQL volume flow

rates, MQL nozzle orientations, and cutting tool geometries.

134

134

• Investigating a hybrid nano-cutting-fluid; for example, a mixture between MWCNTs and

Al2O3 nanoparticles using volume fractions ranging from 0 to 8 %, as suggested in the

literature.

• Developing an analytical model to explain the relationship between induced nanoparticle

abrasion and nanoparticle concentration and size.

• Performing further research to develop a function to represent the friction coefficient under

MQL operations, which in turn could be used to more accurately model the temperature

generated at the tooltip.

• Enhancing the FEM model and CFD simulation to simulate the MQL-nanofluid.

135

135

References:

[1] Gupta K., Laubscher R. F., and Davim P., "Recent developments in sustainable manufacturing

of gears: a review," Journal of Cleaner Production, vol. 112, pp. 3320-3330, 2016.

[2] Tazehkandi A. H., Shabgard M., Kiani G., and Pilehvarian F., "Investigation of the influences

of polycrystalline cubic boron nitride (PCBN) tool on the reduction of cutting fluid consumption

and increase of machining parameters range in turning Inconel 783 using spray mode of cutting

fluid with compressed air," Journal of Cleaner Production, vol. 135, pp. 1637-1649, 2016.

[3] Davoodi B. and Tazehkandi A. H., "Experimental investigation and optimization of cutting

parameters in dry and wet machining of aluminum alloy 5083 in order to remove cutting fluid,"

Journal of Cleaner Production, vol. 68, pp. 234-242, 2014.

[4] Niketh S. and Samuel G. L., "Surface texturing for tribology enhancement and its application

on drill tool for the sustainable machining of titanium alloy," Journal of Cleaner Production,

vol. 167, pp. 253-270, 2017.

[5] Tazehkandi H., Pilehvarian F.,and Davoodi B., "Experimental investigation on removing cutting

fluid from turning of Inconel 725 with coated carbide tools," Journal of Cleaner Production,

vol. 80, no. 1, pp. 271-281, 2014.

[6] Park D., "The occupational exposure limit for fluid aerosol generated in metalworking

operations: limitations and recommendations," Safety and Health at Work 3.1, pp. 1-10, 8

September 2012.

[7] Bordin A., Sartori S., Bruschi S., and Ghiotti A., "Experimental investigation on the feasibility

of dry and cryogenic machining as sustainable strategies when turning Ti-6Al-4V produced by

Additive Manufacturing," Journal of Cleaner Production, vol. 142, pp. 4142-4151, 2017.

[8] Çolak O., "Optimization of machining performance in high-pressure assisted turning of Ti-6Al-

4V Alloy," Journal of Mechanical Engineering, vol. 60, no. 10, pp. 675-681, 2014.

[8] Metal Asian., "ASIAN METAL , Metal pedia, Titanium uses," ASIAN METAL , [Online].

Available: http://metalpedia.asianmetal.com/metal/titanium/application.shtml. [Accessed 15 04

2016].

[10] Meena A. and El Mansori M., "Study of dry minimum quantity lubrication drilling of novel

austempered ductile iron (ADI) for automotive applications," Elsevier, Wear, vol. 271, no. 9-

10, pp. 2412-2416, 2011.

[11] Klocke F. and Klopper C., "Machining of ADI," WZL Laboratory for Machine Tools and

Production Engineering, Aachen, Germany.

[12] Brandenberg K., "Successfully machining austempered ductile iron (ADI)," Applied Process

Inc., Michgen, USA, 2001.

136

136

[13] Datt D., and Batra U., "Influence of composition and austempering temperature on

machinability of asutempered dutile iron," International Journal of Chemical, Molecular,

Nuclear, Metallurgical Engineering, vol. 7, no. 2, pp. 132-137, 2013.

[14] Padan D. S., "Microalloying in austempered ductile iron (ADI)," American Foundry Society, IL,

USA, 2012.

[15] Nofal A., "Advanced in the metallurgy and applications of ADI," Journal of Metallurgical

Engineering (ME), vol. 2, no. 1, 2013.

[16] ADI Treatments Limited, "Austempered Ductile Iron," ADI Treatments Ltd, [Online].

Available: http://www.aditreatments.com/?p=424. [Accessed 27 11 2016].

[17] esa, "IMPRESS, Supporting Education Across Europe," esa, [Online]. Available:

http://www.spaceflight.esa.int/impress/text/education/Circular%20Motion/Turbine_Blades.ht

ml. [Accessed 24 11 2016].

[18] Veiga C., Davim J. P., and Loureiro A., "Review on machinability of titanium alloys:the process

perspective," Reviews on Advanced Materials science, vol. 34, pp. 148-164, 2013.

[19] Arcam EBM system, "Arcam-Ti-6Al-4V-Titanium-Alloy," Arcam AB, [Online]. Available:

http://www.arcam.com/wp-content/uploads/Arcam-Ti6Al4V-Titanium-Alloy.pdf. [Accessed 7

6 2018].

[20] Akdemir A., "The effect of cutting speed and depth of cut on machinability characteristics of

austempered ductile iron," Journal of Manufacturing Science and Engineering, vol. 134, no. 2,

2012.

[21] Parhad P., Likhite A., Bhatt J., and Peshwe D., "The effect of cutting speed and depth of cut on

surface roughness during machining of austempered ductile iron," Trans Indian Inst Metals, vol.

68, no. 1, pp. 99-108, 2015.

[22] Hegde A., and Sharma S., "Machinability study of manganese alloyed austempered ductile

iron," Journal of the Brazilian Society of Mechanical Sciences and Engineering, pp. 40: 338.

https://doi.org/10.1007/s40430-018-1258-6, 2018.

[23] Meena A., and El Mansori M.,"Specific cutting force, tool wear and chip morphology

characteristics during dry drilling of austempered ductile iron (ADI)," International Journal of

Advance Manufacture Technology, vol. 69, p. 2833–2841, 2013.

[24] Polishetty A., "Machinability and microstructural studies on phase transformations in

Austempered Ductile Iron," PhD Thesis, Auckland University, pp. 46-56, 2011.

[25] Arft M., Klocke F., and Lung D., "Evaluation of the machining aspects of austempered ductile

iron," International Journal of Metalcasting , pp. 35-42, 2012.

[26] Polishetty A., Moshe G., and Littlefair G., "Wear characteristics of ultra-hard cutting tools when

machining ADI," International Journal of Mechanical and Mechatronics Engineering, vol. 10,

no. 1, 2011.

137

137

[27] Meena A., El Mansori M., and Ghidossi P., "Machinability of austempered ductile iron (ADI)

produced by integrated green technology of continuous casting-heat treatment processes," in

International Conference on Advances in Materials and Processing Technologies (AMPT2010),

Paris, France, 2010.

[28] Sinlah A., Handayani D., Voigt R. C., Hayrynen K., Saoub R. M., and Saldana C., "Effects of

microstructure and strength on wear performance in rough milling of austempered ductile iron,"

International Journal of Cast Metals Research, Vols. 1-2, no. 29, pp. 62-67, 2016.

[29] Hegab H., Umer U., Deiab I., and Kishawy H., "Performance evaluation of Ti-6Al-4V

machining using nano-cutting fluids under minimum quantity lubrication," The International

Journal of Advanced Manufacturing Technology, vol. 95, no. 9-12, pp. 4229-4241, 2018.

[30] Hartung P., and Kramer D., "Tool wear in titanium machining," CIRP, , vol. 31, pp. 75-80, 1982.

[31] Maekawa K. and Kitagawa, "Temperature and wear of cutting tools in high-speed machining of

Inconel 718 and Ti-6Al-6V-2Sn," Wear, vol. 2, 1997.

[32] Tuğrul Ö. and Karpat Y., "Multi-objective optimization for turning processes using neural

network modeling and dynamic-neighborhood particle swarm optimization," International

Journal Advance Manufacturing Technology, vol. 8, pp. 234-247, 2007.

[33] Venkata Ramana M., Krishna Mohan Rao G., and Hanumantha Rao D., "Optimization and effect

of process parameters on tool wear in turning of titanium alloy under different machining

conditions," International Journal of Materials, Mechanics and Manufacturing, vol. 2, no. 4,

pp. 272-277, 2014.

[34] Ojolo S. T. and Ogunkomaiya O., "A Study of effect of machining parameters on tool life,"

International Journal of Materials Science and Applications, vol. 5, no. 3, pp. 183-199, 2014.

[35] Narutaki N. and Murakoshi A., "Study on machining of titanium alloys," CIRP Annals -

Manufacturing Technology, vol. 32, no. 1, pp. 65-69, 1983.

[36] Liu Z., Xu J., Han S., and Chen M., "A Coupling method of response surfaces (CRSM) for

cutting parameters optimization in machining titanium alloy under minimum quantity

lubrication (MQL) condition," International Journal of Precision Engineering and

Manufacturing, vol. 14, no. 5, pp. 693-702, 2013.

[37] Narasimhulu A., Venkateswara R., and Sudarsan G., "Dry Machining of Ti-6Al-4V using PVD

coated TiAlN tools," in Proceedings of the World Congress on Engineering, Vol III, London,

UK, 2012.

[38] Ying-lin, Hui-yue, LIU G., and Ming Z., " Use of nitrogen gas in high-speed milling of Ti-6Al-

4V," International Journal of Machine Tools & Manufacture, vol. 49, p. 739–743, 2009.

[39] Heigel J., Whitenton E., Lane B., Donmez M., Madhavan V., and Moscoso-Kingsley W.,

"Infrared measurement of the temperature at the tool–chip interface while machining Ti–6Al–

4V," Journal of Materials Processing Technology, vol. 243, p. 123–130, 2017.

138

138

[40] Ramana V., "Optimization and influence of process parameters on surface roughness in Turning

of titanium alloy under different lubricant conditions," Materials Today: Proceedings, vol. 4 ,

p. 8328–8335, 2017.

[41] Boswell B. and Islam M., "Sustainable cooling method for machining titanium alloy," Materials

Science and Engineering, vol. 114, pp. doi:10.1088/1757-899X/114/1/012021, 2016.

[42] Shaw M. C., "Metal Cutting Principles", New York: Oxford University Press, 1998.

[43] Sharma A. K., Katiyar J. K., Bhaumik S., and Roy S., "Influence of alumina/MWCNT hybrid

nanoparticle additives on tribological properties of lubricants in turning operations," Friction,

pp. https://doi.org/10.1007/s40544-018-0199-5, 2018.

[44] Goindi G. and Sarkar P., "Dry machining: A step towards sustainable machining-challenges and

future directions," Journal of Cleaner Production, vol. 165, pp. 1557-1571, 2017.

[45] Minton T., Ghani S., Sammler F., Bateman F., Furstmann R., and Roeder P., "Temperature of

internally-cooled diamond-coated tools for dry-cutting titanium," International Journal of

Machine Tools and Manufacturing, vol. 75, pp. 27-35, 2013.

[46] Shokrani A., Dhokia V., and Newman S., "Environmentally conscious machining of difficult-

to-machine materials with regard to cutting fluids," International Journal of Machine Tools and

Manufacturing, vol. 83, pp. 83-101, 2012.

[47] Kishawy H., Li L., and El-Wahab A., "Prediction of chip flow direction during machining with

self-propelled rotary tools," International Journal of Machine Tools Manufacture, vol. 12, no.

46, pp. 1680-1688, 2006.

[48] Pervaiz S., "Investigation cooling and lubrication strategies for sustainable machining of

titanium alloys: impact on machinability and environmental performance," KTH Royal Institute

of Technology, Stockholm, 2014.

[49] Pervaiz S., Rashid A., Deiab I., and Nicolescu M., "An experimental investigation on effect of

minimum quantity cooling lubrication (MQCL) in machining titanium alloy (Ti6Al4V) ", The

International Journal of Advanced Manufacturing Technology, vol. 87, pp. 5-8, 2016.

[50] Kitagawa T., Kubo A., and Maekawa K., "Temperature and wear of cutting tools in high speed

machining of Inconel 718 and Ti-6Al-6V2Sn," Wear, vol. 202, pp. 142-148, 1997.

[51] Bagherzadeh A. and Budak E., "Investigation of machinability in turning of diffcult-to-cut

materials using a new cryogenic cooling approach," Tribology International, vol. 119, pp. 510-

520, 2018.

[52] Krajnik P., Pusavec F., and Rashid A., "Nanofluids: properties, applications and sustainability

aspects in materials processing technology," In Advances in Sustainable Manufacturing, pp.

107-113, 2011.

139

139

[53] Raza S. W., Pervaiz S., and Deiab I., "Tool wear patterns when turning of titanium alloy using

sustainable lubrication strategies," International Journal of Precision Engineering and

Manufacturing, vol. 15, no. 9, pp. 1979-1985, 2014.

[54] Krishna P., Srikant R., Padmini R., and Viswaditya, "Application of nanomaterials as

coolants/lubrications in machining," International Advanced Nanomaterials and Emerging

Engineering Technologies (ICANMEET), pp. 674-682, 2013.

[55] Behera B. C., Ghosh S., and Venkateswara P., "Modeling of cutting force in MQL machining

environment considering chip tool contact friction," Tribology International, vol. 117, pp. 283-

295, 2018.

[56] Kaynak Y., "Evaluation of machining performance in cryogenic machining of Inconel 718 and

comparison with dry and MQL machining," International Journal Advanced Manufacture

Technology, vol. 72, p. 919–933, 2014.

[57] Boswell B., Islam M., Davies J., Ginting Y., and Ong A., "A review identifying the effectiveness

of minimum quantity lubrication (MQL) during conventional machining," International Journal

of Advanced Manufacturing and Technology, vol. 92, pp. 321-340, 2017.

[58] Dureja J., Singh R., Singh P., Dogra M., Singh R. and Bhatti M., "Performance evaluation of

coated carbide tool in machining of stainless steel (AISI 202) under minimum quantity

lubrication (MQ), "International Journal of Precision Engineering and Manufacturing Green

Technology, vol. 2, pp. 123-129, 2015.

[59] Klock F. and Eisenblatter G., "Dry cutting," CIRP Annual Manufacturing Technology, vol. 46,

pp. 519-526, 1997.

[60] Tschatsch H. and Reichelt A., "Cutting fluid (coolants and lubricants)," In Applied Machining

Technology, pp. 349-352, 2009.

[61] Tai B., Stephenson D., Furness R., and Shih A., "Minimum quantity lubrication (MQL) in

automotive power train machining," Procedia CIRP, vol. 14, pp. 523-528, 2014.

[62] Leppert T., "Effect of cooling and lubrication conditions on surface topography and turning

process of C45 steel," International Journal of Machine Tools and Manufacture, vol. 51, no. 2,

pp. 120-126, 2011.

[63] Chinchanikar S. and Choudhury S. K., "Hard turning using HiPIMS-coated tools: Wear behavior

under dry and minimum quantity lubrication (MQL)," Measurement, vol. 55, pp. 536-548, 2014.

[64] Weinert K., Inasaki I., Sutherland J., and Wakabayashi T., "Dry machining and minimum

quantity lubrication," 54th General Assembly of CIRP, vol. 53, no. 2, pp. 511-537, 2004.

[65] Revankar G. D., Shetty R., Srinivas Rao S., and Gaitonde V. N., "Analysis of surface roughness

and hardness in titanium alloy machining with polycrystalline diamond tool under different

lubricating modes," Materials Research, vol. 17, no. 4, pp. 1010-1022, 2014.

140

140

[66] Deiab I., Raza S., and Pervaiz, S, "Analysis of lubrication strategies for sustainable machining

during turning of titanium Ti-6Al-4V alloy," CIRP, vol. 17, pp. 766-771, 2014.

[67] Dhar N., Kamruzzaman M., and Ahmed M., "Effect of minimum quantity lubrication (MQL)

on tool wear and surface roughness in turning AISI-4340 steel," Journal of Materials Processing

Technology, vol. 172, pp. 299-304, 2006.

[68] Chetan S. and Ghosh P. R., "Specific cutting energy modeling for turning nickel-based Nimonic

90 alloy under MQL condition," International Journal of Mechanical Sciences, vol. 146–147,

pp. 25-38, 2018.

[69] Khan A. and Maity K., "Influence of cutting speed and cooling method on the machinability of

commercially pure titanium (CP-Ti) grade II," Journal of Manufacturing Processes, vol. 31, p.

650–661, 2018.

[70] Ginting Y., Boswell B., Biswas W., and Islam W., "Advancing environmentally conscious

machining," Procedia CIRP, vol. 26, pp. 391-396, 2015.

[71] Eltaggaz A., Zawada P., Hegab H., Deiab I., and Kishawy H., "Coolant strategy influnce on tool

life and surface roughness when machining ADI," The International Journal of Advanced

Manufacturing Technology, vol. 94, no. 9-12, pp. 3875-3887, 2018.

[72] Sharma K., Tiwari A. K., and Dixit A., "Effects of minimum quantity lubrication (MQL) in

machining processes using conventional and nanofluid based cutting fluids: A comprehensive

review," Journal of Cleaner Production, vol. 127, pp. 1-18, 2016.

[73] Sarikaya M., Yilmaz V., and Gullu A., "Analysis of cutting parameters and cooling/lubrication

methods for sustainable machining in turning of Haynes 25 superalloy," Journal of Cleaner

Production, vol. 133, pp. 172-181, 2016.

[74] Sartori S., Ghiotti A., and Bruschi S., "Solid lubricant-assisted minimum quantity lubrication

and cooling strategies to improve Ti-6Al-4V machinability in finishing turning," Tribology

International, vol. 118, p. 287–294, 2018.

[75] Sadeghi M. H., Haddad M. J., Tawakoli T., and Emami M., "Minimal quantity lubrication-MQL

in grinding of Ti–6Al–4V titanium alloy," International Journal Advanced Manufacture

Technology, vol. 44, pp. 487-500, 2008.

[76] Stachurski W., Sawicki J., Wojcik R., and Nadolny K., "Influence of application of hybrid

MQL-CCA method of applying coolant during hob cutter sharpening on cutting blade surface

condition," Journal of Cleaner Production, vol. 171 , pp. 892-910, 2018.

[77] Revuru R., Posinasetti R., Ramana N., and Amrita V., "Application of cutting fluid in machining

of titanium alloys-a review," The International Journal of Advanced Manufacturing

Technology, vol. 91, p. 2477–2498 , 2017.

[78] Zhang S., Li J., and Wang Y., "Tool life and cutting forces in end milling Inconel 718 under dry

and minimum quantity cooling lubrication cutting conditions," Journal of Cleaner Production,

vol. 32, pp. 81-87, 2012.

141

141

[79] Boswell B. and Islam M. N., "Feasibility study of adopting minimal quantities of lubrication for

end milling aluminium," in Proceedings of the World Congress on Engineering, vol III, London,

U.K., July 4 - 6, 2012.

[80] Su Y., He N., Li L., Iqbal A., Xiao M., and Xu S., "Refrigerated cooling air cutting of difficult-

to-cut materials," International Journal of Machine Tools & Manufacture, vol. 47 ,pp. 927–933,

2007.

[81] Huang W.T., Wu D.H., and Chen J.T., "Robust design of using nanofluid/MQL in micro-

drilling," The International Journal of Advanced Manufacturing and Technology, vol. 85, pp.

2155-2161, 2016.

[82] Hajiyan M., Ebadil S., Mahmud S., Biglarbegian M., and Abdullah H., "Experimental

investigation of the effect of an external magnetic field on the thermal conductivity and viscosity

of Fe3O4–glycerol," Journal of Thermal Analysis and Calorimetry, vol. 133, pp. 1-14, 2018.

[83] Sharma A., Tiwari A., and Dixit A., "Progress of nanofluid application in machining: a review,"

Material Manufacturing Processes, vol. 30, no. 7, pp. 813-828, 2015.

[84] Saidur R., Leong K., and Mohammad H., "A review on applications and challenges of

nanofluids," Renewable and Sustainable Energy Reviews, vol. 15, pp. 1646-1668, 2011.

[85] Alberts M., Kalaitzidou K., and Melkote S., "An investigation of graphite nanoparticles as

lubricant in grinding," International Journal of Machine Tool and Manufacture, vol. 49, pp.

966-970, 2009.

[86] Lee P., Naam J., Li C., and Lee S., "An experimental study on micro-grinding process with

nanofluid minimum quantity lubrication (MQL)," International Journal Precission Engineering

Manufacturing, vol. 13, pp. 331-338, 2012.

[87] Lee P., Nam T., Li C., and Lee S., "Environmentally friendly nano-fluid minimum quantity

lubrication (MQL) meso-scale grinding process using nano-diamond particles, in manufacturing

automation (ICMA)," in Conference on IEEE, The University of Hong Kong, Hong Kong, 2010.

[88] Huang W., Liu W., and Wu D., "Investigation into lubrication in grinding processes using

MWCNTs nanofluids with ulrasonic-assisted dispersion," Journal of Cleaner Production, vol.

137, pp. 1553-1559, 2016.

[89] Pashmforoush F. and Bagherinia R. D., "Influence of water-based copper nanofluid on wheel

loading and surface roughness during grinding of Inconel 738 superalloy," Journal of Cleaner

Production, vol. 178, pp. 363-372, 2018.

[90] Molaie M., Akbari J., and Movahhedy M., "Ultrasonic assisted grinding process with minimum

quantity lubrication using oil-based nanofluids," Journal of Cleaner Production, vol. 129, pp.

212-222, 2016.

[91] Hsieh R., Liu H., and Yeh Y., "Nanofluids spray heat transfer enhancement," International

Journal of Heat and Mass transfer, vol. 94, pp. 104-118, 2016.

142

142

[92] Setti D., Sinha M., Ghosh S., and Rao P., "Performance evaluation of Ti-6Al-4V grinding using

chip formation and coefficient of friction under the influence of nanofluids," International

Journal of Tools Manufacture, vol. 88, pp. 237-248, 2015.

[93] Yuan S., Hou X., Wang L., and Chen B., "Experimental investigation on the compatibility of

nanoparticles with vegetable oils for nanofluid minimum quantity lubrication machining,"

Tribology Letters, 66: 106. https://doi.org/10.1007/s11249-018-1059-1, 2018.

[94] Nguyen T., Nguyen D., Howes P., Kwon P., and Park K.H., "Minimum quantity lubrication

(MQL) using vegetable oilwith nano-platelet solid lubricant in milling titanium alloy," in

Proceedings of the ASME 2015 International Manufacturing Science and Engineering

Conference MSEC2015, June 8-12, 2015, Charlotte, North Carolina, USA , 2015.

[95] Kim J., Woo J., and Lee W.,"Experimental characterization on micro-end milling of titanium

alloy using nanofluid minimum quantity lubrication with chilly gas," The International Journal

of Advanced Manufacturing Technology, vol. 91, no. 5-8, p. 2741–2749, 2017.

[96] Sahu N., Andhare A., and Raju R., "Evaluation of performance of nanofluid using multiwalled

carbon nanotubes for machining of Ti-6Al-4V," Machining Science and Technology An

International Journal, vol. 22, no. 3, pp. 476-492, 2018.

[97] Su Y., Gong L., Li B., Liu Z., and Chen D., "Performance evaluation of nanofluid MQL with

vegetable-based oil and ester fluids in turning," The International Journal of Advanced

Manufacturing Technology, vol. 83, pp. 2083-2089, 2016.

[98] Eltaggaz A., Hegab H., Deiab I., and Kishawy H., "Hybrid nano-fluid-minimum quantity

lubrication strategy for machining austempered ductile iron (ADI)," International Journal on

Interactive Design and Manufacturing (IJIDeM), vol. 12, no. 47, pp. 1-9, 2018.

[99] Nam J. and Lee S. W., "Machinability of titanium alloy (Ti-6Al-4V) in environmentally-friendly

micro-drilling process with nanofluid minimum quantity lubrication using nanodiamond

particles," International Journal of Precision Engineering and Manufacturing-Green

Technology, vol. 5, no. 1, pp. 29-35, 2018.

[100] Hegab H., Umer U., Soliman M., and Kishawy H., "Effects of nano-cutting fluids on tool

performance and chip morphology during machining Inconel 718," The International Journal

of Advanced Manufacturing Technology, vol. 96, no. 9-12, pp. 3449-3458, 2018.

[101] Hegab H., Darras B., and Kishawy H., "Sustainability assessment of machining with nano-

cutting fluids," Procedia Manufacturing, vol. 26, pp. 245-254, 2018.

[102] Hegab H. and Kishawy H., "Towards sustainable machining of Inconel 718 using nano-fluid

minimum quantity lubrication," Journal Manufacturing and Materials Processing, vol. 2, no. 3,

p. 50, 2018.

[103] Srikant R., Rao D., Subrahmanyam M., and Krishna P., "Applicability of cutting fluids with

nanoparticle inclusion as coolants in machining," Journal of Engineering Tribology , vol. 223,

no. 2, pp. 221-225, 2009.

143

143

[104] Kishawy H., "An experimental evaluation of cutting temperatures during high speed machining

of hardened D2 tool steel," Journal Machining Science and Technology, vol. 6, no. 1, pp. 67-

79, 2002.

[105] Haglund A., Kishawy H., and Rogers R., "An exploration of friction models for the chip-tool

interface using an arbitrary Lagrangian-Eulerian finite element model," Wear , vol. 265, no. 3-

4, p. 2008, 252-260.

[106] El-Wardany T., Kishawy H., and M. Elbestawi, "Surface integrity of die material in high speed

hard machining, part 1: micrographical analysis," Journal of Manufacturing Science and

Engineering, vol. 122, no. 4, p. 620–631, 2000.

[107] Sharma A. K., Singh R., Dixit A., and Tiwari A., "Characterization and experimental

investigation of Al2O3 nanoparticles based cutting fluid in turning of AISI 1040 steel under

minimum quantity lubrication (MQL)," Material Today, vol. 3, pp. 1899-1906, 2016.

[108] Sharma A. K., Singh R., Dixit A., and Tiwari A.,"Novel uses of alumina-MoS2 hybrid

nanoparticle enriched cutting fluid in hard turning of AISI 304 steel," Journal of Manufacturing

Processes, vol. 30, pp. 467-482, 2017.

[109] Zhang Y., Li C., Jia D., Zhang D., and X. Zhang, "Experimental evaluation of MoS2

nanoparticles in jet MQL grinding with different types of vegetable oil as base oil," Journal of

Cleaner Production, vol. 87, pp. 930-940, 2015.

[110] Alberts M., Kalaitzidou K., and Melkote S., "An investigation of graphite nanoplatelets as

lubricant in grinding," International Journal of Machine Tools and Manufacture, vol. 49, no.

12, p. 966–970 , 2009.

[111] Vasu V. and Reddy G., "Effect of minimum quantity lubrication with Al2O3 nanoparticles on

surface roughness, tool wear, temperature dissipation in machining Inconel 600 alloy," SAGE

Journals, vol. 225, no. 1, pp. 3-16, 2011.

[112] Wang Y., Li C., Zhang Y., Yang M., and Zhang X., "Experimental evaluation on tribological

performance of the wheel/workpiece interface in minimum quantity lubrication grinding with

different concentrations of Al2O3 nanofluids," Journal of Cleaner Production, vol. 142, pp.

3571-3583, 2017.

[113] Kishawy H., Mihaela D., and M. Elbestawi, "Effect of coolant stratgey on tool performance,

chip morphology and surface quality during high-speed machining of A356 aluminum alloy,"

International Journal of Machine Tools and Manufactur, vol. 2, no. 45, pp. 219-227, 2005.

[114] Montgomery D. C., Hand book, "Design and analysis of experiments, " 8th Edition, Arizona,

USA: John Wiley & Sons, Inc., 2012.

[115] Pervaiz S., "Numerical and experimental investigations of the machinability of Ti-6Al-4V,"

KTH Institute of Technology, School of Industrial Technology and Management, Stockholm,

Sweden, October 2015.

144

144

[116] Seker U. and Hasirci H., "Evaluation of machinability of austempered ductile iron in terms of

cutting forces and surface quality," Journal of Materials Processing Technology, vol. 173, pp.

260-268, 2006.

[117] Brandenberg K., "Successfully machining austempered ductile iron (ADI)," Applied Process

Inc., Michgen, USA, 2001.

[118] Aslantas K., "Preduction of cutting and surface roughness in turning of austempered ductile Iron

using fuzzy logic approach," Electronic Journal of Machine Technology, vol. 2, pp. 13-21, 2008.

[119] Shimizu K, Noguchi T, and Doi S. "Basic study on the erosive wear of austempered ductile

iron," Trans Am. Foundry Soc. 101:225– 229, 1993.

[120] Guesser W. L., Koda F., Martinez J., and Da Silva C. H., "Austempered ductile iron for gears,"

SAE International, vol. 36, 2012.

[121] Hayrynen K. L., "The production of austempered ductile iron (ADI)," in 2002 World Conference

on ADI, Livonla. MI, USA, 2002.

[122] Coromant Sandvik, "Sandvik Coromant," 2018. [Online]. Available:

https://www.sandvik.coromant.com/engb/products/Pages/productdetails.aspx?c=CNMG%201

2%2004%2016-MRR%20%20%201115, [Accessed 11 9 2018].

[123] KENNAMETAL, "Kennametal.com," 2018. [Online]. Available:

https://www.kennametal.com/en/products/20478624/47535256/63745063/63745065/6384030

3/63840318/55735598/100002321.html, [Accessed 27 9 2018].

[124] Hwang Y., Park H., Lee J., and Jung W., "Thermal conductivity and lubrication characteristics

of nanofluids," Current Applied Physics, vol. 6, pp. 67-71, 2006.

[125] Li X., Zhu D., Wang X., Wang N., and Gao J., "Thermal conductivity enhancement dependent

pH chemical surfactant for Cu-H2O nanofluids," Thremochim. Acta, vol. 469, no. 1-2, pp. 98-

103, 2008.

[126] Zhu X., Zhang C., Tang Y., Wang J., and Ren B., "Preparation and thermal conductivity of

suspension of graphite nanoparticles," Carbon, vol. 45, no. 1, pp. 226-228, 2007.

[127] Accu-Svenska, "Accu-Svenska AB, Minimum Quantity Lubrication Systems," [Online].

Available: http://www.en.accu-svenska.se/, [Accessed 4 9 2018].

[128] Kistler, measure, analyze, innovate. "Kistler, measure, analyze, innovate," 2018. [Online].

Available:https://www.kistler.com/en/applications/sensor-technology/cutting-force-

measurement/, [Accessed 4 9 2018].

[129] Grote K. and Atonsons E., "Handbook of Mechanical Engineering, " New York: Springer, 2009.

[130] MitutoyoCompany,[Online].Available:https://www.mitutoyo.co.jp/eng/products/gazoukogaku/

kenbikyo.html, [Accessed 5 9 2018].

145

145

[131] Power Sight Company, Summit Technology, [Online]. Available:

http://www.powersight.com/HTML/PRODUCTS/PS2500.html, [Accessed 9 September 2018].

[132] OMEGA Company, "User Guide HH374 Data logger thermometer," 2018. [Online]. Available:

Http://www.omega.com/manuals/manualpdf/M5055.pdf, [Accessed 12 10 2018].

[133] Venugopal K., Paul S., and Chattopadhyay A., "Tool wear in cryogenic turning of Ti-6Al-4V

alloy," Cryogenics, vol. 47, no. 1, pp. 12-18, 2007.

[134] Cordes S., Hübner F., and Schaarschmidt T., "Next generation high performance cutting by use

of carbon dioxide as cryogenics," Procedia CIRP, vol. 14, pp. 401-405, 2014.

[135] Stephenson D. A., Skerlos S. J., King A. S., and Superkar S. D., "Rough turning Inconel 750

with supercritical CO2-based minimum quantity lubrication," Journal Material Process

Technology, vol. 214, no. 3, pp. 673-680, 2014.

[136] Nguyen , Do I., and Kwon P., "A tribological study of vegetable oil enhanced by nano-platelets

and implication in MQL machining," International Journal of Precission Engineering

Manufacture , vol. 13, no. 7, p. 1077–1083, 2012.

[137] Bhowmick S. and Alpas A., "Minimum quantity lubrication drilling of aluminim-silicon alloys

in water using diamond-like carbon coated drills," International Journal Machine Tool and

Manufacture, vol. 48, no. 12, pp. 1429-1443, 2008.

[138] Bhowmick S. and Alpas , "The role of diamond-like carbon coated drill on minimum quantity

lubrication drilling of magnesium alloys," Surface Coat Technology, vol. 205, no. 23, pp. 5302-

5311, 2011.

[139] Klocke F. and Klöpper C., "Machining of ADI," [Online]. https://www.google.ca/url?sa,

[Accessed 2108].

[140] Kim S. J., Bang I. C,. Buongiorno J., and Hu W., "Effects of nanoparticle depostion on surface

wettability influncing boiling heat transfer in nanofluids," Application Physicas Letters, vol. 89,

no. 15, 2006.

[142] Wang X. Q. and Mujumdar A., "Heat transfer characteristics of nanofluids: a review,"

International Journal of Thermo Science, vol. 46, no. 1, pp. 1-19, 2007.

[143] Khandekar. S., Ravi Sankar, Agnihotri V., Ramkumar J., "Nano-cutting fluid for enhancement

of metal cutting performance," Materials and Manufacturing Processes, vol. 27, no. 9, pp. 963-

967, 2012.

[144] Hsieh S., Liu H. and Yeh Y., "Nanofluids spray heat transfer enhancement," International

Journal of Heat Mass Transfer, vol. 94, pp. 104-118, 2016.

[145] Rahmati B., Sarhan A. and Sayuti M., "Morphology of surface generated by end milling Al

6061-T6 using molybdenum disulfide (MoS2) nanolubrication in end milling machining,"

Journal of Cleaner Production, vol. 66, pp. 685-691, 2014.

146

146

[146] Phadke M., "Quality engineering using design of experiments, quality control, robust design,

and the Taguchi method, " Los Angeles: Wadsworth, 1998.

[147] Öktem H., Erzurumlu T., and Kurtaran H., "Application of response surface methodology in the

optimization of cutting conditions for surface roughness," Journal of Material Process

Technology, vol. 170, no. 1, pp. 11-16, 2005.

[148] Odelros, S , "Tool wear in titanium machining," , PhD Thesis, Uppsala University, 2012.

[149] Byers J., Hand book," Metalworking fluids," CRC Press.

[150] De Oliveira V. and Dos Santos M., "Analysis of the cutting tool's wear behavior for turning

machining of austempered ductile," in 19th International Congress of Mechanical Engineering,

Brasilia, 2007.

[151] Mehta S., Chauhan K.P., and Kanagaraj S., "Modeling of thermal conductivity of nanofluids by

modifying Maxwell’s equation using cell model approach, " Journal of Nanoparticle Res. 13(7),

2791–2798., 2011.

[152] Loos M., "Carbon nanotube reinforced composites, " CNT Polymer Science and Technology,

Elsevier, London, 2014.

[153] Eltaggaz A., Hegab H., Deiab I., and Kishawy H., "On using nano-cutting fluid when machining

austempered ductile iron (ADI)," The 26th Canadian Congress of Applied Mechnaics, Victoria,

Canada, 2017.

[154] Khandekar S., Sankar M. R., Agnihotri V., and Ramkumar J.,"Nano-cutting fluid for

enhancement of metal cutting performance, " Materials and Manufacturing Processes, 27(9),

963-967, 2012.

[155] Sharma A.K., Tiwari A. R., and Dixit A.R., "Progress of nanofluid application in machining: a

review, " Materials and Manufacturing Processes, 30(7): p. 813-828, 2015.

[156] Rajmohan T., Sathishkumar S., and Palanikumar K., "Effect of a nanoparticle-filled lubricant in

turning of AISI 316L stainless steel (SS)," Particulate Science and Technology, 35(2): p. 201-

208, 2017.

[157] Singh R. K., Sharma, Dixit A. K., Tiwari A. R., Pramanik A. K., and Mandal A, "Performance

evaluation of alumina-graphene hybrid nano-cutting fluid in hard turning, " Journal of Cleaner

Production, 162, 830-845, 2017.

[158] Bakalova T., Svobodová L., Rosická P., Borůvková K., Voleský L.,and Louda P., "The

application potential of SiO2, TiO2 or Ag nanoparticles as fillers in machining process fluids,"

Journal of Cleaner Production, 142, 2237-2243, 2017.

[159] Liu G., Li C., Zhang Y., Yang M., Jia D., Zhang X., and Zhai H., "Process parameter

optimization and experimental evaluation for nanofluid MQL in grinding Ti-6Al-4V based on

grey relational analysis," Materials and Manufacturing Processes, 1-14, 2017.

147

147

[160] Kandile N.G., and Harding D.R., "Nanotechnology and performance development of cutting

fluids," Surfactants in Tribology, 2017.

[161] Lee K., Hwang Y., Cheong S., Choi Y., Kwon L., Lee J., and Kim S. H., "Understanding the

role of nanoparticles in nano-oil lubrication, " Tribology Letters, 35(2), 127-131, 2009.

[162] Kaynak Y., "Evaluation of machining performance in cryogenic machining of Inconel 718 and

comparison with dry and MQL machining," International Journal Advanced Manufacture

Technology, vol. 72, p. 919–933, 2014.

[163] Boswell B., Islam M., Davies J., Ginting Y. and Ong A., "A review identifying the effectiveness

of minimum quantity lubrication (MQL) during conventional machining," International Journal

of Advanced Manufacturing and Technology, vol. 92, pp. 321-340, 2017.

[164] Dureja J., Singh R., Singh P., Dogra M., Singh R., and Bhatti M., "Performance evaluation of

coated carbide tool in machining of stainless steel (AISI 202) under minimum quantity

lubrication (MQ)," International Journal of Precision Engineering and Manufacturing Green

Technology, vol. 2, pp. 123-129, 2015.

[165] Klock F. and Eisenblatter G., "Dry cutting," CIRP Annual Manufacturing Technology, vol. 46,

pp. 519-526, 1997.

[166] Tschatsch H. and Reichelt A., "Cutting fluid (coolants and lubricants)," In Applied Machining

Technology, pp. 349-352, 2009.

[167] Tai B., Stephenson D., Furness R. and Shih A., "Minimum quantity lubrication (MQL) in

automotive power train machining," Procedia CIRP, vol. 14, pp. 523-528, 2014.

[168] Weinert K., Inasaki I., and Wakabayashi T., "Dry machining and minimum quantity

lubrication," CIRP Annual Manufacture Technology, vol. 53, pp. 511-537, 2004.

[169] Dhar N., Kamruzzaman M., and Ahmed M., "Effect of minimum quantity lubrication (MQL)

on tool wear and surface roughness in turning AISI-4340 steel," Journal of Materials Processing

Technology, vol. 172, pp. 299-304, 2006.

[170] Ginting Y., Boswell B., Biswas W., and Islam W., "Advancing environmentally conscious

machining," Procedia CIRP, vol. 26, pp. 391-396, 2015.

[171] Sharma K., Tiwari A. K. ,and Dixit A., "Effects of minimum quantity lubrication (MQL) in

machining processes using conventional and nanofluid based cutting fluids: A comprehensive

review," Journal of Cleaner Production, vol. 127, pp. 1-18, 2016.

[172] Sarikaya M., Yilmaz V., and Gullu A., "Analysis of cutting parameters and cooling/lubrication

methods for sustainable machining in turning of Haynes 25 superalloy," Journal of Cleaner

Production, vol. 133, pp. 172-181, 2016.

[173] Huang W.T., Wu D.H., and Chen J.T., "Robust design of using nanofluid/MQL in micro-

drilling," The International Journal of Advanced Manufacturing and Technology, vol. 85, pp.

2155-2161, 2016.

148

148

[174] Hajiyan M., Ebadil S., Mahmud S., Biglarbegian M., and Abdullah H., "Experimental

investigation of the effect of an external magnetic field on the thermal conductivity and viscosity

of Fe3O4–glycerol," Journal of Thermal Analysis and Calorimetry, vol. 133, pp. 1-14, 2018.

[175] Saidur R., Leong K., and Mohammad H., "A review on applications and challenges of

nanofluids," Renewable and Sustainable Energy Reviews, vol. 15, pp. 1646-1668, 2011.

[176] Alberts M., Kalaitzidou K., and Melkote S., "An investigation of graphite nanoparticles as

lubricant in grinding," International Journal of Machine Tool and Manufacture, vol. 49, pp.

966-970, 2009.

[177] Lee P., Naam J., Li C., and Lee S., "An experimental study on micro-grinding process with

nanofluid minimum quantity lubrication (MQL)," International Journal Precission Engineering

Manufacturing, vol. 13, pp. 331-338, 2012.

[178] Hsieh R., Liu H. and Yeh Y., "Nanofluids spray heat transfer enhancement," International

Journal of Heat and Mass transfer, vol. 94, pp. 104-118, 2016.

[179] Su Y., Gong L., Li B., Liu Z. and Chen D., "Performance evaluation of nanofluid MQL with

vegetable-based oil and ester fluids in turning," The International Journal of Advanced

Manufacturing Technology, vol. 83, pp. 2083-2089, 2016.

[180] Sharma A., Singh R., Dixit A. and Tiwari A., "Novel uses of alumina-MoS2 hybrid nanoparticle

enriched cutting fluid in hard turning of AISI 304 steel," Journal of Manufacturing Processes,

vol. 30, pp. 467-482, 2017.

[181] Zhang Y., Li C., Jia D., Zhang D. and Zhang X., "Experimental evaluation of MoS2

nanoparticles in jet MQL grinding with different types of vegetable oil as base oil," Journal of

Cleaner Production, vol. 87, pp. 930-940, 2015.

[182] Vasu V. and Reddy G., "Effect of minimum quantity lubrication with Al2O3 nanoparticles on

surface roughness, tool wear, temperature dissipation in machining Inconel 600 alloy," SAGE

Journals, vol. 225, no. 1, pp. 3-16, 2011.

[183] Wang Y., Li C., Zhang Y., Yang M. and Zhang X., "Experimental evaluation on tribological

performance of the wheel/workpiece interface in minimum quantity lubrication grinding with

different concentrations of Al2O3 nanofluids," Journal of Cleaner Production, vol. 142, pp.

3571-3583, 2017.

[184] Yu F., Chen Y., Liang X., Xu J., Lee C., Liang Q., Tao P., and Deng T., "Dispersion stability

of thermal nanofluid," Progress in Natural Science: Materials International , vol. 27, p. 531–

542, 2017.

[185] Hwang Y., "Thermal conductivity and lubrication characteristics of nanofluids," Current

Applied Physics, vol. 6, pp. 226-228, 2006.

[186] Zhu H., Zhang C., Tang Y., Wang J., Ren B., and Yin Y., "Preparation and thermal conductivity

of suspensions of graphite nanopaticles," Carbon, vol. 45, no. 1, pp. 226-228, 2007.

149

149

[187] Hwang Y., Lee J., Lee C., Jung Y., Cheong S., Lee C., Ku B. and Jang S., "Stability and thermal

conductivity characteristics of nanofluids," Thermochim Acta, vol. 455, no. 1, pp. 70-74, 2007.

[188] Uuniversity of Guelph, Health and Safety, "University of Guelph Hazardous Wsate," 2017.

[Online]. Available: https://www.uoguelph.ca/hr/hr-services-environmental-health-safety-

programs-laboratory-safety/hazardous-waste. [Accessed 5 05 2017].

[189] Ghaednia H. and Jackson R., "The effect of nanopartricles on the real area of contact, friction,

and wear," Journal of Tribology, vol. 4, no. 135, 2013.

[190]

Trent E. M., "Metal cutting and the tribology of seizure: I Seizure in metal cutting," Wear, vol.

128, pp. 29-45, 1988.

[191] Joshi, K., Venkatachalam, A. and Jawahir I. , "A new methodology for transforming 3R concept

into 6R concept for improved product sustainability," International IV Global Conference on

Sustainable Product Development and Life Cycle Enginerring, 2006.

[192] Eltaggaz A. and Deiab I., "Comparison of between direct and peck drilling for large aspect ratio

in Ti-6Al-4V alloy," The International Journal of Advanced Manufacturing Technology, 1-9.

2019.

[193] Ezugwu E. O.,"Key improvements in the machining of difficult-to-cut aerospace super-

alloys," International Journal of Machine Tools and Manufacture, 45(12-13), 1353-1367. 2005

[194] Attanasio A., Gelfi M., Giardini C., and Remino C., "Minimal quantity lubrication in turning:

effect on tool wear," Wear, 260(3), 333-338, 2006.

[195] Weinert K., Inasaki I., Sutherland J. W., and Wakabayashi T., "Dry machining and minimum

quantity lubrication," CIRP annals, 53(2), 511-537, 2004.

[196] Hegab H., Kishawy H. A., Umer U., and Mohany A.,"A model for machining with nano-

additives based minimum quantity lubrication,” The International Journal of Advanced

Manufacturing Technology, 1-16., 2019.

[197] Chowdhury S. A., Islam M. N., and Boswell B.,"Effectiveness of using CFD for comparing tool

cooling methods," In Proceedings of the World Congress on Engineering, vol. 2, 2014.

[198] Balan A. S., Kullarwar T., Vijayaraghavan L., and Krishnamurthy R., "Computational fluid

dynamics analysis of MQL spray parameters and its influence on superalloy

grinding," Machining Science and Technology, 21(4), 603-616., 2017.

[199] Klöpper C. F. and Klocke F. "Untersuchungen zur Zerspanbarkeit von austenitisch-ferritischem

Gusseisen mit Kugelgraphit (ADI), " No. RWTH-CONV-124078, Lehrstuhl für Technologie

der Fertigungsverfahren., 2006.

[200] Arft M. and Klocke F.,"High performance turning of austempered ductile iron (ADI) with

adapted cutting inserts," Procedia CIRP, 8, 129-134., 2103.

150

150

[201] Brodmann, M., "Schädigungsmodell für schlagartige Beanspruchung metallischer Werkstoffe,

" Verlag nicht ermittelbar, 2001.

[202] Klocke F., and Klopper C., "Machining of ADI," Production Engineering, Aachen, Germany,

2013.

[203] Pervaiz S., Deiab I., Wahba E. M., Rashid A., and Nicolescu M., "A coupled FE and CFD

approach to predict the cutting tool temperature profile in machining, " Procedia CIRP, 17, 750-

754, 2014.

[204] Systèmes, D.J.S.C.P., RI, USA, Abaqus analysis user’s manual. 2007.

151

151

APPENDIX I:

Assessment study:

- Sustainable Machining Metrics and indictors

-CO2 Emission rate

152

152

- Proposed algorithm of sustainability assessment of machining processes

- Sustainability assessment of machining processes

153

153

Case No. 2:

154

154

Case No. 3:

155

155

Case No.4:

156

156

APPENDIX II

Energy model