You are on page 1of 114

ARAB ACADEMY FOR SCIENCE, TECHNOLOGY

AND MARITIME TRANSPORT


College of Engineering and Technology
Department of Construction and Building Engineering

UNCERTAINTIES IN THERMO-MECHANICAL
CHARACTERIZATION OF NI-TI SHAPE MEMORY
ALLOYS FOR CIVIL ENGINEERING APPLICATION

By
Ossama Alaa Mekawy
In metrology, measurement uncertainty is the expression of the statistical dispersion of
the values attributed to a measured quantity. All measurements are subject to uncertainty
and a measurement result is complete only when it is accompanied by a statement of the
associated uncertainty, such as the standard deviation.
A thesis submitted to AASTMT in partial
Fulfillment of the requirements for the award of the degree of
MASTER OF SCIENCE
in
CONSTRUCTION AND BUILDING ENGINEERING
In the world of metrology, there is no way to achieve 100% accuracy. It is only a goal. For
every measurement, even the most careful and precise, there is always a margin of doubt
or uncertainty.
Supervisors
Prof. Nabil H. El-Ashkar Prof. Alaa Mahmoud Morsy
Professor of Structural Engineering Professor of Structural Engineering
Construction and Building Engineering department
College of Engineering and Technology
Arab Academy for Science, Technology and Maritime Transport
Alexandria
Declaration

I certify that all the material in this thesis that is not my own work has been identified,
and

that no material is included for which a degree has previously been conferred on me.

The contents of this thesis reflect my own personal views, and are not necessarily
endorsed by the University.

Name: Ossama Alaa Mekawy

Signature: .............................................................................................

Date: .............................................................................................

II
We certify that we have read the present work and that, in our opinion, it is fully adequate in scope and
quality as a thesis towards the partial fulfillment of the Master's Degree requirements in

Construction and Building Engineering


Ossama Alaa Mekawy from
College of Engineering and Technology
AASTMT
31 December 2023
Supervisor(s):
Name: Prof. Nabil H. El-Ashkar
Position: Professor of Structural Engineering, Construction and Building Department, College of
Engineering and Technology, Arab Academy for Science, Technology and Maritime
Transport, Alexandria

Signature:

Name: Prof. Alaa Mahmoud Mohamed Morsy


Position: Professor of Structural Engineering, Construction and Building Department, College of
Engineering and Technology, Arab Academy for Science, Technology and Maritime
Transport, Alexandria
Signature:
Examiners:
Name: Prof. Khandaker M. Anwar Hossain
Position: Professor of Structures and Materials, Department of Civil Engineering, Faculty of
Engineering & Architectural Science, Toronto Metropolitan University, Canada

Signature:

Name: Prof. Amr Amin Elhefnawy


Position: Professor of Building Materials and Quality Control
Housing and Building National Research Center
Signature:
Name: Prof. Nabil H. El-Ashkar
Position: Professor of Structural Engineering, Construction and Building Department, College of
Engineering and Technology, Arab Academy for Science, Technology and Maritime
Transport, Alexandria
Signature:

III
Acknowledgement
First and foremost, piers and thanks to Allah, the Most Gracious, the Most Merciful, and
peace are upon His Prophet. I would like to express my sincere appreciation and gratitude to
my supervisors, Professors Dr. Nabil H. El-Ashkar and Dr. Alaa Morsy for their continual
support, encouragement, and guidance throughout my study. Specifically, for the teachings
they provided, the time they invested even during a chaotic schedule, and their heartfelt
sincerity to help me achieve my goals. Thanks, professors, for your valuable guidance,
generous help, great support, critical comments, and encouragement throughout my
research.

I would especially like to express my sincere appreciation and gratitude to Prof. Khandaker
M. Anwar Hossain for his unwavering support and guidance. From the inception of the
experimental program to the culmination of this thesis, his consistent assistance, weekly
meetings, and mentorship have been invaluable to the success of my study.

Thanks also are due to the staff of the R.C. laboratory, college of engineering and
technology, Arab academy for Science, technology and Marine Transport, and, Al Ezz
Dekheila Steel (EZDK) in Alexandria, Egypt, for their generous support during the
experimental work of this study, which played a pivotal role in the success of this research
endeavor.
I am so grateful to my dear friend Ossama Nabil and my uncle Shabban Mahmoud who
have been supportive and encouraging through my Master journey.

Finally, I couldn’t have done this without the support, love and prayers of my beloved
family, my parents, my brothers, my sisters, my wife, and last but not least my little
daughter and my upcoming child inshallah.

Ossama Alaa Mekawy

IV
DEDICATION

To

My loving parents, Alaa Mekawy & Hanan Shabban

My brothers, Eslam, Ismail, & Ali

My sisters, Asmaa, & Oswa

My wife, Mai and my daughter, Asawer and my upcoming

child inshallah.

To the brave people of Palestine

Ossama Alaa Mekawy

V
PUBLISHED RESEARCHES

1- Unraveling the Impact of Temperature Variation on Shape


Memory Alloys in Reinforced Concrete Structures.

Conference: the ELEVENTH ALEXANDRIA International Conference


on Structural, Geotechnical Engineering and Management “AICSGE-11”

2- Uncertainties in Thermo-Mechanical Characterization of Ni-Ti


Shape Memory Alloys for Civil Engineering Application

Journal: Construction and Building Materials (under review)

Organization: Elsevier.

VI
Abstract
Shape memory alloys (SMAs) have gained increasing attention in civil engineering
research due to their remarkable properties like superelasticity, shape memory effect, high
ductility, and corrosion resistance. However, SMAs exhibit complex material behaviour
influenced by various parameters such as chemical composition, thermo-mechanical
treatment, and more. Even when two material specimens have identical composition and
heat treatment, they can behave dissimilarly under tension tests, resulting in varying
mechanical properties, failure shapes, and elongation post-fracture. These disparities
persist even when testing is conducted under the standards of ASTM, highlighting a crucial
challenge. It means that materials factor should be defined and is expected to be higher than that of
reinforcing steel.

In this study, we conducted a comprehensive investigation to uncover the underlying


reasons for these discrepancies. Our focus was on understanding the intricate relationship
between established standards and their procedural intricacies and how they contribute to
Reasons for these inconsistencies. We explored several key factors, including gauge length variation,
disparities.
gauge length location, strain rates, stress rates, temperature variation and gripping pressure.
The results from this study showed the contribution of every studied parameter on this
variation, shedding light on the intricate behaviour of SMAs.

It means that One of the most important of our findings reveal that increasing gripping pressure, a
ASTM should
be modified. parameter not previously incorporated into standardized testing protocols, significantly
reduces slippage, providing more accurate values for the modulus of elasticity, precise
That is true for determination of critical stresses, and minimized differences between mechanical
any materials
not only SMA. properties obtained from extensometer and crosshead readings. Moreover, higher gripping
pressure leads to extended plastic deformation regions with necking formation, indicating
ductile behaviour, and substantial reductions in area (40-50%). Conversely, lower gripping
pressure results in less ductile behaviour with minimal post-fracture plastic deformation
and smaller reductions in area (20-30%). These insights underscore the imperative need to
account for a multitude of parameters during tensioning tests for SMAs, to secure accurate
and precise assessments of their mechanical properties.

VII
Table of Content
Acknowledgement............................................................................................................................. IV
PUBLISHED RESEARCHES ..................................................................................................... VI
Abstract ............................................................................................................................................ VII
Table of Content .............................................................................................................................. VIII
List of Tables..................................................................................................................................... XI
List of Figures .................................................................................................................................. XII
CHAPTER 1....................................................................................................................................... 2
1 INTRODUCTION................................................................................................................. 2
1.1. General ................................................................................................................................2
1.2. Problem Statement ..............................................................................................................3
1.3. Research Objective..............................................................................................................6
1.4. Research Methodology........................................................................................................6
1.5. Organization of Thesis ........................................................................................................7
CHAPTER 2....................................................................................................................................... 9
2 LITERATURE REVIEW...................................................................................................... 9
2.1. Introduction to Shape Memory Alloy (SMA) .....................................................................9
2.2. Phase Transformation in Shape Memory Alloys ..............................................................10
2.3. Shape Memory Effect (SME) ............................................................................................12
2.4. Superelasticity ...................................................................................................................14
2.5. Effect of Ambient Temperatures on the Thermo-Mechanical Properties of SMA ...........16
2.6. Alloy Composition of SMA ..............................................................................................17
2.6.1 Nickel titanium (Ni-Ti) based alloys ......................................................................17
2.6.2 Copper (Cu) based alloys ........................................................................................17
2.6.3 Iron (Fe) based alloys ..............................................................................................18
2.6.4 General discussion on the types of SMA ...............................................................18
2.7. Thermo-Mechanical Properties of SMAs..........................................................................20
2.7.1 Effect of strain localization phenomena ................................................................20
2.7.2 Stress-strain behaviour of SMAs under repeated cyclic loading ........................24
2.7.3 Effect of strain rate 28
2.7.4 Effect of stress rate 29
2.7.5 Effect of thermal hysteresis band on recovery stress ...........................................31
2.8. Research Gaps ...................................................................................................................32
CHAPTER 3..................................................................................................................................... 34
3 EXPERIMENTAL PROGRAM ......................................................................................... 34
3.1. Materials and Specimens ...................................................................................................34
3.2. Experimental Setup ...........................................................................................................35
VIII
3.3. Experimental Objectives ...................................................................................................36
3.4. Study Parameters ...............................................................................................................37
CHAPTER 4..................................................................................................................................... 43
4 RESULTS AND DISCUSSION ......................................................................................... 43
4.1 Mechanical Parameters .....................................................................................................43
4.1.1 Influence of gauge length variations ......................................................................43
4.1.1.1 Modulus of elasticity .............................................................................. 43

4.1.1.2 Stresses ................................................................................................... 46

4.1.1.3 Strains ..................................................................................................... 47

4.1.2 Influence of gauge length location .........................................................................49


4.1.2.1 Modulus of elasticity .............................................................................. 49

4.1.2.2 Stresses ................................................................................................... 51

4.1.2.3 Strains ..................................................................................................... 52

4.1.3 Influence of strain rate increase .............................................................................54


4.1.3.1 Modulus of elasticity (EA) ..................................................................... 54

4.1.3.2 Stresses ................................................................................................... 56

4.1.3.3 Strains ..................................................................................................... 57

4.1.4 Influence of stress rate transition...........................................................................58


4.1.4.1 Modulus of elasticity .............................................................................. 58

4.1.4.2 Stresses ................................................................................................... 60

4.1.4.3 Strains ..................................................................................................... 61

4.1.5 Influence of gripping pressure ...............................................................................63


4.1.5.1 Modulus of Elasticity ............................................................................. 63

4.1.5.2 Stresses ................................................................................................... 65

4.1.5.3 Strains ..................................................................................................... 66

4.1.6 Influence of temperature variation ........................................................................68


4.1.6.1 General Discussion................................................................................. 68

4.1.6.2 Modulus of elasticity .............................................................................. 70

4.1.6.3 Stresses ................................................................................................... 72

4.1.6.4 Strains ..................................................................................................... 73

IX
4.2 Post-Fracture Elongation Characteristics ..........................................................................74
4.2.1 Influence of gauge length variations ......................................................................75
4.2.2 Influence of gauge length location .........................................................................76
4.2.3 Influence of strain rate increase .............................................................................76
4.2.4 Influence of stress rate transition...........................................................................77
4.2.5 Influence of gripping pressure increase ................................................................78
4.2.6 Influence of temperature variation ........................................................................79
4.3 Temperature Fluctuations Assessment ..............................................................................81
4.3.1 Stress Temperature Curves ....................................................................................81
4.3.2 Rise in Temperature Strain Curves .......................................................................82
4.4 Analysis of failure characteristics .....................................................................................85
CHAPTER 5..................................................................................................................................... 92
5 SUMMARY AND CONCLUSIONS.................................................................................. 92
REFERENCES ................................................................................................................................. 95
References ........................................................................................................................................ 96

X
List of Tables
Table 2-1 Atomic composition for selected SMA’s [30]. ................................................... 18
Table 2-2 Mechanical properties for selected SMA’s [30]. ................................................ 19
Table 2-3 Martensite and austenite start and finish temperatures [45]................................ 31
Table 3-1 Chemical Composition (wt %) From Supplier. .................................................. 34
Table 3-2 Transformation Temperatures (°C) From Supplier. ............................................ 35
Table 3-3 Mechanical Properties from Supplier.................................................................. 35
Table 3-4 Experimental Program. ....................................................................................... 40
Table 4-1 Modulus of Elasticity for gauge length variation................................................ 45
Table 4-2 variation for stress values with gauge length from extensometer and crosshead
readings................................................................................................................................ 46
Table 4-3 Strain values with variation of gauge lengths from extensometer and crosshead
readings................................................................................................................................ 48
Table 4-4 Modulus of elasticity for gauge length variation. ............................................... 50
Table 4-5 Stress values with variation of gauge location. ................................................... 51
Table 4-6 Variation for strain values with gauge location from extensometer and crosshead
readings................................................................................................................................ 53
Table 4-7 Modulus of Elasticity for strain rate variation. ................................................... 55
Table 4-8 Variation of stress values with strain rate from extensometer and crosshead
readings................................................................................................................................ 56
Table 4-9 Variation of strains with strain rate from extensometer and crosshead readings.
............................................................................................................................................. 57
Table 4-10 Modulus of Elasticity for stress rate variation. ................................................. 59
Table 4-11 Variation of stress with stress rate from extensometer and crosshead readings.
............................................................................................................................................. 60
Table 4-12 Variation for strains with stress rate from extensometer and crosshead readings.
............................................................................................................................................. 62
Table 4-13 Variation of modulus of elasticity with gripping pressure. ............................... 64
Table 4-14 Variation of stress with gripping pressure from (a) extensometer and (b)
crosshead readings y. ........................................................................................................... 65
Table 4-15 Variation of strain with gripping pressure from extensometer and crosshead
readings................................................................................................................................ 67
Table 4-16 Modulus of Elasticity variation between group (2) and group (11). ................. 71

XI
List of Figures
Figure 2-1 Thermo-mechanical characteristics of SMA [18].............................................. 10
Figure 2-2 Illustration of SMA deformation phases [30]. ................................................... 11
Figure 2-3 Shape Memory Effect of SMA [18]. ................................................................. 13
Figure 2-4 Superelasticity behaviour of SMA [18]. ............................................................ 14
Figure 2-5 Effect of ambient temperatures on the thermo-mechanical properties of SMA
[32]. ..................................................................................................................................... 15
Figure 2-6 Measured nominal stress – elongation response obtained from Test 1 and 2 [14].
............................................................................................................................................. 21
Figure 2-7 Measured nominal stress – crosshead displacement response in Test 2 [14]. ... 22
Figure 2-8 Strain contours showing the formation and evolution of TBs along the length in
Test 2 [14]............................................................................................................................ 23
Figure 2-9 Strain distribution along the length of the specimen at various stages from
initiation to complete propagation of TBs in Test 2 [14]. ................................................... 24
Figure 2-10 Measured nominal stress – elongation response of a virgin NiTiNb wire
Specimen in Test 3 [14]. ...................................................................................................... 25
Figure 2-11 Strain contour of Test 3 specimen showing evolution of TBs during loading
and unloading for strain range of 0−9% [14]. ..................................................................... 26
Figure 2-12 Nominal stress – elongation response measured at (a) TB-2b; (b) TB-3; (c)
fracture location (FrL); (d) at location FL-1; (region between TB-1 and TB-3) [14]. ........ 27
Figure 2-13 Comparison of unloading control methods for pseudoelastic testing: (a) stress-
strain response, (b) specimen surface temperature response [18]. ...................................... 30
Figure 3-1 (a) Test set-up for tensile tests, (b) close-up view of the specimen and (c) Gauge
marks. .................................................................................................................................. 39
Figure 3-2 (a) groups with constant temperature are exposed to RT and (b) group 11 with
temperature variation in the gauge length exposed to ice pack. .......................................... 39
Figure 3-3 (a) Stress vs. extensometer strain and stress vs. crosshead strain curves, (b) and
(c) stress-strain curves depicting tangent lines to determine the mechanical properties. .... 41
Figure 4-1 Variation of EA and EM with gauge length from (a) extensometer and (b)
crosshead readings. .............................................................................................................. 46
Figure 4-2 Variation for stress values with gauge lengths (a) from extensometer and (b)
crosshead readings. .............................................................................................................. 47

XII
Figure 4-3 Variation for strain values with gauge lengths (a) extensometer and (b)
crosshead readings. .............................................................................................................. 48
Figure 4-4 Variation of EA and EM with gauge length location from (a) extensometer and
(b) crosshead readings. ........................................................................................................ 51
Figure 4-5 (Variation for Stress with gauge location from (a) extensometer and (b)
crosshead readings. .............................................................................................................. 52
Figure 4-6 Variation for strain values with gauge location from (a) extensometer and (b)
crosshead readings. .............................................................................................................. 53
Figure 4-7 Variation of EA and EM with strain rate from (a) extensometer and (b)
crosshead readings. .............................................................................................................. 56
Figure 4-8 Variation of stress values with strain rate from (a) extensometer and (b)
crosshead readings. .............................................................................................................. 57
Figure 4-9 Variation of strains with strain rate from (a) extensometer and (b) crosshead
readings................................................................................................................................ 58
Figure 4-10 Variation of EA and EM with stress rate from (a) extensometer and (b)
crosshead readings. .............................................................................................................. 60
Figure 4-11 Variation of stress with stress rate from (a) extensometer and (b) crosshead
readings................................................................................................................................ 61
Figure 4-12 Variation of strains with stress rate from (a) extensometer and (b) crosshead
readings................................................................................................................................ 62
Figure 4-13 Variation of (EA and EM with gripping pressure from (a) extensometer and
(b) crosshead readings. ........................................................................................................ 65
Figure 4-14 Variation of stress with gripping pressure from (a) extensometer and (b)
crosshead readings. .............................................................................................................. 66
Figure 4-15 Variation of strain with gripping pressure from (a) extensometer and (b)
crosshead readings. .............................................................................................................. 67
Figure 4-16 Stress vs. both extensometer strain and crosshead strain curves for (a) Group
(2) and (b) Group (11). ........................................................................................................ 69
Figure 4-17 Stress vs strain curves for both Group (2) and Group (11) from (a)
extensometer and (b) crosshead readings. ........................................................................... 70
Figure 4-18 Variation of EA and EM between group 2 and group 11 from extensometer
and crosshead readings. ....................................................................................................... 72
Figure 4-19 Variation for stress values between Group 2 and Group 11 from extensometer
(Ex) and crosshead (CH) readings. ...................................................................................... 73

XIII
Figure 4-20 Variation for strain values between Group 2 and Group 11 from extensometer
(Ex) and crosshead (CH) readings. ...................................................................................... 74
Figure 4-21 Influence of gauge length on (a) recovery strain and (b) retained strain
distribution along the specimen's entire length. .................................................................. 75
Figure 4-22 Influence of gauge length location on (a) recovery strain and (b) retained strain
distribution along the specimen's entire length. .................................................................. 76
Figure 4-23 Influence of strain rate increase on (a) recovery strain and (b) retained strain
distribution along the specimen's entire length. .................................................................. 77
Figure 4-24 Influence of stress rate transition on (a) recovery strain and (b) retained strain
distribution along the specimen's entire length. .................................................................. 78
Figure 4-25 Influence of gripping pressure increase on (a) recovery strain and (b) retained
strain distribution along the specimen's entire length.......................................................... 79
Figure 4-26 The variation in both plastic retained strain and resulting recovery strain
between Group 2 and Group 11, from extensometer and crosshead readings. ................... 80
Figure 4-27 The distribution of plastic strain between Group 0 and Group 1 throughout the
entire length of the specimen, with the failure location serving as the reference point. ..... 80
Figure 4-28 Stress-temperature curves with thermocouples (T1, T2, and T3) placed in the
upper third, middle, and lower third of the specimens. ....................................................... 82
Figure 4-29 The rise in temperature with extensometer strain through T1, T2 and T3
Thermo-Couples for group (2). ........................................................................................... 84
Figure 4-30 The rise in temperature with extensometer strain through T1, T2 and T3
Thermo-Couples for group (11). ......................................................................................... 84
Figure 4-31 Reduction of area (%) distribution across groups............................................ 87
Figure 4-32 Failure shapes in addition to the fracture surfaces for each specimen in every
group from group 0 to 8....................................................................................................... 87
Figure 4-33 Failure shapes in addition to the fracture surfaces for each specimen in groups
9, 10 and 11. Note that Group (9) is based on only two specimens due to limited
availability. .......................................................................................................................... 88
Figure 4-34 Grip impressions observed within Groups 3, 5, 7, 8, 9, and 10. ...................... 89
Figure 4-35 Grip impressions observed within Group 2 and Group 11. ............................. 90

XIV
CHAPTER 1

INTRODUCTION

1
CHAPTER 1

1 INTRODUCTION

1.1. General
In recent decades, the development of various shape memory alloys (SMAs) has garnered
significant attention due to their unique material properties. These properties have not only
intrigued researchers across different domains but have also found practical applications in
civil engineering. Notably, among SMAs, Ni-Ti based and Fe-based alloys are the primary
choices for civil engineering applications. Ni-Ti based SMAs, among all known
polycrystalline SMAs [1], exhibit exceptional super elasticity (SE) and shape memory effect
(SME). SMAs manifest in two distinctive phases: austenite and martensite. The forward
transformation denotes the shift from austenite to martensite, while the reverse
transformation entails the transition from martensite back to austenite. In the cooling
process, the forward transformation begins at the martensitic start temperature (Ms), with
the austenite phase converting to twinned martensite and completing the transformation to
martensite at the martensitic finish temperature (Mf). During heating, the reverse
transformation initiates at the austenitic start temperature (As) and concludes at the austenitic
finish temperature (Af). Austenite is characterized by a symmetric crystalline structure,
stable at high temperatures and low stresses, while martensite features a low-symmetry
lattice structure and is stable at low temperatures and high stresses [2].

SE refers to the material's ability to recover “Pseudoelastic” large strain upon the removal
of an applied load, while SME pertains to its capacity to regain strain upon heating. Both
these remarkable attributes, SE and SME, are intricately linked to martensitic transformation
(MT), a diffusionless phase transformation primarily driven by reversible shear lattice
distortion. When SMAs are deformed in their austenite phase, they undergo stress-induced
martensitic (SIM) phase transformation and give rise to multiple localized bands of
martensites.

In the realm of civil engineering, the selection of a specific SMA type is contingent upon the
intended application. SMAs have proven to be versatile materials in civil engineering,
serving purposes that capitalize on their pseudoelasticity, as exemplified by superelastic
bracing systems for seismic applications [3, 4], or their SME, encompassing applications

2
such as active confinement [5, 6] and heat activated Prestressing [7-10]. As SMAs continue
to gain recognition and acceptance as a viable construction material, it becomes imperative
to comprehend their intricate material behaviour from the perspective of civil engineering.

Numerous studies have previously focused on the material characterization [11-14],


shedding light on the sensitivity of SMA constitutive properties to various factors, including
alloy composition. Indeed, even slight alterations in the balance between elemental
constituents, particularly in systems like Ni-Ti, can significantly impact material behaviour.
Furthermore, the effects of heat treatments, such as high-temperature soaks and rapid
quenches, play a substantial role in shaping the microstructure of the material, inducing
changes in elastic, plastic, and ultimate failure properties. In addition to heat treatments,
processing techniques like cold-working and hot-working can also exert a profound
influence on material behaviour. Crucially, SMAs are capable of undergoing microstructural
changes that set them apart from other metallic systems. Consequently, they exhibit
heightened sensitivity to historical events and processes, making it vital for designers and
experimentalists to consider these details when selecting and testing SMA materials.

It is essential to recognize that the scope of this study does not aim to provide an exhaustive
review of the materials science and mechanics of SMAs, as the existing literature is
extensive. Readers seeking a comprehensive understanding of the thermo-mechanical
behaviour of SMAs and the parameters influencing their behaviour are directed to relevant
references [1, 15-19].

1.2. Problem Statement


The primary challenge at hand pertains to the observation that, even when subjecting two
material specimens to an identical set of conditions, encompassing factors such as
composition, atomic equilibrium, heat treatment, and manufacturing procedures, and
meticulously adhering to the well-established tensioning testing protocols, including ASTM
E8/E8M [20] for metallic materials and incorporating the additional provisions of ASTM
F 2516-07 [21] for SMAs, these specimens frequently manifest contrasting behavioral
patterns. Such variations give rise to differences in critical mechanical properties, notably
including the modulus of elasticity, critical stress thresholds, and their corresponding strains.
These properties hold paramount significance in the application and design of SMAs,
especially within the domain of civil engineering. The existing inconsistency poses a
potential risk to the reliable deployment of SMAs in civil engineering projects.

3
To address this pivotal concern, this study is dedicated to a comprehensive investigation
aimed at unraveling the underlying causes of these precarious discrepancies. Central focus
is to gain a profound understanding of the multifaceted relationship between the established
Standards and their procedural intricacies, and how they contribute to the emergence of these
inconsistencies.

Relation One pivotal aspect governed by the Standards is the extensometer gauge length and its
between
gauge location, with most round specimens requiring a length of 4D (diameter) for ASTM E8 and
length and
bar diameter
5D for ASTM E8M. For specimens without a reduced section, the extensometer gauge length
for determining yield behaviour should not exceed 80% of the distance between grips [20].
This parameter is of critical importance due to SMAs' propensity to form localized bands of
large strain during the SIM phase transformation, often referred to as transformation bands
(TBs) [22, 23]. These TBs can initiate at discrete locations and propagate along the length
of the test section in a wave-like manner [24]. Studies have revealed that loading specimens
to a displacement control mode with an 8% strain and measuring a consistent virtual gauge
length of 10 mm at four different locations results in significant variations. Some locations
may be fully transformed with an 8% strain, while others exhibit partial transformation with
a measured strain of 4.5%, and some remain linear elastic with a measured strain of 0.8%.
These discrepancies are attributed to the non-uniform strain evolution during the SIM phase
transformation [14]. It has been further observed that extensometers with short gauge lengths
are not particularly effective at measuring macroscopic strain, as they tend to indicate
"jumps" in strain when transformation is initiated in the local region where the extensometer
is attached [25].

Another noteworthy consideration is the method employed for strain measurements.


Research has indicated that the initiation of phase transformation often occurs near the grips,
resulting in inconsistencies when comparing strain measurements acquired via crosshead
displacement, covering the entire length of the specimen, with those obtained in a region of
homogeneous stress through extensometry [18]. Other studies show that, the start stress (σMs)
which is the critical stress for the start of martensite reorientation obtained from
extensometer readings being more than 100 MPa higher than those derived from crosshead
readings. This discrepancy can be attributed to the nucleation of the first TB, corresponding
to the onset of yielding, near the grips and outside the gauge length. Consequently, the strain
produced due to nucleation is not captured by the extensometer [14]. An additional facet of
concern is the role of ASTM F 2516-07 [21], which relies on the crosshead speed during

4
tensioning of SMAs. This parameter interacts with the slip induced by the thinning of the
specimen's cross-section due to nucleation or extension of TBs inside the anchors [14].

Moreover, the strain rate, and its influence, is a key consideration. The forward
transformation into martensite is exothermic, leading to various studies that explore the
effects of loading rates on material response, with particular focus on pseudoelastic loading
[25-27]. These studies have uncovered the significance of thermal effects, especially at
higher rates. In contrast, it has been demonstrated that an increase in the strain rate from
10-4 (1/s) to 10-1 (1/s) results in a higher number of TBs in Ni-Ti SMAs, contributing to a
more homogeneous overall deformation due to the increased number of TBs [23].

Furthermore, stress rates play a substantial role in testing systems, most testing systems also
provide the option to use force control, which may present advantages in some situations.
For example, using force control during pseudoelastic loading allows the experimentalist to
set an exact ending force value, such as 7MPa for ASTM standard testing [21]. However,
[18] showed that a constant force rate that is suitable during the elastic portion of loading
quickly leads to an unacceptably high strain rate during stress-induced transformation as the
testing system seeks to provide the same constant force rate.

Moreover, the issue of temperature variation in SMAs applications becomes apparent in


instances where certain methods induce temperature differences within the SMA
components, particularly notable in the case of Ni-Ti SMAs. A clear illustration of this
problem is evident in the work of Oudah and El-Hacha [28], where they proposed a
retrofitting solution for beam-column connections. Their approach aimed to shift the plastic
hinge region away from the face of the column by creating a vertical slot of 40 mm.
Superelastic Ni-Ti SMA rebars were used to reinforce the connection with all the rebar
embedded in concrete except the part in this slot, which was exposed to the surrounding
temperature. Another example involves an externally anchored reinforcement system where
a prestrained SMA bar or strip is initially fixed at both ends with grouting mortar.
Subsequently, the middle part of the specimen is heated to generate recovery stress.
However, this region is left exposed without grouting to leverage the corrosion-resistant
behaviour, particularly beneficial for Ni-Ti SMA, which exhibits corrosion performance
similar to that of stainless steel. Consequently, these methodologies introduce temperature
variation within the specimen, creating distinctions between the exposed section and the rest
of it. The narrow temperature difference of 15–34°C between the martensite and austenite

5
start temperatures in Ni-Ti SMAs poses a critical issue. This modest temperature variance
could lead to a transformation of the exposed part of the specimen from the austenite phase
in sunlight to the martensite phase in the evening. Given the substantially lower mechanical
properties of martensite compared to austenite, this transformation has the potential to induce
premature failure in this region.

Arguably, the most significant, yet often overlooked aspect, is the gripping pressure. This
parameter signifies the pressure at which a specimen is held during a tensile test.
Surprisingly, despite its potential influence, this factor remains unincorporated into
standardized testing protocols. Moreover, it has yet to be investigated to the best knowledge
of the authors.

1.3. Research Objective


In light of these multifaceted considerations, this study aims to comprehensively investigate
material behaviour through tensile testing. The focus is on elucidating the reasons behind
the notable variation observed in nominally identical SMA specimens. By scrutinizing how
aspects of Standards and their procedures contribute to these variations, the study aims to
provide valuable insights into these complex phenomena, ultimately enhancing the safe and
informed utilization of SMAs in civil engineering applications.

1.4. Research Methodology


This study is undertaken in two phases:

Phase 1

A literature review will be conducted to include a general introduction and a brief


background about shape memory alloys and their characteristics.

6
Phase 2

Experimental program was conducted to investigate the mechanical behaviour of binary


Ni-Ti SMAs under monotonic tensile loading conditions. It was aimed to elucidate the
impact of various parameters on the mechanical properties, post-fracture elongation,
temperature fluctuations, and failure shapes.

1.5. Organization of Thesis


This introductory chapter (Chapter 1) gives a brief introduction to the shape memory alloys

A review of the literature with a brief background about shape memory alloys and their
characteristics has been presented in Chapter 2.

Chapter 3 presents the detailed experimental study which is conducted to investigate the
mechanical behaviour of binary Ni-Ti SMAs under monotonic tensile loading conditions
with various parameters.

Chapter 4 deals with the results and discussions which are extruded from the tensile tests on
the binary Ni-Ti SMAs.

Chapter 5 presents a summary and the conclusions.

7
CHAPTER 2

LITERATURE REVIEW

8
CHAPTER 2
2 LITERATURE REVIEW
2.1. Introduction to Shape Memory Alloy (SMA)
Shape Memory Alloys (SMA’s) are categorized as smart materials that are composed of
elements such as Nickel (Ni), Titanium (Ti), Iron (Fe), Hafnium (Hf), Copper (Cu),
Aluminum (Al), Beryllium (Be), Zinc (Zn), Manganese (Mn), Zirconium (Zr), Palladium
(Pd), Gold (Au), Cadmium (Cd), and Silver (Ag). Combinations of these elements form
SMA’s with unique engineering properties. Several types of SMA’s reported in the literature
include: NiTi, NiTiFe, TiNiHf, TiNiCu, CuAlBe, CuZnAlMnZr, FePd, MnCu, FeMn,
CuAlNi, CuZnAl, CuZn, AuCd and AgCd. Among them, the most common type of SMA’s
is the Nickel-Titanium (NiTi), which was discovered by Buehler and coworkers during their
investigation for materials suitable for heat shielding applications [29]. They found that the
Ni-Ti SMA composition exhibited mechanical properties that were comparable to
commonly used engineering materials, but most importantly, the material also possessed a
shape recovery capability.

Since then, the Ni-Ti SMA has been extensively used in biomedical applications and
industries such as aviation [18]. Other applications of Ni-Ti SMA also include actuators,
switches, clamping devices, electronic cable connectors, valves, air conditioning vents, and
a variety of products [18]. Recently, the unique characteristic properties of Ni-Ti SMA were
found to have potential opportunities to be implemented for various civil engineering
applications [30]. These unique characteristic properties of Ni-Ti SMA lie in their ability to
undergo large deformations that can return to their original form through heating (shape
memory effect) or stress removal (superelasticity). Depending on the nature of the
application, the SMA’s can be formed into various shapes such as wires, bars, rings and
plates. The fabrication process of SMA’s usually involves melting alloys in a high-vacuum
or inert-gas environment. The alloys can then be hot-worked or cold-worked to the desired
shape. The final SMA product undergoes specific thermo-mechanical processing to exhibit
its distinctive shape memory effect and superelasticity behaviour [31].

9
2.2. Phase Transformation in Shape Memory Alloys
SMA’s have two phases, a high temperature phase called austenite (A) and a low temperature
phase called martensite (M). The austenite phase is considered as the parent phase and the
martensitic phase as the product phase. The transformation from one phase to the other can
occur by either applying an external stress or by changing the temperature. The SMA’s are
characterized by four distinctive temperatures (Figure 2-1): austenite start temperature (As),
austenite finish temperature (Af), martensite start temperature (Ms), and martensite finish
temperature (Mf).

Figure 2-1 Thermo-mechanical characteristics of SMA [18].


Similar to many metals and alloys, SMA exhibit polymorphism, which is more than one
crystal structure. When the material is cooled, the crystal structure transforms from austenite
to martensite through a process referred to as the forward transformation. Upon heating, a
reverse transformation occurs from the martensite back to the austenite phase (Figure 2-1).
During the forward transformation process, the austenite structure begins to transform to the
martensitic structure at the martensite start temperature (Ms), and the complete
transformation to the martensitic structure occurs at the martensite finish temperature (Mf).
The crystal structure formed at this stage is called a twinned martensitic structure. The
application of a mechanical load will deform the twinned martensitic structure through a

10
detwinning process giving rise to detwinned martensitic structure (Figure 2-1). To initiate
the detwinning process, the applied load must be sufficiently large. The minimum stress
required to initiate the detwinning process is termed as the detwinning start stress (σMs) and
the stress level required to complete the detwinning process is referred to as the detwinning
finish stress (σMf). Similarly, during the reverse transformation process (Figure 2-1), the
martensitic structure begins to transform back to the austenite structure at the austensite start
temperature (As), and a full transformation to the austenite structure is complete at the
austenite finish temperature (Af).

Figure 2-2 Illustration of SMA deformation phases [30].

The austenite phase generally has a cubic crystal structure. During the forward
transformation process, the twinned martensitic crystal formed can have different orientation
directions called a variant. The arrangement of these variants occurs such that the average
macroscopic shape change from austenite to martensite is negligible [18]. Under the
application of the load, the twinned martensitic structure transforms to a detwinned
martensitic structure by orienting a certain number of variants through a detwinning process.
The detwinning process involves the shear and movement of twin boundaries that allow for
high deformability without causing any slip (damage) to the material (Figure 2-2).
Subsequent heating of the SMA above the Af temperature will lead to a complete shape
recovery.

11
2.3. Shape Memory Effect (SME)
One of the unique thermo-mechanical properties of the SMA’s is their capability to exhibit
the shape memory effect (SME) phenomenon. The SME simply describes the forward and
reverse transformations by changing the temperature of the SMA. To better describe this
phenomenon, the thermo-mechanical loading path in a combined stress-strain-temperature
space is schematically shown in Figure 2-3. Considering a stress-free state, cooling the SMA
material from its austenite (parent) phase below the martensite finish temperature (Mf)
results in the formation of twinned martensite. At this stage, if the material is loaded, a linear
elastic stress-strain response is maintained until the minimum detwinning stress (σMs) is
reached. This marks the initiation of the detwinning process that continues with a yielding
plateau characterized by large strains and low stiffness until the detwinning finish stress
(σMf) is reached. Some alloys exhibit an increase in the stiffness once the detwinned
martensite phase is complete. Then, unloading the material to a state of zero stress will still
preserve the detwinned martensite structure and retain its deformation. Upon heating the
material above the austenite finish temperature (Af), the reverse transformation to the parent
austenite phase occurs with complete strain recovery. The recovered strain is usually termed
as the transformation strain (Ԑt). Subsequent cooling will again result in the forward
transformation of the SMA from the austenite phase to the twinned martensite phase and the
entire cycle can be repeated (Figure 2-3). This phenomenon is mainly referred to as the SME
of the SMA.

12
1-Compare with reinforcing steel stress-strain curve.
Figure 2-3 Shape Memory Effect of SMA [18]. 2- Is the behaviour in compression is the same?

If the SMA is already sustaining an applied load, the SME process is still effective but under
different transformation temperatures. The transformation temperatures (As, Af, Ms, Mf)
are directly related to the applied load. As the applied load (tension or compression)
increases, the transformation temperatures also increase. This relationship is demonstrated
as a positive slope of the transformation temperatures in the stress-temperature space (Figure
2-1). Thus, applying a load with a corresponding stress σ results in new transformation
temperatures denoted by Msσ, Mf σ, Asσ and Af σ. If the applied load in the austenite phase
is greater than the minimum detwinned stress (σMs), cooling the SMA will directly result
in detwinned martensite. Then, subsequent heating will induce the reverse transformation
from detwinned martensite to austenite, accompanied with a complete shape recovery under
the applied load (Figure 2-3).

13
2.4. Superelasticity
Another unique characteristic property of the SMA’s is the superelasticity behaviour, in
which the transformation (forward and reverse) process can be triggered by applying an
external mechanical load without the need to apply heat. To further expand on this
phenomenon, the unique stress-strain characteristics of the superelastic SMA is displayed in
Figure 2-4.

During the loading and unloading process, the stress levels at which the forward
transformation and reverse transformation occur are denoted by σMs, σMf, σAs and σAf, as
shown in Figure 2-1. The initial loading stages of the superelastic SMA exhibit a linear
elastic stress-strain response up to a stress level corresponding to σMs. This stress level marks
the initiation of a stress induced transformation from the austenite phase to the martensite
phase. At this stage, a yield plateau is observed characterized by increasing strains and a
reduction in stiffness with increasing loads. The transformation process to a detwinned
martensite is complete at a stress level corresponding to σMf. Upon the release of the load,
spontaneous recovery of the strain occurs. During the unloading process, another yield
plateau is experienced by the SMA that initiates at a stress level σAs corresponding to the
onset of the reverse transformation process. The material returns to its austenite state below
a stress level of σAf, in which an elastic recovery to a zero stress-strain state is followed
(Figure 2-4).

How can we define this stress level?

What is the recovery rate?

Figure 2-4 Superelasticity behaviour of SMA [18].

14
Figure 2-5 Effect of ambient temperatures on the thermo-mechanical properties of SMA
[32].

15
2.5. Effect of Ambient Temperatures on the Thermo-Mechanical
Properties of SMA
The thermo-mechanical properties of SMA are significantly dependent on temperature. For
typical structures, the ambient temperatures can be assumed to lie between -20°C and +60 °C.
The effect of temperature on the thermo-mechanical properties of the SMA can be categorized
into three cases (Case A, B and C) shown in Figure 2-5. In these figures, the parameter “Z” is
defined as the martensitic phase fraction that is used to track the progression of the
transformation phase on a scale from 0 % to 100 %. Specifically, Z = 100% when the material
is fully martensitic, and Z = 0% when the material is fully austenitic. The first case (Case A)
represents a state in which the SMA is stable at ambient temperatures in its austenitic form
(Figure 2-5). The corresponding thermo-elastic properties of the SMA in this state exhibit
superelastic behaviour, where loading the material will result in a stress induced martensite while
undergoing deformations. This deformation is completely recovered upon unloading the
material.

The second case (Case B) characterizes the martensitic SMA as stable at ambient temperatures
(Figure 2-5). In this case, loading the material will result in deformations that are sustained when
the material is unloaded. Upon heating, the retained deformations are completely recovered at
the temporary austenite high temperature phase followed by a forward transformation to the
martensitic phase as the temperature drops to ambient levels. If the deformed SMA is restrained
from movement during the heating process, the recovery stress during the temporary austenite
state will soon disappear once the material returns to its martensitic stable state at ambient
temperatures.

The third case (Case C) has a dual stable state of either martensite or austenite at ambient
temperatures, that is determined based on the direction of the transformation (Figure 2-5). If the
material is utilized during its forward transformation process (Point A), the SMA will exhibit an
austenitic superelastic response upon loading and unloading. Similarly, the SMA will experience
a martensitic state if utilized during the reverse transformation process (Point B). Loading the
SMA in this phase will induce deformations that are retained when the material is unloaded.
Under a strain-free state, heating the martensitic SMA above the Af temperature will cause
complete shape recovery as the material transforms to its austenitic form and sustains this form
as the temperature drop to ambient temperatures. If the material is constrained from movement,
the application of heat will generate a recovery stress that is maintained even after the removal
of heat. Unlike Case B, the recovery stress is maintained in this case because the SMA material
remains in its austenite phase as the temperature drops to ambient levels.

16
2.6. Alloy Composition of SMA
A wide variety of SMA’s were developed over the years by combining different alloying
elements. The alloy composition can be adjusted to suit the nature of the intended
application. The relative proportion of the constitutive metals can have a significant impact
on the thermomechanical properties of the SMA. For instance, it has been reported that a 1%
shift in the Nickel content of the Ni-Ti SMA material results in a 100°C change in the Ms or
As temperatures [18]. Therefore, selecting the type of the SMA alloying elements and their
composition is a key factor to achieve the desired results of a structural strengthening
application. In this section, several types of SMA’s are discussed including their capabilities
and limitations.
Can we refer to corrosion of SMA used with
2.6.1 Nickel titanium (Ni-Ti) based alloys concrete in different exposure conditions?

The most extensively studied and commonly used type of SMA’s are the Nickel Titanium
(Ni-Ti) based alloys. The transformation temperatures depend on the atomic percentage of
Ni. At the equiatomic composition of Ni-Ti (50% Ni and 50% Ti), the maximum recorded
Af temperature is 120°C. Decreasing the atomic percentage of Ni does not alter the
transformation temperatures, but increasing the atomic percentage of Ni can significantly
reduce the transformation temperatures of the Ni-Ti based SMA’s. For example, altering the
atomic percentage of Ni to 51 % instead of 50 % can reduce the Af temperature from 120°C
to -40°C [18]. The Ni-Ti based SMA’s are highly resistant to corrosion. A recovery stress
up to 900 MPa can be generated, and strains up to 8 % can be recovered by the shape memory
effect material characteristics [30]. Over the years, researchers have investigated the
possibility of adding a third alloying element to the Ni-Ti based SMA’s to enhance the
thermo-mechanical properties of the SMA. As a result, SMA types such as Ni-Ti-Cu and
Ni-Ti-Nb were produced.

2.6.2 Copper (Cu) based alloys


The main distinctive features of the Cu-based SMA’s is their good thermal and electrical
conductivity, in addition to their lower cost compared to the Ni-Ti based SMA’s. One of the
major drawbacks associated with the Cu-based SMA’s is their extreme sensitivity to the
composition of the alloying elements. A precise change in the composition from 10-3 % to
10-4 % can cause a change in the transformation temperatures within a range of 5°C. This
means that high difficulty is involved to produce Cu-based SMA’s with consistent thermo-
mechanical properties. Some of the most commonly used Cu-based SMA’s are Cu-Zn-Al

17
and Cu-Al-Ni alloys. In general, these alloys exhibit limited recovery stress levels of around
280 MPa and recovery strains of approximately 3-4%.

2.6.3 Iron (Fe) based alloys

The inclusion of Iron (Fe) to SMA’s was considered as a major breakthrough as it


significantly reduced the cost of SMA’s. In order to possess the shape memory effect, the
Fe-based SMA’s require several treatments such as thermo-mechanical training, centrifugal
casting, precipitation, and high-speed deformation [30]. The most commonly used ferrous
SMA’s are the Fe-Ni-Co-Ti and the Fe-Mn-Si. These alloys exhibit high strength and
Young’s modulus compared to the Ni-Ti based and the Cu-based SMA’s. The maximum
recovery stress and strain for the Fe-based SMA’s was reported to be around 400 MPa and
3 %, respectively.
2.6.4 General discussion on the types of SMA

The atomic composition and the mechanical properties of selected SMA’s are summarized
in Table 2-1 and Table 2-2, respectively. It can be recognized from Table 2-1 that the thermo-
elastic properties of the SMA’s depend on the type of alloy and their corresponding atomic
percentage in the SMA composition. In general, a variety of the Ni-Ti based and the Cu-
based SMA’s experience a thermo-elastic response, and most of Fe-based SMA’s do not
exhibit a thermo-elastic behaviour.

Table 2-1 Atomic composition for selected SMA’s [30].

18
Table 2-2 Mechanical properties for selected SMA’s [30].

Several conclusions can be drawn from Table 2-2 related to the mechanical properties of the
SMA’s. The elastic stiffness and the yield strength of the SMA’s in the austenite phase are
typically higher than that of the martensite phase. In general, the Cu-based and the Fe-based
SMA’s possess higher strength, stiffness, and elongation compared to the commonly used
Ni-Ti based SMA’s. However, reflecting on the objective of this project to utilize the SMA
for confinement applications, the two main parameters of the SMA’s mechanical properties
for this type of application are the recovery strain and the recovery stress. Experimental
studies reported in the literature [30, 32, 33] show that the Ni-Ti based SMA’s exhibit higher
recovery stress (up to 900 MPa) and recovery strain (up to 8 %), when compared to the
Cu-based and the Fe-based SMA’s (Table 2-2).

Other concerns related to the Cu-based and the Fe-based SMA’s are their susceptibility to
corrosion that can limit their external use for structural strengthening applications. In
contrast, the alloying elements used in the composition of the Ni-Ti based SMA’s do not
corrode, and therefore have a high potential for use in a wide spectrum of structural
strengthening applications. Another major concern is the production consistency in the
mechanical properties of the Cu-based and the Fe-based SMA’s. The long-term research
conducted on Ni-Ti based SMA’s for numerous applications supplied production companies

What is the electrochemical effect when using with steel rebars?


19
with sufficient information to produce highly consistent mechanical properties of the Ni-Ti
based SMA’s at large productions. In contrast, high variations in the mechanical properties
of the relatively newly developed Cu-based and Fe-based SMA’s were reported in the
literature, mainly attributed to the limited research conducted using these types of SMA’s.

2.7. Thermo-Mechanical Properties of SMAs

SMAs exhibit coupled thermomechanical behaviour, requiring the experimentalist to


carefully consider the particular inputs, or loading paths, applied to the material. To quantify
the complex behaviour of SMA materials, various loading paths are imposed while
phenomena associated with the phase transformation are recorded. These phenomena
include the shape memory effect and superelasticity, as discussed earlier. This should be
done through the application and measurement of three thermomechanical fields in
particular: stress, strain and temperature.

2.7.1 Effect of strain localization phenomena

It is well known that SMAs have the propensity to form localized bands of large strain during
the SIM phase transformation [22, 23]. The mechanical behaviour of SMAs is governed by
solid-solid phase transition process between different crystal structures [34].The phase
transition undergoes through discrete events of nucleation and growth of micro-domains at
grain level and involves intrinsic material instability and dissipative evolution of the micro
domains [23]. The phase transformation is accompanied by the localized release/absorption
of latent heat. Therefore, thermo-mechanical coupling is inherent in the process and
significantly influences the material behaviour both at micro- and macro-level [15, 23, 25,
35] Under the quasi-static tensile loading, the SIM phase transformation takes place through
the manifestation of macroscopically observable domain(s) (transformation bands), the
evolution of which is strongly coupled with heat transfer within the material and the
surrounding environment. The transformation bands (TBs) appear at the start of the stress
plateau in the stress-strain curve. The beginning and the end of the stress plateau in the stress-
strain curve of SMAs are associated with initiation and complete propagation of the localized
reorientation bands (TBs) [22]. Phase transformation, however, starts before the start of the
stress plateau in the stress-strain curve and continues even after the end of the stress plateau
[36, 37]. Over the stress plateau region, the strain in the TBs is limited to characteristic
transformation strain, which is generally indicated by the strain corresponding to the end of

20
stress plateau, until the complete propagation of TBs take place in the SMA [38]. The
characteristic transformation strain of an SMA depends on parameters such as grain size,
composition of alloy, strain-rate and the thermomechanical treatment [18, 19].

R. Suhail, J.F. Chen, G. Amato and D. McCrum [14] made two tensile tests on Ni-Ti-Nb
based SMAs, Test 1 ‘Strain’ (δ/l) was measured using the video-extensometer over a gauge
length (l) of 50 mm, which was specified at the mid height of the specimen using two target
pieces. In Test 2, full-field strain measurement of the specimen was obtained using Digital
Image Correlation (DIC). A gauge length of 50 mm was defined on strain contour maps to
compare the stress-elongation response of Test 1 with that of Test 2 that was shown in (Figure
2-6).

Figure 2-6 Measured nominal stress – elongation response obtained from Test 1 and 2 [14].

21
Please check.

From (Figure 2.7), The authors obtained also the Stress-Strain for test 2 from the crosshead
and it is could be seen based on the tangent lines the start stress (σMs) obtained from DIC
was 700 MPa from (Figure 2.6) and the same stress obtained from crosshead was 608 MPa
as in (Figure 2.7). This difference in the stress value was explained as the nucleation of the
first TB that corresponded to the onset of yielding was formed near the grip and outside the
virtual gauge length. As a result, the strain produced due to nucleation was not recorded by
the virtual extensometer. Any increase in the stress after the nucleation of TB appears as a
part of linear elastic portion of the stress-elongation curve until either one or more TB
extends inside the gauge length or a new TB is nucleated inside the gauge length.

Figure 2-7 Measured nominal stress – crosshead displacement response in Test 2 [14].

During Test 2, images at different stages were correlated with a reference image to establish
the strain distribution in the specimen through Digital Image Correlation (DIC) as shown in
(Figure 2.8). Strain contours were then sequentially plotted to illustrate the temporal
evolution of the full-field strain distribution. The entire length of the specimen, except for
3–6 mm near the grips, was investigated. The area analyzed, termed the Area of Interest
(AOI), excluded the nucleation points of the first two Transformation Bands (TB-2 and TB-

22
4) near the grips. The nucleation of TBs led to the propagation of high-strain fronts, acting
as a heat source. The stress plateau showed a positive slope due to the release of latent heat
during nucleation, with subsequent TB nucleations influencing the stress-drop and front
propagation. As adjacent TBs coalesced, the specimen transformed fully into martensite.
The strain became nearly uniform (8%) over the stress plateau, while non-uniform strain
distribution, resembling Lüders-like bands, was observed after full transformation. This non-
uniformity, particularly in Ni-Ti-Nb SMAs, poses challenges in structural applications,
affecting deformation patterns and potentially influencing bond behaviour in applications
like reinforcing bars.

Figure 2-8 Strain contours showing the formation and evolution of TBs along the length in
Test 2 [14].

The increase of strain from the start of the stress plateau (σMs) to the end of the stress plateau
(σMf) was measured equal to 7.3%. Over the stress plateau, the local strain within all TBs
was limited to ~8% (see Figure 2.9).

23
Figure 2-9 Strain distribution along the length of the specimen at various stages from
initiation to complete propagation of TBs in Test 2 [14].

2.7.2 Stress-strain behaviour of SMAs under repeated cyclic loading


The authors also in [14] illustrated in test 3 the stress-elongation response of a Ni-Ti-Nb wire
under cyclic loading. The virgin specimen underwent cyclic loading and unloading under
displacement control. The loading and unloading cycles involved three phases, as illustrated
in (Figure. 2.10). In the 1st cycle, the specimen was loaded from Point a to d and then
unloaded to zero stress (Point e). The 2nd cycle involved loading from Point e to g and
unloading to zero stress (Point i). Lastly, the specimen was loaded from Point i until fracture
occurred at Point k. Elongation responses were measured using DIC over a virtual gauge
length of 50 mm at the mid-height of the specimen. The strain rate for the test was
1.5 × 10−3 (1/s).

24
Figure 2-10 Measured nominal stress – elongation response of a virgin Ni-Ti-Nb wire
Specimen in Test 3 [14].

In the examination of cyclic tensile represented by Test 3, the full-field strain distribution of
the Ni-Ti-Nb wire was analyzed. The test involved multiple cycles of loading and unloading,
each marked by the formation and propagation of transformation bands (TBs). TB
nucleation, propagation, and coalescence were observed, revealing distinctive behaviors
during the 1st loading cycle. The TBs exhibited a limiting strain of approximately 8%,
consistent with findings from previous tests. Upon unloading, strain recovery and residual
strain patterns were examined, providing insights into the wire's mechanical response. In the
2nd loading cycle, the deformation initially occurred within the previously formed TBs,
followed by broadening and coalescence. A significant stress-drop was noted during
unloading, potentially attributed to slip at the grips. The test concluded with the complete
transformation of the specimen to the martensite phase during the 3rd loading cycle.

25
Figure 2-11 Strain contour of Test 3 specimen showing evolution of TBs during loading and
unloading for strain range of 0−9% [14].

In the examination of cyclic loading behaviour, the stress-elongation response of the Ni-Ti-
Nb wire was found to exhibit significant variations due to the non-uniform evolution of strain
during the SIM phase transformation. This variability was clearly depicted in (Figure 2.12),
where stress-elongation curves were constructed based on local measurements at four
distinct locations on the specimen: TB-2b, TB-3 (representing transformed regions), and FrL
and FL-1 (representing non-transformed areas) as shown in (Figure 2.11).

For TB-2b and TB-3, which underwent complete transformation by the end of the 1st loading
cycle, their stress-elongation curves displayed overlapping branches in subsequent cycles.
In contrast, FrL experienced partial transformation during the 1st loading cycle, resulting in
a measured strain of 4.5% within the gauge length, significantly less than the ~8% measured
in the transformed regions (TB-2b and TB-3). The unloading in the 1st cycle occurred at a
strain of 4.5%, retaining a residual strain of about 2.8% at FrL, in comparison to the ~6% in
the transformed regions. During the 2nd loading cycle, the strain in the specimen increased

26
but was constrained to the transformation strain until complete transformation. Unloading in
the 2nd cycle took place at a strain of ~8%, aligning with the complete transformation within
the gauge length at this stage. Notably, FL-1 remained linear elastic during the 1st cycle,
causing the unloading branch to overlap with the loading one. The reloading of the 2nd cycle
followed the initial elastic branch, reaching yielding during the process. Unloading occurred
at a strain of 5.5%, with a residual strain of 4.2%. Full transformation of FL-1 occurred only
in the 3rd loading cycle. A crucial insight derived from this analysis is the non-homogeneous
nucleation of transformation bands (TBs) along the length of SMA wires. This emphasizes
the importance of carefully selecting an appropriate gauge length to obtain an accurate
representation of the stress-elongation response of SMA wires.

Figure 2-12 Nominal stress – elongation response measured at (a) TB-2b; (b) TB-3; (c)
fracture location (FrL); (d) at location FL-1; (region between TB-1 and TB-3) [14].

27
2.7.3 Effect of strain rate

The forward transformation into martensite is exothermic while the reverse transformation
into austenite is endothermic [18]. This introduction of the latent heat phenomenon and
resulting possible temperature changes during the characterization process imposes
additional constraints on the loading rates applied to SMA specimens. The challenge is
particularly applicable to pseudoelastic testing. Such experiments are often performed at
constant temperature to simplify interpretation of the transformation behaviour. During
loading, however, thermal energy can be added to the specimen if the heat generated during
the exothermic forward transformation is not dissipated. If loading is performed slow
enough, convective and conductive processes will remove this additional heat without
noticeably raising the specimen temperature. If loading or unloading progresses too rapidly,
the temperature of the specimen will rise during loading and fall during unloading, violating
any isothermal assumptions. Several studies have been performed to assess the influence of
loading rates on material response. Many of these address pseudoelastic loading [25-27] and
have found the thermal effects present during this exothermic/ endothermic cycle to be
significant, especially at higher rates. Throughout the various research efforts in the area of
shape memory alloys, many displacement and thermal loading rates have been suggested.
For isothermal loading at any temperature, rates should not exceed ∼ 0.01-0.05% strain/sec
[21, 25]. This helps to ensure quasi-static testing. Others have also explored the martensitic
reorientation (detwinning) rate dependency in various stress configurations and in various
alloys [39, 40], and have shown that the detwinning process is generally independent of the
strain rate.

28
2.7.4 Effect of stress rate
The use of reasonable strain rates is straightforward in strain-driven or deformation-driven
experimentation, but not all thermomechanical experiments on metals are performed using
prescribed strain or displacement inputs. Most testing systems also provide the option to use
force control, which may present advantages in some situations. For example, using force
control during pseudoelastic unloading allows the experimentalist to set an exact ending
force value, such as 0MPa or 7MPa for ASTM standard testing [21]. However, for SMAs
exhibiting “flat” pseudoelastic loading/unloading plateaus, force control can lead to
problems. A constant force rate that is suitable during the elastic portion of loading quickly
leads to an unacceptably high strain rate during stress-induced transformation as the testing
system seeks to provide the same constant force rate. An example of this is shown in Fig.
2.13. Here the main result involves a specimen undergoing two pseudoelastic loading cycles
at an ambient temperature of 37 °C (nominal). For the first test, a strain rate of 0.05%/s was
applied during both loading and unloading. For the second test, a strain rate of 0.05%/s was
applied during loading, but a 0.08 N/s force rate was applied during unloading. In
(Figure. 2.13a) the forward loading (forward transformation) paths for both tests are nearly
identical while the unloading (reverse transformation) curves are not. During unloading, the
use of force control in the second test caused high strain rates as the material response
reached the lower plateau stress. This in turn caused a reduction in specimen temperature
from 37 °C to 35°C as seen in (Figure. 2.13b.) This temperature change causes the unloading
curve to follow a different isothermal load path (i.e., the result for a test temperature of 35
°C) at the end of reverse transformation.

29
Figure 2-13 Comparison of unloading control methods for pseudoelastic testing:
(a) stress-strain response, (b) specimen surface temperature response [18].

30
2.7.5 Effect of thermal hysteresis band on recovery stress

The SME is most pronounced in SMAs that have a wide thermal hysteresis between their
martensite and austenite phases. The start and finish temperatures of martensite and austenite
phases are listed in Table 2 for Ni-Ti, Ni-Ti-Nb, Cu, and Fe-SMA. Notably, the difference
between the martensite and austenite start temperatures is greater than 100 °C for Ni-Ti-Nb
and Fe-SMA, which makes them suitable for generating large recovery stresses in the order
of 130–580 MPa and 354–560 MPa, respectively. In contrast to Ni-Ti-Nb and Fe-SMAs, the
difference in the martensite and austenite start temperatures is 15–34 °C for Ni-Ti SMAs
and 20 °C for Cu-SMAs as shown at Table 2-3. This (relatively) small temperature difference
minimizes the potential of Ni-Ti and Cu-SMAs to exhibit a notable SME. Studies conducted
on strengthening with Ni-Ti SMAs report the generation of high recovery stresses in the
order of 400–700 MPa at activation temperatures of 100–200 °C after the SMAs are
prestrained in the range of 5%–8% [41, 42]. However, a significant drop in the generated
recovery stresses occurs because of short-term relaxation once the heating process is
completed and the temperature decreases. This is because the ambient temperature should
remain high enough for the SMA to remain in the austenite phase and maintain the recovery
stress. However, because the ambient temperature is often close to the martensite phase
temperature of Ni-Ti SMAs, they transform back to their martensite phase upon cooling,
resulting in the loss of generated recovery stresses [43]. Notably, to effectively use the SME
of SMAs for civil engineering applications, the austenite start temperature should be higher
than the high ambient temperature (i.e., 40 °C) and the martensite start temperature should
be lower than the low ambient temperature (-10 °C) [42].

Table 2-3 Martensite and austenite start and finish temperatures [45].

31
2.8. Research Gaps
Existing literature reveals significant gaps in the study of shape memory alloys (SMAs).
Firstly, the majority of studies have concentrated on subjecting SMAs to a constrained
prestrain value, often neglecting the exploration of behaviour until failure. Secondly, crucial
parameters, including gauge location, gauge length, temperature variation and gripping
pressure during tensile tests, have not received adequate attention. These oversights are
noteworthy as understanding their influence is crucial for accurately assessing the
mechanical properties resulting from tensioning tests, particularly for the safe utilization of
SMAs in civil engineering applications.

32
CHAPTER 3

EXPERIMENTAL PROGRAM

33
CHAPTER 3
3 EXPERIMENTAL PROGRAM
Experimental program was conducted to investigate the mechanical behaviour of binary Ni-
Ti SMAs under monotonic tensile loading conditions. It was aimed to elucidate the impact
of various parameters on the mechanical properties, elongation post-fracture, temperature
fluctuations, and failure shape of the specimens.

3.1. Materials and Specimens


The experimental program began by utilizing binary Ni-Ti SMA specimens, characterized
by a composition typical of Ni-Ti alloys with 55.8 wt. % Ni and a diameter (D) of 4 mm.
Notably, due to the austenite finish temperature (AF) being 20°C, which falls below room
temperature (RT) in the range of 22-25 °C, all specimens were maintained in their austenite
phase throughout the experiments. The chemical composition of the SMA, expressed in
equivalent weight percentage (wt.%), and the phase transformation temperatures were
provided by the supplier, (Baoji Lihua Non-ferrous Metals Co. Ltd), and can be found in
Table 3-1 and Table 3-2, respectively. Key mechanical properties, including tensile strength
and maximum elongation, were consistent across all specimens and were obtained from the
supplier, as documented in Table 3-3. It is worth noting that no information regarding
thermo-mechanical treatment or heat treatment history was available for these specimens.

Table 3-1 Chemical Composition (wt %) From Supplier.

Alloying
Ti Ni Co Cu Nb Fe C O+N H
element

< < <


Ni, Ti Remainder 55.80 0.011 0.008 0.039 0.0022
0.01 0.005 0.01

34
Table 3-2 Transformation Temperatures (°C) From Supplier.

Mf Ms As Af Room Temperature (RT)

-10 0 12 20 22-25

Table 3-3 Mechanical Properties from Supplier.

Elastic Modulus (GPa) Tensile Strength (MPa) Yield Strength (MPa) Elongation (%)

20 840 405 17.5

3.2. Experimental Setup


Tensile tests on the Ni-Ti specimens were performed using a Zwick Roell universal testing
machine (UTM) equipped with a 600 kN load cell. Testing continued until the specimens
reached failure. Strain measurements were acquired through two distinct methods:

i. Extensometer Strain: An extensometer, attached directly to the specimen via knife


edges mounted on the sensor arms, measured the total deformation within the gauge
length.
ii. Crosshead Strain: This method involved measuring the total crosshead displacement
within the entire specimen length between the grips, Fig 3-3a represents the stress vs
extensometer strain and stress vs crosshead strain curves.

To minimize potential inaccuracies arising from slip within the grips, all tensile tests
underwent a pre-load value with 20 MPa before the actual testing commenced. It is essential
to note that due to the distribution of the preload, measurements of extensometer strain and
crosshead strain were taken along the modified length of the gauge and the modified total
length of the specimen between the grips, respectively.

Throughout the experiments, the specimens' length remained constant at 120 mm (clear
distance between the grips) and the gripping length inside each grip was 85 mm. To ensure
35
consistency and accuracy, temperature monitoring was conducted using three Type-K
thermocouples affixed directly to the specimen's surface, with placements in the upper third,
middle, and lower third. To eliminate any residual transformation strain and ensure
consistent starting conditions, all specimens underwent a controlled thermal treatment. Prior
to testing, each specimen was heated above its (Af) to 100°C using a heat gun. This step was
critical for the recovery of any potential transformation strain within the Ni-Ti (SMA). To
facilitate post-fracture elongation measurements, gauge marks were created on the
specimen's surface using a metal marking machine. These marks were spaced at 1 cm
intervals along the entire specimen length, Fig 3-1 represents the full test setup.

Several attempts were made to induce temperature variation in the extensometer gauge
length region compared to the rest, employing cooling sprays and cooler devices. However,
these initiatives faced challenges in maintaining a consistently lower temperature before the
test and encountered stochastic variations. The most successful solution involved encasing
the extensometer gauge length with an ice pack Fig 3-2b, ensuring a nearly constant and
lower temperature for the attached region. This method resulted in maintaining an almost
constant lower temperature of approximately 12°C for this specific region, ensuring it
remained in the austenitic phase as it was above its Ms and Mf degrees. Meanwhile, the rest
of the specimen maintained a temperature of about 22°-25°C.

3.3. Experimental Objectives


The primary objectives of the tensile tests were to evaluate the effect of various
parameters on:

a) Mechanical Properties: This included analyzing data from stress-strain curves


to determine the modulus of elasticity of Austenite (EA), modulus of elasticity
of Martensite (EM ) post-forward transformation into martensite, stress start (σMs) and
strain start (Ԑs) at the initiation of forward transformation into martensite (referred to as
elastic yield stress), stress finish (σMf ), and strain finish (Ԑf ) at the completion
of forward transformation into martensite, yield plastic stress (σy) and yield plastic
strain (Ԑy), ultimate stress (σu) and ultimate strain (Ԑu), The values for the stresses σMs,
σMf and σy and their corresponding strains Ԑs, Ԑf and Ԑy were obtained by drawing
tangents on both stress- strain curves for extensometer and crosshead readings. Fig 3-
3(b) and (c) represent tangent lines on the both curves.

36
b) Post-Fracture Elongation Analysis: After fracture, elongation measurements were
conducted by aligning the two fractured specimen halves and precisely measuring the
distance between gauge marks. Subsequently, these measurements were compared with
elongation data acquired through both extensometer and cross-head methods. The
primary aim of this analysis was to assess the extent of plastic strain retained along the
entire specimen length and its distribution. Furthermore, this investigation included an
assessment of recovery strain, an aspect that has been relatively underexplored in
previous studies, particularly in the context of post-failure behaviour.
Is there a c) Temperature Fluctuations Assessment: A systematic analysis involved the
correlation
between continuous monitoring of temperature variations using three strategically placed thermo-
STRESS-STRAN-
TEMPERATURE
couples. This endeavour aimed to meticulously observe temperature fluctuations and
discern the potential influence of individual parameters on these fluctuations.
d) Analysis of Failure Characteristics: A comprehensive examination was conducted to
scrutinize the ramifications of each parameter on the failure shape of the specimens.
Additionally, reductions in specimen area subsequent to failure were meticulously
valuated to gain insights into the post-failure structural changes.

3.4. Study Parameters


Table 4 provides a comprehensive summary of the parameters examined in this experimental
program. Each group within the study comprised three pristine specimens subjected to
monotonic tensile loading, with the exception of Group (9), which utilized only two
specimens due to limited availability.
In Group (0), serving as the reference, adherence to ASTM 'E8/E8M' standards was
maintained [19]. The extensometer gauge length was set at 40 mm, centered within the
specimen to ensure it remained below the recommended maximum length, equivalent to
80% of the distance between the specimen's grips. To determine elastic yield properties, a
strain rate of 2.5 x 10-4 (1/s), corresponding to 0.015 mm/mm/min as recommended, was
applied. After capturing the yield behaviour, the testing machine's speed transitioned to a
crosshead displacement rate of 6.7 x 10-3 (1/s), aligning with the recommended range of 0.05
to 0.5 mm/mm/min.
Groups (1,2) investigated the impact of varying extensometer gauge lengths, setting them at
20 mm and 80 mm, corresponding to 5D while remaining below the recommended
maximum length respectively [19]. Recognizing that phase transformation initiation often
occurs near the grips, leading to measurement inconsistencies between strain acquired via

37
crosshead and extensometer readings, Groups (3,4) strategically positioned the extensometer
gauge length in the upper and lower thirds, adjacent to the movable upper grip and fixed
lower grip, respectively.
To assess the effect of increased strain rates, Groups (5,6) exposed specimens to strain rates
of 2.5 x 10-3 and 2.5 x 10-2 (1/s), adjusting the crosshead displacement rate to match at
5 x 10-2 (1/s) after yielding. Additionally, the study explored the transition from strain rate
control to stress rate control. Group (7) employed a speed of 3 MPa/s, falling within the
recommended range of 1.15 to 11.5 MPa/s [19], while Group (8) increased it to 30 MPa
above recommendation values.
A novel parameter introduced in this study was gripping pressure, representing the pressure
Gripping at which the specimen grips held during the tensile test, ranging from 40 to 480 bar according
pressure
to the machine's manual. The base group maintained the minimum pressure with 40 bar,
while Groups (9) and (10) applied gripping pressures of 80 and 120 bar, respectively.
In Group (11), the study explored the impact of lowering the temperature for a specific part
of the specimen, specifically the extensometer gauge length, compared to the rest. The values
for other parameters remained consistent with those of Group (2) see Fig 3-2 (a) and (b).

These meticulously selected parameter variations aimed to provide a comprehensive


understanding of the mechanical behaviour of Ni-Ti SMAs under diverse testing conditions,
yielding valuable insights into their responses to varying parameters.

38
(a) (b) (c)

Figure 3-1 (a) Test set-up for tensile tests, (b) close-up view of the specimen and (c) Gauge
marks.

(a) (b)
Figure 3-2 (a) groups with constant temperature are exposed to RT and (b) group 11 with
temperature variation in the gauge length exposed to ice pack.

39
Table 3-4 Experimental Program.
- Strain - Strain
Loading Rate Rate
Extensometer Loading Gripping Specimen
Extensometer Method (1/s) (1/s)
Parameter Gauge Length Method After Pressure Temperature
Location Till - Stress - Stress
(mm) Yielding (bar) (°C)
Yielding Rate Rate
(MPa) (MPa)

6.7 x
Strain 2.5 x 10-4 Displacement 22-25 °C for
Group (0) 40 Middle 10-3 40
Control (1/s) Control entire length.
(1/s)

6.7 x
Strain 2.5 x 10-4 Displacement 22-25 °C for
Group (1) 20 Middle 10-3 40
Control (1/s) Control entire length.
(1/s)

6.7 x
Strain 2.5 x 10-4 Displacement 22-25 °C for
Group (2) 80 Middle 10-3 40
Control (1/s) Control entire length.
(1/s)

6.7 x
Strain 2.5 x 10-4 Displacement 22-25 °C for
Group (3) 40 Upper Third 10-3 40
Control (1/s) Control entire length.
(1/s)

6.7 x
Strain 2.5 x 10-4 Displacement 22-25 °C for
Group (4) 40 Lower Third 10-3 40
Control (1/s) Control entire length.
(1/s)

Strain 2.5 x 10-3 Displacement 5 x 10-2 22-25 °C for


Group (5) 40 Middle 40
Control (1/s) Control (1/s) entire length.

Strain 2.5 x 10-2 Displacement 5 x 10-2 22-25 °C for


Group (6) 40 Middle 40
Control (1/s) Control (1/s) entire length.

6.7 x
Stress Displacement 22-25 °C for
Group (7) 40 Middle 3 (MPa) 10-3 40
Control Control entire length.
(1/s)

6.7 x
Stress Displacement 22-25 °C for
Group (8) 40 Middle 30 (MPa) 10-3 40
Control Control entire length.
(1/s)

6.7 X
Strain 2.5 x 10-4 Displacement 22-25 °C for
Group (9) 40 Middle 10-3 80
Control (1/s) Control entire length.
(1/s)

6.7 x
Strain 2.5 x 10-4 Displacement 22-25 °C for
Group (10) 40 Middle 10-3 120
Control (1/s) Control entire length.
(1/s)

22-25 °C
6.7 x except the
Strain 2.5 x10-4 Displacement
Group (11) 80 Middle 10-3 40 extensometer
Control (1/s) Control
(1/s) length was
12 °C

40
(a)

σMf
σMs

σMf
σMs

Figure 3-3 (a) Stress vs. extensometer strain and stress vs. crosshead strain curves, (b) and
(c) stress-strain curves depicting tangent lines to determine the mechanical properties.

41
CHAPTER 4

RESULTS AND DISCUSSION

42
CHAPTER 4
4 RESULTS AND DISCUSSION

4.1 Mechanical Parameters


The investigation of mechanical parameters entailed a comprehensive analysis of stress-
strain curves, providing valuable insights into the complex behaviour of SMAs. These
curves, obtained from both extensometer and crosshead strain measurements, revealed
notable differences.

4.1.1 Influence of gauge length variations

4.1.1.1 Modulus of elasticity


The examination of how different extensometer gauge lengths influence the Modulus of
Elasticity (EA) in Ni-Ti SMAs under diverse testing conditions is a central focus of this
study. A comprehensive perspective on this impact is offered through Table 4-1 and
Fig 4-1, which provide valuable insights. One of the main target is to explore the effect of gauge length
variation on tensile properties of SMA.

At an extensometer strain rate control of 2.5 x 10-4 (1/s), the influence of distinct gauge
lengths (20, 40, and 80 mm) on EA from extensometer readings become apparent. These
variations lead to different force increase rates, resulting in values of 8.4, 12.5, and 6 MPa/s,
What is the respectively. The 40 mm gauge length stands out with the highest EA value, reaching 50
relation between
specimen GPa, while the 20 mm and 80 mm gauges yield 30.86 and 23.76 GPa, respectively. It is
dimeter and optimum
gauge length? important to highlight that these significant variations become apparent within the elastic
region, occurring prior to the formation of transformation bands (TB). These TBs are
recognized for inducing strain localization effects and typically develop after the elastic
region has concluded. Cross head readings are not used to evaluate any properties
due to grips slippage?
Interestingly, EA values obtained from crosshead readings deviate from extensometer
readings, yielding EA values of 23.94, 27.23, and 17 GPa for 20, 40, 80mm gauge lengths.
This deviation highlights the complex behaviour of SMAs, as both 20 mm and 80 mm gauge
lengths require the same crosshead speed of approximately (3.4 x 10-4) (1/s), despite
differing force increase rates (8.4 and 6 MPa/s).

The notable disparity in EA values between the extensometer and crosshead readings for
each parameter becomes more understandable when considering the substantial impact of
gauge length alterations along the specimen on the extensometer readings. This observation

43
underscores the consistent influence of varying the gauge length, from a portion of the
specimen to its entire length, on EA readings.

Regarding EM, after the forward transformation into Martensite and the completion of TB
propagation through the specimen's entire length, following the transition to crosshead speed
rate control (approximately 6.7 x 10-3 1/s), the strain increased along the entire length of the
specimen, although non-uniformly, until the fracture occurred. This non-uniformity in strain
distribution immediately after yielding is a characteristic feature of SMAs, distinguishing
them from common metals such as mild steel [27]. The EM values obtained from
extensometer readings exhibited slight variations, ranging from 17 to 21 GPa.

However, examining EM values derived from crosshead strain readings reveals an intriguing
observation: they remain consistent at around 10 GPa. This close resemblance in values
might seem initially to be a positive indication. However, this can be attributed to slip at the
grips, a phenomenon observed due to the continued extension of TB inside the grips during
the Martensite transformation. This action considerably and rapidly diminishes the local
cross-sectional area of the specimen [13].

And the evidence for that is that the crosshead speed rate values are nearly double that of
extensometer readings, especially for 80 mm gauge length which is 67% from the length
between grips. In addition, a relatively low required force increase rate for the crosshead
speed needed, which is approximately 69 MPa/s. This force increase rate is considerably
lower compared to the force increase rate needed to achieve a similar speed by increasing
gripping pressure, as will be discussed in section 4.1.5.

44
Table 4-1 Modulus of Elasticity for gauge length variation.

Extensometer Readings

Force
Parameter Strain rate Force Increase
EA Strain rate Increase EM
Gauge for EM rate for EM
length (GPa) for EA (1/s) rate for EA (GPa)
variation (1/s) (MPa/s)
(MPa/s)

20 (mm) 30.86 0.00025 8.4 19.66 0.0037 69.7

40 (mm) 50.04 0.00025 12.5 17.37 0.004 65.3

80 (mm) 23.76 0.00025 6.0 21.22 0.00345 69.2

Crosshead Readings

Force Crosshead
Parameter Crosshead Force Increase
EA Increase EM Speed
Speed rate for EM
(GPa) rate for EA (GPa) for EM
for EA (1/s) (MPa/s)
(MPa/s) (1/s)

20 (mm) 23.94 0.00034 8.4 10.69 0.0067 69.7

40 (mm) 27.23 0.00046 12.5 10.03 0.00665 65.3

80 (mm) 17 0.00034 6.0 10.84 0.0067 69.2

45
(a) (b)
Figure 4-1 Variation of EA and EM with gauge length from (a) extensometer and (b)
crosshead readings.

4.1.1.2 Stresses
Table 4-2 and Fig 4-2, provide that varying extensometer gauge lengths has no effect on the
stress values from extensometer and crosshead reading. The remarkable thing is that the σMs
obtaining from crosshead reading is always lower than the corresponding extensometer
reading because nucleation of TB always happens near the grips which is not captured with
extensometer unless expanding the TB inside the gauge length. This was observed for all
following groups for every parameter.

Table 4-2 variation for stress values with gauge length from extensometer and crosshead
readings.

Extensometer Readings (MPa) Crosshead Readings (MPa)


Parameter
Gauge
length σMs σMf σy σu σMs σMf σy σu
variation

20 (mm) 309.53 456.12 1198.83 1258.48 289.38 443.98 1190.53 1258.48

40 (mm) 315.56 470.91 1117.63 1189.84 285.7 461.54 1117.83 1189.84

80 (mm) 291.75 438.21 1210.8 1269.07 260.09 429.48 1206.93 1269.07

46
(a) (b)
Figure 4-2 Variation for stress values with gauge lengths (a) from extensometer and (b)
crosshead readings.

4.1.1.3 Strains

The influence of varying extensometer gauge lengths on strain parameters is a fascinating


aspect of the study. Table 4-3, in conjunction with Fig 4-3, reveal that these variations have
a minimal effect on the strain values obtained from both extensometer and crosshead
readings.

When each group's values from extensometer readings are compared with their
corresponding values from crosshead readings, very slight differences in Ԑs and Ԑf were
observed. However, these differences become more pronounced when examining Ԑy and Ԑu.

The significant divergence especially in Ԑu values, in particular, can be attributed to two key
factors. Firstly, as previously mentioned, the substantial slip, especially after the
transformation into Martensite, contributes to this variation. Secondly, the identical cross-
sectional area at the ends of a wire specimen and within the gauge length results in the
concentration of excessive stress, leading to premature failure near the grips.

A noteworthy trend is that groups with the highest modulus of elasticity EA consistently
exhibit the highest start stress (σMs) and the lowest initial strain (Ԑs). This pattern extends to
form the lowest Modulus of Elasticity (EM), resulting in the lowest ultimate stress (σu) and
the highest strain values for ultimate strain (Ԑu). Conversely, groups with lower EA values

47
exhibit the opposite pattern, emphasizing the intricate relationship between strain
parameters, modulus of elasticity, and gauge length variations.

Table 4-3 Strain values with variation of gauge lengths from extensometer and crosshead
readings.

Extensometer Readings (%) Crosshead Readings (%)


Parameter
Gauge
length Ԑs Ԑf Ԑy Ԑu Ԑs Ԑf Ԑy Ԑu
variation

20 (mm) 0.83 5.65 9.43 16.19 1.09 6.82 13.8 26.11

40 (mm) 0.61 4.9 8.67 18.06 1.02 6.88 13.43 29.51

What is the
reason? 80 (mm) 1.12 5.12 8.76 15.12 1.35 6.98 14.15 27.33

(a) (b)
Figure 4-3 Variation for strain values with gauge lengths (a) extensometer and (b)
crosshead readings.

48
4.1.2 Influence of gauge length location

4.1.2.1 Modulus of elasticity

Examination of EA at an extensometer strain rate of 2.5 x 10-4 (1/s) at different gauge length
locations and utilizing Table 4-4 along with Fig 4-4, it becomes evident that gauge length
locations in the upper third and lower third, closer to the grips, yield almost the same and
the lowest force increase rates. This results in nearly identical EA values of approximately
31 GPa from extensometer readings and about 23 GPa from crosshead readings. In contrast,
the gauge length location at the middle, which requires the highest force increase rate, results
in the highest EA values, with around 50 GPa from extensometer readings and about 27 GPa
from crosshead readings.

This yields that the variation in the beginning of the test near the grips regions whether the
upper or the lower one are almost the same. In contrast, the variation in the middle region
has a stochastic and significant variation. This was especially evident in the EA values
obtained when altering the gauge length in the middle region, as seen in the previous group.
All exhibited the difference between the extensometer and crosshead reading for the same
parameter.

For the Modulus of Elasticity for Martensite (EM), a similar trend was observed to the
previous group, indicating that not only are EM values similar between extensometer and
crosshead readings but the values also closely align with those from the earlier group,
ranging from 17 to 20 GPa for extensometer readings and an average of around 10 GPa for
crosshead readings. This suggests that changing the gauge lengths or their locations has no
significant impact on the EM values for both readings.

49
Allways extensometer is mounted on the
Table 4-4 Modulus of elasticity for gauge length variation. middle of the test specimen.

Extensometer Readings

Force Force
Parameter
EA Strain rate Increase EM Strain rate Increase
Guage (GPa) for EA (1/s) rate for EA (GPa) for EM (1/s) rate for EM
location
(MPa/s) (MPa/s)

Upper Third 30.8 0.00025 7.7 20.88 0.0036 69.9

Middle 50.04 0.00025 12.5 17.37 0.004 65.3

Lower Third 31.83 0.00025 8.1 19.24 0.004 71.2

Crosshead Readings

Force Force
Parameter Crosshead Crosshead
EA Increase EM Increase
Speed Speed
(GPa) rate for EA (GPa) rate for EM
for EA (1/s) for EM (1/s)
(MPa/s) (MPa/s)

Upper Third 22.94 0.00033 7.7 10.79 0.0067 68.7

Middle 27.23 0.00046 12.5 10.03 0.0066 65.3

Lower Third 23.16 0.00034 8.1 11.08 0.0067 71.0

50
(a) (b)
Figure 4-4 Variation of EA and EM with gauge length location from (a) extensometer and
(b) crosshead readings.

4.1.2.2 Stresses

Moving on to the discussion of stresses, the same trend as observed in the previous group
persisted when considering varying gauge length locations. Fig 4-5 and Table 4-5 provided
insight into this analysis. There was no discernible impact on the stress values obtained from
either extensometer or crosshead readings when compared with each other. This suggests
that altering the gauge length locations near the grips does not substantially influence the
stress measurements. There is shloud be no relation between
location of extensometer and the stress
value which is Failure load divided by bar
Table 4-5 Stress values with variation of gauge location. cross section area.

Extensometer Readings (Mpa) Crosshead Readings (Mpa)


Parameter
Gauge σMs σMf σy σu σMs σMf σy σu
location

Upper
304.01 444.16 1198.87 1248.93 277.76 435.14 1189.07 1248.93
Third
Strain is
expected
to be maximum Middle 315.56 470.91 1117.63 1189.84 285.7 461.54 1117.83 1189.84
value at the
Middel
Lower
289.89 448.29 1192.4 1249.81 274.59 417.4 1183.23 1249.81
Third

51
(a) (b)
Figure 4-5 (Variation for Stress with gauge location from (a) extensometer and (b)
crosshead readings.

4.1.2.3 Strains

Finally, for strains, the findings remained consistent with previous observations regarding
gauge length variation. The differences in strain values obtained from two different curves
were particularly significant, especially when examining yield strain (Ԑy) and ultimate strain
(Ԑu).

The highest modulus of elasticity (EA) and initial stress (σMs) continued to yield the lowest
values for corresponding strains, including initial strain (Ԑs). Conversely, the group
characterized by the lowest modulus of elasticity (EM) and lower ultimate stress (σu)
produced the highest strain values for ultimate strain (Ԑu). These observations are detailed
in Fig 4-6 and Table 4-6.

52
Table 4-6 Variation for strain values with gauge location from extensometer and crosshead
readings.

Extensometer Readings (%) Crosshead Readings (%)


Parameter
Strain values
Gauge
expected to
location
Ԑs Ԑf Ԑy Ԑu Ԑs Ԑf Ԑy Ԑu
be the same
at lower and
upper third.
Also the Upper
highest value 0.91 5.22 8.84 17.07 1.05 6.82 13.8 27.82
expected to be
Third
at the Middel

Middle 0.61 4.9 8.67 18.06 1.02 6.88 13.43 29.51

Lower
0.82 5.26 9.13 17.27 1.06 6.6 13.51 26.91
Third

(a) (b)
Figure 4-6 Variation for strain values with gauge location from (a) extensometer and (b)
crosshead readings.

53
4.1.3 Influence of strain rate increase

4.1.3.1 Modulus of elasticity (EA)

The influence of increasing strain rates on the EA is described using Table 4-7 and Fig 4-7.
The impact of elevated strain rates was noticeable in both extensometer and crosshead
readings.

When the strain rate was increased from 2.5 x 10-4 to 2.5 x 10-3, a tenfold acceleration in
speed, the rate of force increase exhibited a remarkable surge of approximately 7.7 times its
initial value. Consequently, the EA, as determined from extensometer readings, decreased
to 42.49 GPa, compared to the initial value of 50 GPa obtained at a strain rate of 2.5 x 10-4.
In contrast, in the case of crosshead readings, where the crosshead speed increased by a
sevenfold factor, EA reached its highest value among all test parameters, measuring 31.34
GPa.

The substantially elevated EA in the crosshead readings for the higher strain rate group
suggests that the group with lower strain rate experienced higher slippage. This is evident as
the group with a 2.5 x 10-3 strain rate exhibited an EA of 42.49 GPa from extensometer
readings, and due to the high EA, it also yielded a high EA value of 31.34 GPa in crosshead
readings. Conversely, the group tested at a strain rate of 2.5 x 10-4, despite having the highest
EA according to extensometer readings at 50 GPa, produced a lower EA from crosshead
readings, measuring 27.23 GPa.

Efforts to push the strain rate to 2.5 x 10-2, a hundredfold increase, exceeded the machine's
operational capacity, causing it to operate only at 1.1 x 10-2. This exceptionally high speed
led to a decrease in stiffness, resulting in an EA of 30.5 GPa for extensometer readings and
23.15 GPa for crosshead readings.

In the case of EM (Modulus of Elasticity for Martensite), despite attempts to increase


crosshead speed from 6.7 x 10-3 to 5 x 10-2, the machine's limitations led it to operate at
2.6 x 10-2 for Group 5 and 3.5 x 10-2 for Group 6. However, these variations had no impact
on EM values, which remained consistent at around 10 GPa for crosshead readings and
ranged between 17-19 GPa for extensometer readings.

54
Table 4-7 Modulus of Elasticity for strain rate variation.

Extensometer Readings

Force Force
Parameter Strain rate Strain rate
Strain rate EA Increase EM Increase
for EA for EM
(GPa) rate for EA (GPa) rate for EM
(1/s) (1/s)
(MPa/s) (MPa/s)

The highest 0.00025 (1/s) 50.04 0.00025 12.5 17.37 0.004 65.3
EA value

0.0025 (1/s) 42.49 0.0024 96.4 19.12 0.014 265.6

0.025 (1/s) 30.5 0.011 320.4 19.02 0.02 364.3

Crosshead Readings

Crosshead Force Crosshead Force


Parameter
EA Speed Increase EM Speed Increase
(GPa) for EA rate for EA (GPa) for EM rate for EM
(1/s) (MPa/s) (1/s) (MPa/s)

0.00025 (1/s) 27.23 0.00046 12.5 10.03 0.0066 65.3

0.0025 (1/s) 31.34 0.0032 96.4 10.52 0.026 265.6

0.025 (1/s) 23.15 0.015 320.4 10.94 0.035 364.3

55
(a) (b)
Figure 4-7 Variation of EA and EM with strain rate from (a) extensometer and (b)
crosshead readings.

4.1.3.2 Stresses

Table 4-8 and Fig 4-8 demonstrated that varying strain rates had no discernible effect on
stress values for both readings. The only notable trend observed was a slight increase in the
values of σy and σu with higher strain rates.

Table 4-8 Variation of stress values with strain rate from extensometer and crosshead
readings.

Extensometer Readings (MPa) Crosshead Readings (MPa)


Parameter
Strain rate σMs σMf σy σu σMs σMf σy σu

0.00025
315.56 470.91 1117.63 1189.84 285.7 461.54 1117.83 1189.84
(1/s)

0.0025 (1/s) 324.86 505.64 1164.1 1218.05 290.98 496.43 1158.93 1218.05

0.025 (1/s) 324.56 446.98 1172.8 1240.08 309.25 442.73 1173.57 1240.08

56
(a) (b)
Figure 4-8 Variation of stress values with strain rate from (a) extensometer and (b)
crosshead readings.

4.1.3.3 Strains

Examination of strain readings in Table 4-9 and Fig 4-9, similar to the previous two
parameters, revealed significant differences when comparing strain readings from
extensometers and crosshead measurements. The disparities were particularly pronounced
when examining yield strain (Ԑy) and ultimate strain (Ԑu). Notably, the ultimate strain (Ԑu)
from both curves for the highest strain rate exhibited the lowest value.

Table 4-9 Variation of strains with strain rate from extensometer and crosshead readings.

Extensometer Readings (%) Crosshead Readings (%)


Parameter
Strain rate
Ԑs Ԑf Ԑy Ԑu Ԑs Ԑf Ԑy Ԑu

0.00025
0.61 4.9 8.67 18.06 1.02 6.88 13.43 29.51
(1/s)

0.0025 (1/s) 0.72 5.23 8.68 17.95 0.83 7.06 13.36 26.97

0.025 (1/s) 0.97 5.35 9.17 14.89 1.21 6.64 13.34 23.28

57
(a) (b)
Figure 4-9 Variation of strains with strain rate from (a) extensometer and (b) crosshead
readings.

4.1.4 Influence of stress rate transition

4.1.4.1 Modulus of elasticity

Table 4-10, in conjunction with Fig 4-10, shows the transition from strain rate control to
force rate control, ranging from 3 to 30 MPa/s, had a distinct influence on the EA for both
the extensometer and crosshead readings. This shift resulted in a notable reduction in EA
values, marking the lowest point across all the parameters for both reading with
extensometer readings ranged from 17 to 19 GPa, while crosshead readings remained nearly
constant at approximately 13 GPa.

An intriguing observation emerged with extensometer readings, as for the first time, the EM
surpassed EA by a considerable margin, reaching almost 23 GPa. This value represented the
highest EM reading across all parameters for extensometer measurements. In contrast, EM
values from crosshead readings remained lower than EA readings due to slip, hovering
around 10 to 11 GPa.

It's worth highlighting that a similar observation to that of Group 5 and 6 was noted in Group
7 and 8. After the transition to crosshead speed control at 6.7 x 10-3 following yielding, both
extensometer and crosshead readings yielded identical EM values, emphasizing the
consistent nature of this outcome.

58
Table 4-10 Modulus of Elasticity for stress rate variation.

Extensometer Readings

Force Force
Parameter Strain rate Strain rate
EA Increase EM Increase
Stress rate for EA for EM
(GPa) rate for EA (GPa) rate for EM
(1/s) (1/s)
(MPa/s) (MPa/s)

0.00025
50.04 0.00025 12.5 17.37 0.004 65.3
(1/s)

3 (Mpa) 17.21 0.00011 1.9 23.66 0.0034 77.6

30 (Mpa) 19.24 0.00081 15.7 23.24 0.003 66.2

Crosshead Readings

Crosshead Force Crosshead Force


Parameter
EA Speed Increase EM Speed Increase
(Gpa) for EA rate for EA (Gpa) for EM rate for EM
(1/s) (MPa/s) (1/s) (MPa/s)

0.00025
27.23 0.00046 12.5 10.03 0.0066 65.3
(1/s)

3 (Mpa) 13.81 0.00013 1.9 11.85 0.0067 76.8

30 (Mpa) 13.27 0.0013 17.6 10.6 0.0067 69.5

59
(a) (b)
Figure 4-10 Variation of EA and EM with stress rate from (a) extensometer and (b)
crosshead readings.

4.1.4.2 Stresses

Table 4-11 and Fig 4-11 illustrate that increasing the stress rate from 3 to 30 MPa/s led to a
general increase in stress values for σy and σu for both readings. Furthermore, the ultimate
stress control rate yielded the highest ultimate stress (σu) among all parameters reaching to
1308 MPa.

Table 4-11 Variation of stress with stress rate from extensometer and crosshead readings.

Extensometer Readings (MPa) Crosshead Readings (MPa)


Parameter
Stress rate
variation σMs σMf σy σu σMs σMf σy σu

0.00025
315.56 470.91 1117.63 1189.84 285.7 461.54 1117.83 1189.84
(1/s)

3 (Mpa) 273.63 373.12 1225.17 1287.31 253.13 371.8 1214.6 1287.31

30 (Mpa) 330.85 420.27 1260.43 1308.25 308.95 405.13 1235.23 1308.25

60
(a) (b)
Figure 4-11 Variation of stress with stress rate from (a) extensometer and (b) crosshead
readings.

4.1.4.3 Strains

Regarding strains, Table 4-12 and Fig 4-12 show that the transition from strain rate to force
rate control, ranging from 3 to 30 MPa/s, maximized Ԑs for both readings, resulting for the
lowest EA values and minimized Ԑu for extensometer readings, resulting for the maximum
σu values. Notably, significant differences were observed when comparing strain readings
from extensometers with crosshead measurements, particularly concerning yield strain (Ԑy)
and ultimate strain (Ԑu).

An important observation was that increasing the stress rate to 30 MPa/s resulted in more
slip, evidenced by the Ԑu. Although Group 8 and Group 7 yielded the same Ԑu in
extensometer readings, Group 8 achieved a higher Ԑu for crosshead readings, indicating a
5% difference between the two groups. Additionally, although both groups maintained the
same crosshead speed rate after the transformation into Martensite at 6.7 x 10-3, Group 8
required a lower force increase rate, registering 69.5 MPa/s compared to Group 7's 76.8
MPa/s.

61
Table 4-12 Variation for strains with stress rate from extensometer and crosshead readings.

Extensometer Readings (%) Crosshead Readings (%)


Parameter

Stress rate Ԑs Ԑf Ԑy Ԑu Ԑs Ԑf Ԑy Ԑu

0.00025 (1/s) 0.61 4.9 8.67 18.06 1.02 6.88 13.43 29.51

3 (Mpa) 1.46 5.45 9.05 13.79 1.59 6.65 13.77 24.1

30 (Mpa) 1.46 5.23 8.84 14.89 2.06 7.11 14.97 29.42

Extensometer Readings Extensometer Readings


35 35
Ԑs Ԑs
30 Ԑf 30 Ԑf
Ԑy Ԑy
Strain (%)

25
Strain (%)

25 Ԑu Ԑu
20 20
15 15
10 10
5 5
0 0
0.00025 (1/S) 3 (Mpa) 30 (Mpa) 0.00025 (1/S) 3 (Mpa) 30 (Mpa)

(a) (b)
Figure 4-12 Variation of strains with stress rate from (a) extensometer and (b) crosshead
readings.

62
4.1.5 Influence of gripping pressure

4.1.5.1 Modulus of Elasticity

Gripping pressure emerged as the most crucial parameter. Increasing the gripping pressure
from 40 bar to 80 bar and then to 120 bar had a noticeable impact on the EA and EM, as
shown in Table 4-13 and Fig 4-13.

The trend reveals that raising the gripping pressure decreased EA values and increased EM
values for both readings. It also reduced the disparity between extensometer and crosshead
readings for the same parameter. For instance, with 80 bar gripping pressure in Group 9, the
difference between EA values from extensometer and crosshead readings was only about 2.5
GPa. This suggests that increasing gripping pressure reduces the reading differences,
indicating a contribution to slip in the previous parameter, even within the elastic region.

Concerning EM, while minimizing the difference between extensometer and crosshead
readings for the same parameter, it achieved the highest-ever force increase rate of 97.4 GPa
to match the previously adapted crosshead rate of 6.7 x 10-3. This led to EM reaching its
maximum value, recording 15.1 GPa for crosshead readings, although a slight degree of slip
still persisted.

63
Table 4-13 Variation of modulus of elasticity with gripping pressure.

Extensometer Readings

Force Force
Parameter Strain rate Strain rate
EA Increase EM Increase
Gripping for EA for EM
pressure (GPa) rate for EA (GPa) rate for EM
(1/s) (1/s)
(MPa/s) (MPa/s)

40 (bar) 50.04 0.00025 12.5 17.37 0.0040 65.3

80 (bar) 22.28 0.00019 4.4 19.1 0.0044 79.6

120 (bar) 20.57 0.00025 5.1 21.49 0.0047 95.7

Crosshead Readings

Crosshead Force Crosshead Force


Parameter
EA Speed Increase EM Speed Increase
(Gpa) for EA rate for EA (Gpa) for EM rate for EM
(1/s) (MPa/s) (1/s) (MPa/s)

40 (bar) 27.23 0.00046 12.5 10.03 0.0066 65.3

80 (bar) 19.78 0.00022 4.4 12.34 0.0067 78.7

120 (bar) 16.13 0.00031 5.1 15.1 0.0067 97.4

64
(a) (b)
Figure 4-13 Variation of (EA and EM with gripping pressure from (a) extensometer and
(b) crosshead readings.
When we use extensometer, there should not be change in extensometer reading due gripping pressure

4.1.5.2 Stresses

Table 4-14 and Fig 4-14 reveal that an increase in gripping pressure led to stress values from
both curves becoming more similar for every parameter. This increase in gripping pressure
also resulted in the smallest σMs values among all previous parameters. The lower
nucleation stress for TB compared to the high σMs values from previous parameters was a
consequence of the contribution to slippage.

Table 4-14 Variation of stress with gripping pressure from (a) extensometer and (b)
crosshead readings y. It is expected to get similar values between extensometer reading and crosshead
readings by increasing griping pressure.

Extensometer Readings (MPa) Crosshead Readings (MPa)


Parameter
Gripping σMs σMf σy σu σMs σMf σy σu
pressure

Why? 40 (bar) 315.56 470.91 1117.63 1189.84 285.7 461.54 1117.83 1189.84

80 (bar) 254.74 414.9 1108.25 1148.08 231.34 422.09 1101.35 1148.08

120 (bar) 251.32 374.48 1134.2 1181.11 243.84 364.78 1125.97 1181.11

65
(a) (b)
Figure 4-14 Variation of stress with gripping pressure from (a) extensometer and (b)
crosshead readings.

4.1.5.3 Strains

In Table 4-15 and Fig 4-15, the impact of increasing gripping pressure on strain values
obtained from both readings is evident, particularly for Ԑy and Ԑu. The percentage
differences for Ԑy decreased from 4.76 to 3.23 and then to 2.33, while Ԑu decreased from
11.45 to 6.34 and then to 4.88.

Notably, in Fig 4-15 (a), Group 9 yielded the maximum Ԑu for extensometer readings,
whereas Group 0 achieved the highest Ԑu for crosshead readings. Group 0's Ԑu matched that
of Group 10 at approximately 18% for crosshead readings. However, due to higher slip,
Group 0 presented a falsely elevated maximum Ԑu in crosshead readings. Group 10, with the
highest gripping pressure, attained the lowest ever (Ԑu) for crosshead readings, primarily due
to effective slip reduction in the grips.

66
Table 4-15 Variation of strain with gripping pressure from extensometer and crosshead
readings.

Extensometer Readings (%) Crosshead Readings (%)


Parameter
Ԑs Ԑf Ԑy Ԑu Ԑs Ԑf Ԑy Ԑu

40 (bar) 0.61 4.9 8.67 18.06 1.02 6.88 13.43 29.51

80 (bar) 0.71 5.5 9.13 20.8 0.98 6.85 12.36 27.14

It is logic 120 (bar) 1.27 5.76 9.33 17.23 1.5 6.61 11.66 22.11

(a) (b)
Figure 4-15 Variation of strain with gripping pressure from (a) extensometer and (b)
crosshead readings.

67
4.1.6 Influence of temperature variation

4.1.6.1 General Discussion

In general, as depicted in Fig. 4-16 (a), which illustrates the differences between Group 2 for
both extensometer and crosshead readings, it becomes apparent that, even with the
extensometer covering approximately 70% of the entire specimen length, variations between
the two curves emerge early in the test and persist throughout the transformation into the
martensite region. At any stress value, the extensometer readings consistently yield lower
strain values than the crosshead readings. This discrepancy arises due to the tendency of
SMAs to form Transformation Bands (TB) during the transformation into Martensite.
Nucleation of TB occurs near the grips and is not captured by the extensometer unless the
TB expansion is considered within the gauge length [23, 24]. The differences become more
pronounced after the completion of the Martensite transformation until failure. Several
factors contribute to this variation, including slip at grips, a phenomenon observed due to
the continuous extension of TB within the grips during Martensite transformation. This
action rapidly diminishes the local cross-sectional area, causing slip and concentrating
excessive stress at the ends of the wire specimen, leading to premature failure near or inside
the grips [14]. Another contributing factor is the non-uniform strain distribution in the form
of Luders-like bands [22] observed in Ni-Ti wires immediately after yielding, even when
TB expands through the entire length of the specimen.

In contrast, Fig. 4-16 (b) for group 11, reveals that, from the beginning of the test until the
end of the martensite transformation region, the extensometer readings consistently yield
larger strain values than the crosshead readings at any stress value. This indicates a stiffening
effect for the crosshead compared to the extensometer. This observation suggests that most
TB nucleation occurs in the extensometer region at a lower temperature than the rest of the
specimen, contrary to the usual occurrence near grips. This concentration of deformation in
this region initiates martensite transformation, propagating to the entire length of the
specimen.

After transforming all specimens into martensite until failure, Group 11 exhibits a similar
trend to Group 2. At any stress value, extensometer readings yield lower strain values than
crosshead readings for the same reasons mentioned earlier. However, it's important to note
that these differences are smaller, with slight variations between the extensometer and
crosshead readings compared to Group 2.

68
Comparing extensometer readings for both groups, Fig. 4-17 (a) presents identical stress-
strain curves with nearly the same slopes for every region, except the elastic region, resulting
in lower modulus of elasticity for Group 11 than Group 2. Consequently, Group 11 exhibits
lower stress values than Group 2.

Fig. 4-17 (b) compares crosshead readings for both groups, illustrating identical curves.
However, Group 11 shows the first transformation into martensite at a lower value than
Group 2, and the initial plastic yielding is also at a lower value.

For a comprehensive comparison of the differences between extensometer and crosshead


readings for each group and between each group for both readings, the next section provides
an accurate comparison using the exact mechanical properties values derived from the
tangent lines used for stress-strain curves obtained from extensometer and crosshead
readings.

Group (2) Group (11)

(a) (b)
Figure 4-16 Stress vs. both extensometer strain and crosshead strain curves for (a)
Group (2) and (b) Group (11).

69
Group (11) Group (11)
Group (2) Group (2)

(a) (b)
Figure 4-17 Stress vs strain curves for both Group (2) and Group (11) from (a)
extensometer and (b) crosshead readings.

4.1.6.2 Modulus of elasticity

Table 4-16 and Fig 4-18 highlight significant differences in the (EA) and (EM) between
Group 2 and Group 11. For Group 2, at the same stress increase rate of 6 MPa/s, it is observed
that the strain increase rate at the extensometer region (2.5x10-4 1/s) is lower than the strain
increase rate obtained from crosshead readings throughout the entire length (3.5x10-4 1/s).
This results in a larger value for EA derived from extensometer readings (23.76 GPa)
compared to the 17 GPa extracted from crosshead readings. After the transformation into
Martensite, the EM value for extensometer readings (21.22 GPa) nearly doubles the EM
from crosshead readings (10.84 GPa). The significant differences between EM derived from
extensometer readings and EM extracted from crosshead readings are attributed to the
reasons previously mentioned, such as slip at grips and non-uniform strain distribution.

For Group 11, the trend for EA is opposite. At a very low force increase rate of 2.1 MPa/s,
it exhibits a higher strain increase rate at the extensometer region (2.4x10-4) than at the
crosshead region (2.2x10-4), resulting in EA from extensometer readings (9 GPa) lower than
the 9.79 GPa obtained from crosshead readings. Conversely, for EM, it follows the same
trend as Group 2, with a nearly doubled value for EM from extensometer readings (26.98
GPa) compared to the EM from crosshead readings (14.74 GPa). Importantly, Group 11
yields an EM much higher than its EA value with a tripled value for extensometer readings,
and EM remains greater than its EA value for crosshead readings, albeit with smaller
differences. This contrasts with the observed trend in Group 2, where EA is higher than its
EM value for both readings.

70
Comparing Group 2 with Group 11, it is evident that EA values for Group 2 are larger than
EA values for Group 11 for both readings. However, after transforming into Martensite,
Group 11 exhibits higher values for EM than Group 2 for both readings. This observation
may be attributed to the overall stiffness of the material appearing as a fixed value divided
into almost two identical values for Group 2, as seen in its EA and EM values. In contrast,
for Group 11, this value is divided into a very small value as EA, while the rest becomes a
much higher value as EM.

Table 4-16 Modulus of Elasticity variation between group (2) and group (11).

Extensometer Readings

Strain rate Stress increase Strain rate Force increase


EA EM
Parameter for EA rate for EA for EM rate for EM
(Gpa) (GPa)
(1/s) (MPa/s) (1/s) (MPa/s)

Group (2) 23.76 0.00025 6.0 21.22 0.00345 69.2

Group (11) 9.00 0.00024 2.1 26.98 0.0035 88.2

Crosshead Readings

Crosshead Stress increase Crosshead Force increase


EA EM
Parameter speed for rate for EA speed for rate for EM
(Gpa) (GPa)
EA (1/s) (MPa/s) EM (1/s) (MPa/s)

Group (2) 17 0.00035 6.0 10.84 0.0067 69.2

Group (11) 9.79 0.00022 2.1 14.74 0.0067 94.6

71
Figure 4-18 Variation of EA and EM between group 2 and group 11 from extensometer
and crosshead readings.

4.1.6.3 Stresses

Fig 4-19 provides insights into the variations in stress values for both extensometer and
crosshead readings between Group 2 and Group 11. For Group 2, where EA values obtained
from extensometer are higher than those from the crosshead, this results in higher values for
start stress (σMs) and finish stress (σMf) from extensometer readings compared to crosshead
readings. This suggests that the initiation and completion of the transformation into
Martensite occurred primarily outside the extensometer region. Additionally, Group 2
exhibits very close values for yield stress (σy) and ultimate stress (σu) for both readings.

In contrast, Group 11, with EA values from extensometer lower than those from the
crosshead, yields lower values for start stress (σMs) and finish stress (σMf) from extensometer
readings compared to crosshead readings. This indicates that the initiation and completion
of the transformation into Martensite occurred primarily inside the extensometer region.
Similar to Group 2, Group 11 demonstrates a consistent trend for yield stress (σy) and
ultimate stress (σu) with very similar values for both readings.

When comparing stress values between Group 2 and Group 11, the lower EA value for
Group 1 results in lower stresses for all parameters in both readings. Notably, the differences
between each stress value in Group 2 and its corresponding value in Group 11 remain
consistently similar, with approximately 130 MPa for extensometer readings and around 100
MPa for crosshead readings.

72
Figure 4-19 Variation for stress values between Group 2 and Group 11 from extensometer
(Ex) and crosshead (CH) readings.

4.1.6.4 Strains

Fig. 4-20 offers a comprehensive comparison of strain values between Group 2 and Group
11, considering data obtained from both extensometer and crosshead readings. For Group 2,
characterized by higher values of EA and EM in extensometer readings compared to
crosshead readings, it results in lower values for all strain parameters in extensometer data.
The differences between the two readings are slight for (Ԑs) and (Ԑf), becoming higher for
(Ԑy) and significantly higher for (Ԑu).

In Group 11, a similar trend is observed, with lower values for all strain parameters in
extensometer readings compared to crosshead readings. Except (Ԑs) in Group 11, which
yields a higher value for extensometer reading than crosshead reading due to the lower EA
for extensometer than crosshead. The variations between the values for both readings are
lower than in Group 2, showcasing smaller differences.

Comparing Group 2 with Group 11 reveals very slight differences between any strain value
from Group 2 and its adjacent value from Group 11 for extensometer readings, not exceeding
a 0.7% difference between the two readings. In contrast, although crosshead readings exhibit
slight differences for (Ԑs) and (Ԑf), these differences become more pronounced for (Ԑy) and
substantially higher for (Ԑu). Notably, while both groups show similar extensometer (Ԑu)
values at approximately 15%, Group 2 demonstrates a higher crosshead (Ԑu) at 27.33%
compared to Group 11's 21.96%, indicating a higher degree of slip in Group 2.

73
Figure 4-20 Variation for strain values between Group 2 and Group 11 from extensometer
(Ex) and crosshead (CH) readings.

4.2 Post-Fracture Elongation Characteristics


The objective was to evaluate the recovery strain in SMAs following a failure and understand
how various parameters influence it. This assessment involved using three pristine
specimens in each group. Following their failure, these specimens underwent a
transformation into a plastic state. The methodology entailed aligning the halves of the
fractured specimens and manually calculating the difference between the machine-recorded
extensometer failure strain (Ԑfr) % and the plastic retained strain. The plastic retained strain
was measured by examining the gauge marks spaced at 1 cm intervals along the entire length
of the specimen. This approach allowed to determine the recovery strain within the
extensometer location.

Additionally, the crosshead recovery strain throughout the entire specimen length between
the grips after fracture was calculated. This calculation involved finding the difference
between the crosshead failure strain (Ԑfr) % and the plastic retained strain measured from
the gauge marks within the entire length of the specimen.

Furthermore, the distribution of the plastic retained strain along the entire specimen, from
grip to grip, was compared and examined how this distribution interacted with various
parameters. Measurements were taken starting from the failure location and moved
incrementally 1 cm away for every specimen in every group, using the failure location as a
reference point. This process allowed to comprehensively investigate how SMAs perform
post-failure and how different factors influence the recovery strain.

74
4.2.1 Influence of gauge length variations
In Fig 4-21 (a), it is evident that altering the gauge length has no impact on the extensometer,
and the crosshead recovery strains after failure remains (nearly remain constant at around
12.5% and approximately 24%, respectively). The extensometer recovery strain represents
the true value of SMA performance after failure, while the variation in crosshead recovery
strain arises due to the amount of slip at the grips. Since the machine operates with a
consistent gripping pressure of 40 bar, the slip is uniform across all specimens with the same
gripping pressure, resulting in consistent recovery strain.

Fig 4-21 (b) demonstrates that the distribution of plastic retained strain (%) along the entire
length for Group 0 is well-distributed, with the maximum value at the failure location and a
gradual decrease toward the grips. In contrast, Group 1 and Group 2 exhibit less uniform
distribution with the maximum value at the failure location, and a sharper decrease nearby,
especially in the case of Group 1. This uneven distribution persists despite the expansion of
all transformation bands (TBs) along the specimen's length during the transformation into
Martensite and the elimination of strain localization effects. Uneven strain distribution is
still observed along the specimen's entire length, particularly near the movable grip, where
higher strain values are measured compared to the rest of the specimen.

(a) (b)
Figure 4-21 Influence of gauge length on (a) recovery strain and (b) retained strain
distribution along the specimen's entire length.

75
4.2.2 Influence of gauge length location
It is evident from Fig 4-22 (a) that the variations in gauge length location do not significantly
affect the recovery strain values. The values are consistent across different locations, ranging
from 11% to 13% for extensometer reading and an identical 24% for crosshead recovery
strain.

The minor differences in extensometer recovery strain values are due to the proximity of the
extensometer location to the failure location. A closer extensometer location results in more
plastic retained strain and less recovery strain, and vice versa.

Fig 4-22 (b) reveals that varying the gauge length location does not affect the uniformity of
plastic strain distribution along the entire length of the specimen, except for Group 0 in the
middle gauge location, which shows the highest values, followed by Group 4 and Group 3.

(a) (b)
Figure 4-22 Influence of gauge length location on (a) recovery strain and (b) retained strain
distribution along the specimen's entire length.

4.2.3 Influence of strain rate increase


Fig 4-23 (a) shows that the extensometer recovery strain remains consistent across different
groups at around 13%. When comparing crosshead recovery strain, it is noticed that an
increase in strain rate results in a lower value. This may seem counterintuitive, as higher
strain rates usually result in less retained strain, which would suggest more recovery strain
[28]. However, this effect occurs only when cycles of loading and unloading are performed
with the same rate, leading to shorter unloading times and elevated specimen temperatures,
resulting in more recovery strain. It is crucial to note that all recovery strain measurements
are taken after failure without undergoing loading and unloading cycles. Additionally, the

76
specimens reach nearly identical temperatures at failure, as explained in section 3.3. The
reduction in recovery strain with increasing strain rate can be attributed to the reduction in
slip, as observed in Group 5, which exhibits a higher extensometer failure strain than Group
0. However, Group 0 has more crosshead failure strain, indicating greater slip, consistent
with earlier findings.

Fig 4-23 (b) demonstrates that increasing strain rates lead to non-uniform distribution. The
plastic strain increases in the failure region and decreases significantly as we move further
away from the failure location toward the distant grip. This is particularly evident in Group
6, which experiences the highest strain rate and exhibits almost zero plastic strain near the
distant grip.

(a) (b)
Figure 4-23 Influence of strain rate increase on (a) recovery strain and (b) retained strain
distribution along the specimen's entire length.

4.2.4 Influence of stress rate transition


In Fig 4-24 (a), the shift from a strain rate-controlled process to a stress rate-controlled
process yields results that closely resemble those of the previous groups. The extensometer
strain remains at approximately 12%. However, there is variability in the crosshead recovery
strain, ranging from 22% to 27%. This variability can be attributed to the slip phenomenon
as previously discussed. Notably, Group 8 exhibits a lower extensometer failure strain
compared to Group 0. Yet, it demonstrates a higher crosshead failure strain, indicating that
the transition to a high stress rate introduces more slip than that already present in Group 0,
as explained earlier.

77
Fig 4-24 (b) illustrates how the shift from a strain rate to a stress rate control significantly
disrupts the uniformity of strain distribution. Group 7, characterized by a very low stress rate
of 3 MPa/s, displays the highest plastic retained strain at the failure location, with elevated
values near the nearest grip. However, in regions close to the distant grip, the plastic strain
drops nearly to zero. This outcome underscores that relying on an extremely low stress rate,
as opposed to a strain rate, leads to the formation of regions with very high strains and
regions with very low strains. Importantly, none of these regions exceed the elastic strain
limit. In contrast, Group 8, subjected to a high stress rate, exhibits a more uniform
distribution of plastic strain, albeit with lower values compared to Group 0.

(a) (b)
Figure 4-24 Influence of stress rate transition on (a) recovery strain and (b) retained strain
distribution along the specimen's entire length.

4.2.5 Influence of gripping pressure increase


As evident from Fig 4-25 (a), increasing the gripping pressure has no effect on the value of
the recovery strain at the extensometer location, which remains nearly constant at around
12%. However, it significantly decreases the crosshead recovery strain. As gripping pressure
increases from 40 to 80 to 120 bar, the crosshead recovery strain decreases from 24% to 20%
to 16%. This suggests that increasing gripping pressure reduces slip. It appears that further
increasing the gripping pressure might make the crosshead recovery strain match the
extensometer value, as SMAs should yield the same recovery strain value for all measuring
methods. This is supported by the observation that Group 9 exhibits more extensometer
strain at failure than Group 0, and Group 10 yields exactly the same value as Group 0.
However, surprisingly, Group 0 has the highest crosshead strain at failure.

78
Fig 4-25 (b) highlights another important aspect: increasing gripping pressure almost
doubles the plastic strain at the failure location for Group 9 and Group 10 compared to Group
0, resulting in the formation of a necking region for the first time, which will be further
discussed in section 3.4. Additionally, increasing gripping pressure promotes more uniform
distribution along the entire length of the specimen.

(a) (b)
Figure 4-25 Influence of gripping pressure increase on (a) recovery strain and (b) retained
strain distribution along the specimen's entire length.

4.2.6 Influence of temperature variation


Fig. 4-26 illustrates that for extensometer readings, group 11 has a plastic retained strain
value at the extensometer region higher than the value obtained from group 2, resulting in
more extensometer failure strain with almost 18.4% for group 11 than 15.8% for group 2.
Despite this variation, both groups yield almost the same value for recovery strain, nearly
12.4%, which is a consistent value for SMAs of the exact material. The occurrence of
fractures near the grips outside the extensometer region for most specimens results in more
plastic retained strain throughout the entire length of the specimen, as shown in groups 2 and
11. This should have resulted in lower values for the recovery strain obtained from the entire
length of the specimen through crosshead readings than extensometer readings. However,
due to the presence of high slip at grips, it gives falsely higher values for recovery strain for
both groups from cross head readings.

It is observable that group 11, which yields more plastic retained strain, whether in the
extensometer region or throughout the entire length of the specimen than group 2, also

79
yielded more extensometer failure strain. Although group 2 yielded higher crosshead failure
strain, indicating higher slip than already presented in group 11.

Fig. 4-27 illustrates the plastic retained strain (%) distribution throughout the entire length
of the specimen. Both groups (2) and (11) show the same trend with the maximum value at
the failure location and a gradual decrease toward the grips. Group 11 yielded double plastic
retained strain at the failure location with 31.7% compared to 15.5% for group 2. It also
yielded higher values for the rest of the specimen than group 2. The main reason for this is
the formation of necking before failure, giving group 11 more ductility than group 2, as
explained in detail in analysis of failure characteristics.

Figure 4-26 The variation in both plastic retained strain and resulting recovery strain
between Group 2 and Group 11, from extensometer and crosshead readings.

Figure 4-27 The distribution of plastic strain between Group 0 and Group 1 throughout the
entire length of the specimen, with the failure location serving as the reference point.

80
4.3 Temperature Fluctuations Assessment
The characterization process during the loading of specimens in the austenitic phase
triggers a forward transformation into martensite, which is an exothermic process. This
transition introduces the latent heat phenomenon and consequently leads to temperature
fluctuations during the testing procedure.

The stress-induced martensitic (SIM) phase transformation is responsible for the sudden
generation of substantial strain, resulting in significant stress drops. When specimens are
subjected to quasi-static tensile loading, the SIM phase transformation occurs through the
formation of macroscopically observable domains known as transformation bands (TBs).
The development of these TBs is closely intertwined with heat transfer processes within
the material and its surrounding environment. The amount of latent heat generated depends
on factors such as the number of active transformation bands, their propagation fronts, and
the speed of these fronts [16, 23, 25, 35].

As observed in the stress-strain plots shown in Fig 3-3 (a), the formation of several TBs
becomes evident during the stress plateau region. The initiation of a new TB is typically
signified by a stress drop in the stress-strain plots. Notably, the magnitude of the stress drop
correlates with the amount of strain produced and the latent heat released.

4.3.1 Stress Temperature Curves


The stress-temperature plots for the three thermocouples, depicted in Fig 4-28 for four
different groups with no temperature variation, generally exhibit patterns similar to the
stress-strain plots, with each curve displaying four distinct regions. The first region
commences at the beginning of the test and continues until reaching the start stress (σMs),
with no significant increase in specimen temperature. The second region spans from σMs to
the finish stress (σMf), where the SIM phase transformation occurs sharply, resulting in a
notable temperature rise, while stress remains nearly constant. The third region extends from
σMf to the point of yield stress (σy), characterized by a gradual increase in specimen
temperature along with an increase in stress. The fourth region encompasses the stress range
from σy to the ultimate stress (σu) and is characterized by a slight simultaneous increase in
temperature and stress.

It is important to note that transitions between these regions are signaled by changes in the
slope of the curve. Furthermore, variations in the locations of the three thermocouples can
lead to differences in slope changes among them.
81
Figure 4-28 Stress-temperature curves with thermocouples (T1, T2, and T3) placed in the
upper third, middle, and lower third of the specimens.
4.3.2 Rise in Temperature Strain Curves
For assessing the temperature variation for group 2 with uniform temperature than group 11
with temperature variation, rise in temperature-strain curves were plotted. Due to the
significant impact of the failure location on the temperature variation measured by the
nearest thermocouple on the rise in temperature-strain curves, individual specimens were
selected from Group 2 and Group 11, with similar failure locations, for detailed temperature
analysis. A specimen from Group 2 with a failure near the bottom grip near T3 in the lower
third was chosen and compared with a specimen from Group 11 with a similar failure
location.

Fig. 4-29 and Fig. 4-30 illustrate the temperature-extensometer strain curves derived from
monotonic tensile tests for SMAs. The initial point, where the temperature experiences its
first increase, can be identified as the strain start (Ԑs*). This signifies the commencement of
the forward transformation into martensite. The temperature continues to rise until a point is
reached where the rate of temperature increase becomes significantly lower. Using tangent

82
lines to determine this point, it indicates the strain finish (Ԑf *), marking the completion of
the transformation into the martensite phase [44].

In Fig. 4-29, representing Group 2, the sudden increase in temperature for all three
thermocouples commenced at approximately 0.7% strain, which is lower than the concluded
start strain value (Ԑs) from the stress-extensometer strain curve (1.12%). The temperature
rise continued until 7.5% for T1 and T3, and 8.8% for T2, marking their (Ԑf *). This
confirmed the completion of forward transformation, first observed in the upper and lower
thirds near the grips before reaching the middle region where T2 was attached. The average
value for (Ԑf *) for the three thermocouples was equal to 7.9%, closely matched the (Ԑf) from
the stress-crosshead strain curve (6.98%).

The overall rise in temperature for T1, T2, and T3 until their (Ԑf *) values was approximately
24°C. Afterward, the slopes for all thermocouples showed a gradual decrease toward failure,
except for T3, which exhibited a slightly higher slope due to the presence of failure near it.
The rise in temperature from (Ԑf *) to (Ԑu) was only 7°C for T1 and T2 and 10°C for T3.

In Fig. 4-30, representing Group 11, variations in temperature between regions resulted in
differences for all values for the three thermocouples. The sudden increase in temperature
started first for T3 at 0.7%, then T2 at 1.4%, and finally T1 at 2.8%. This temperature rise
continued up to the completion of the forward transformation, reaching 4.6% for T2, 7.1%
for T3, and 8.4% for T1. This confirmed the opposite trend than group 0 with completion of
forward transformation, first observed in the middle region where T2 was attached before
reaching the upper and lower thirds near the grips. The average value for (Ԑs*) for the three
thermocouples (1.63%), closely matched the (Ԑs) indicated from the stress-extensometer
strain curve (1.64%). Similarly, the average value for (Ԑf *) for the three thermocouples was
equal to 6.7%, matching the (Ԑf) from the stress-crosshead strain curve (6.75%).

The rise in temperature until reaching (Ԑf *) was only 9°C for T2, 15°C for T3, and 27°C for
T1. Notably, T1 and T2, after reaching their (Ԑf *), exhibited a horizontal slope with no
further rise in temperature until failure. In contrast, T3 continued to show a continuous rise
in temperature, albeit at a lower slope, until failure, with a total rise from (Ԑf *) to (Ԑu) of
about 20°C, attributed to the presence of failure near it.

83
Figure 4-29 The rise in temperature with extensometer strain through T1, T2 and T3
Thermo-Couples for group (2).

Figure 4-30 The rise in temperature with extensometer strain through T1, T2 and T3
Thermo-Couples for group (11).

84
4.4 Analysis of failure characteristics
In the realm of tensile testing, specimens can manifest one of two distinct types of fracture:
ductile or brittle. These classifications are rooted in the degree of macroscopic plastic
deformation exhibited prior to fracture, thereby characterizing the material's response to
stress [45, 46].

Illustrated in Fig 4-32, all groups except for Group 9, Group10 and Group 11 in
predominantly display less ductile fractures. In these cases, fracture transpires without any
significant necking or extensive plastic deformation. The outcome is abrupt and catastrophic,
often resulting in a fracture surface that appears relatively flat and smooth, exhibiting
minimal to no signs of deformation preceding the fracture point.

Conversely in Fig 4-33, group 9, group 10 and group 11 deviate from this less ductility
behaviour due to the manipulation of gripping pressure as in groups 9 and 10 and temperature
variation as in group 11. In these groups, a ductile fracture is observed, characterized by the
occurrence of necking. The material continues to undergo plastic deformation, resulting in
thinning and elongation until it ultimately ruptures. The fracture surfaces in these cases bear
a distinctive cup-and-cone appearance, indicative of substantial plastic deformation.

An additional measure of ductility can be ascertained by examining the reduction in area.


This parameter is typically quantified by calculating the percentage change in cross-sectional
area before and after the tensile test, as outlined in ASTM E8 [20]. Fig 4-31 highlights that
group 9, group 10 and group 11 exhibit the highest reduction in area, approximately 50%,
40% and 50%, respectively. These significant reductions are indicative of pronounced
ductility. In contrast, the remaining groups, ranging from 20% to 30% reduction, reflect a
comparatively less ductile behaviour. This aspect underscores the relationship between
increasing gripping pressure and the material's ability to exhibit the necking phenomenon
associated with ductile behaviour, in addition, temperature variation for a part of the
specimen could make the specimen to show more ductile behaviour.

The presence and extent of slip within the material can be inferred from the grip impressions
on the specimens. These initial marks are typically imprinted when the specimen is secured
within the grips prior to testing. They are a result of localized compression and deformation
at the contact points between the specimen and the grips. Changes in the appearance of these
marks, particularly when they become less distinct or appear as elongated lines, suggest
movement or sliding of the specimen within the grips during the test. This movement is often
85
indicative of material deformation and can be associated with slip. Conversely, regions
where the marks remain unaltered suggest minimal deformation or slippage within the grips.

To further understand the phenomenon of grip impressions, a comprehensive analysis of the


gripping length inside the clamps is meticulously carried out for all specimen groups. While
all groups within the 0 to 8 range adhere to the consistent gripping pressure of 40 bar, the
illustrative comparison will primarily focus on Groups 3, 5, 7, and 8, which serve as
representative examples within this 40 bar range. Additionally, this analysis encompasses
Group 9, characterized by a doubling of gripping pressure to 80 bar, group 10, where the
gripping pressure triples to 120 bar, in addition, a comparation between group 2 with unifrom
temprature and group 11 with temperature variation ws done.

The insights gleaned from Fig 4-34 vividly depict the transformation of grip impressions in
Groups 3, 5, 7, and 8. These impressions have notably evolved from their initial distinct
appearance to take on the form of elongated lines, spanning across roughly 80% of the
gripping length in all specimens within these groups. In Group 9, where gripping pressure
experiences a twofold increase from 40 to 80 bar, the grip impressions similarly adopt the
elongated line pattern but now cover approximately 50% of the gripping length. In group 10,
where gripping pressure triples to 120 bar, the grip impressions manifest as lines over a
notably shorter region, encompassing about 30% of the gripping length. In this case, the
remaining portion of the grip impressions maintains their clear and unchanged
characteristics, exhibiting no significant alteration.

Also Fig 4-35 show the difference between group 2 and group 11, as the grip impressions
manifest as lines over a notably shorter region, encompassing about 30% of the gripping
length for group 11 than 80% for group 2.

These observations shed light on the multifaceted role of increased gripping pressure. Not
only does it play a pivotal role in promoting necking, which is indicative of ductile
behaviour, but it also serves to minimize slip, thus exerting a discernible influence on the
material's failure behaviour.

86
Figure 4-31 Reduction of area (%) distribution across groups.

Figure 4-32 Failure shapes in addition to the fracture surfaces for each specimen in every
group from group 0 to 8.

87
Group (11)

Figure 4-33 Failure shapes in addition to the fracture surfaces for each specimen in groups
9, 10 and 11. Note that Group (9) is based on only two specimens due to limited availability.

88
Please indicate the direction of specimen w.r.t crosshead movement direction.

Figure 4-34 Grip impressions observed within Groups 3, 5, 7, 8, 9, and 10.

89
Group (11)
Group (2)

Figure 4-35 Grip impressions observed within Group 2 and Group 11.

90
CHAPTER 5

SUMMARY AND CONCLUSIONS

91
CHAPTER 5
5 SUMMARY AND CONCLUSIONS
In this comprehensive study, a series of experimental tests was conducted to explore the
multifaceted behaviour of Shape Memory Alloys (SMAs) under a range of testing
conditions, including gauge length variation, gauge length location, strain rates, stress rates,
and gripping pressure. The insights gained from this research shed light on several critical
aspects of SMAs for civil engineering applications.

The following conclusions are drawn from the test results:

It is true even for  It is evident that relying solely on cross head readings without the inclusion of
steel rebar.
increased gripping pressure is insufficient. This inadequacy is primarily due to
the presence of slippage, even when pre-load is applied. Consequently, all the
conclusions derived from this study are based on the extensometer readings,
which offer a more accurate representation of the material's behaviour.
 Variation in the gauge length (20, 40 and 80 mm) significantly impacts elastic

What is the proper modulus (EA) values (30.86, 50 and 23.76 GPa, respectively). However, these
gauge length to get gauge length variations do not affect the elastic modulus (EM), critical stresses,
the correct value of EA?
corresponding strains, or the distribution of plastic retained strain.
 Modifying the gauge length location, whether in the upper or lower third near the
grips, results in nearly identical values for EA, EM, critical stresses, and
See Table 4-1 corresponding strains. In addition, it has no discernible impact on the uniformity
of plastic strain and uniformity of plastic strain distribution along the specimen's
length.
 Higher strain rates lead to lower EA values, minor variations in EM, an increase
in ultimate stress (σu), and a decrease in ultimate strain (Ԑu). They also induce
non-uniform plastic strain distribution, with a significant increase near the failure
region and a subsequent decrease toward the distant grip. Group 6 with the
highest strain rate, shows almost zero plastic strain near the distant grip.
 Transitioning from strain rate to force rate control results in the lowest EA values
(approximately 17 GPa to 19 GPa), with EM surpassing EA for the first time
(around 23.5 GPa). It also yields the highest ultimate stress (σu) of approximately
1300 GPa and the lowest ultimate strain (Ԑu) of around 14%.

92
 Transitioning to a low stress control rate of 3 MPa/s significantly disrupts the
uniformity of strain distribution. It results in the highest plastic retained strain at
the failure location and elevated values near the nearest grip. In regions close to
the distant grip, the plastic strain drops nearly to zero. Conversely, with an
increased stress rate of 30 MPa/s, a more uniform distribution of plastic strain is
observed.
 Increasing gripping pressure substantially reduces slippage, providing more
reliable values for EA and EM. It also leads to the precise determination of start
stress (σMs) and finish stress (σMf) during the transformation into Martensite. In
addition, reduces differences between mechanical properties obtained from
extensometer and crosshead readings with ultimate strain (Ԑu) decreases from
11.45 to 6.34 and then to 4.88 for 40, 80, and 120 bar gripping pressure.
 Higher gripping pressure enhances plastic strain distribution uniformity along the
specimen length, resulting in substantial plastic strain retention (up to 30%),
promoting ductile behaviour, and a significant reduction in area (40-50%).
Conversely, lower gripping pressure leads to brittle fractures with minimal
plastic deformation retain (up to 15%) and smaller area reduction (20-30%). In
addition, increasing gripping pressure also reduces grip impressions and
slippage, with elongated lines diminishing from 80% (40 bar) to 50% (80 bar)
and approximately 30% (120 bar), highlighting its pivotal role in post-fracture
behaviour.
 All SMAs display a consistent recovery strain of approximately 12%. Minor
variations in extensometer recovery strain values are attributed to the proximity
of the extensometer location to the failure point.
 An interesting trend emerges where groups with the highest EA values
consistently exhibit the highest start stress (σMs) and the lowest initial strain (Ԑs).
This pattern extends to the lowest modulus of elasticity (EM), resulting in the
lowest ultimate stress (σu) and the highest ultimate strain (Ԑu). Conversely,
groups with lower EA values exhibit the opposite pattern.
 Group 11, with lowering the temperature for a specific part of the specimen than
the rest resulted in TB nucleation in this region, leading to a higher stiffness value
for the crosshead compared to the extensometer. Notably, this condition showed
a lower degree of slip, resulting in ductile fractures.

93
 The impact of temperature variation in Group 11 was further emphasized by a
significant reduction in the EA to 9 GPa, approximately one-third of the value
for the material with uniform temperatures as in group 2. This reduction
contributed to lower transformation stresses and higher transformation strains,
ultimately leading to a higher (EM).
 Group 2 with uniform temperature experiences an overall temperature rise for all
the three thermos-couples through the first region of approximately 24°C,
following this, the temperature rise for the second region was limited, measuring
about 8°C for all. For Group 11 with temperature variation, the temperature rise
was in the first region about 27°C for T1, 9°C for T2, and 15°C for T3, with no
additional rise for the second region for T1 and T2, except 20°C rise for T3
attributed to the presence of failure near it.

This study underscores the intricate nature of SMAs under varying testing conditions.
Factors such as gauge length, gauge length location, strain rates, stress rates, gripping
pressure and temperature variation significantly influence mechanical properties, failure
modes, and post-fracture behaviour. These factors must be taken into account in the
process of characterization of SMA thermo-mechanical material properties. The
reliability of material behaviour is paramount for the safe utilization of SMAs in civil
engineering applications. As a result, there is a compelling need for standardized material
specifications for SMAs under tensile testing, ensuring their secure application in civil
engineering projects.

Considering the uncertainty, what do you think regarding SMA material factor, is it higher or lower than
that for steel rebars.

94
REFERENCES

95
References
1. Shaw, J.A., Tips and tricks for characterizing shape memory alloy wire: part 1—differential
scanning calorimetry and basic phenomena. 2008.
2. Hosseini, F., et al., An experimental investigation of innovative bridge columns with
engineered cementitious composites and Cu–Al–Mn super-elastic alloys. 2015. 24(8): p.
085029.
3. Dolce, M., et al., Implementation and testing of passive control devices based on shape
memory alloys. 2000. 29(7): p. 945-968.
4. McCormick, J., et al., Seismic assessment of concentrically braced steel frames with shape
memory alloy braces. 2007. 133(6): p. 862-870.
5. Choi, E., et al., The behavior of concrete cylinders confined by shape memory alloy wires.
2008. 17(6): p. 065032.
6. Shin, M. and B. Andrawes. Uniaxial compression behavior of actively confined concrete
using shape memory alloys. in Structures Congress 2009: Don't Mess with Structural
Engineers: Expanding Our Role. 2009.
7. Czaderski, C., et al., RC beam with variable stiffness and strength. 2006. 20(9): p. 824-833.
8. Shahverdi, M., et al., Iron-based shape memory alloys for prestressed near-surface
mounted strengthening of reinforced concrete beams. 2016. 112: p. 28-38.
9. Suhail, R., Heat activated Prestressing of SMA wires for seismic retrofitting of RC beam-
column joints. 2018, Queen's University Belfast. Faculty of Engineering and Physical
Sciences ….
10. Suhail, R., et al., Heat activated prestressing of Ni48. 46Ti36. 03Nb15. 42 shape memory
alloy for active confinement of concrete sections. 2016: p. 29-30.
11. Dolce, M. and D.J.I.j.o.m.s. Cardone, Mechanical behaviour of shape memory alloys for
seismic applications 2. Austenite NiTi wires subjected to tension. 2001. 43(11): p. 2657-
2677.
12. Dommer, K. and B.J.J.o.M.i.C.E. Andrawes, Thermomechanical characterization of NiTiNb
shape memory alloy for concrete active confinement applications. 2012. 24(10): p. 1274-
1282.
13. Ozbulut, O.E., et al., Feasibility of self-pre-stressing concrete members using shape
memory alloys. 2015. 26(18): p. 2500-2514.
14. Suhail, R., et al., Mechanical behaviour of NiTiNb shape memory alloy wires–strain
localisation and effect of strain rate. 2020. 144: p. 103346.
15. Churchill, C., Tips and tricks for characterizing shape memory alloy wire: part 2—
fundamental isothermal responses. 2009.
16. Churchill, C.B., J.A. Shaw, and M. Iadicola, Tips and tricks for characterizing shape memory
alloy wire: Part 3-localization and propagation phenomena. 2009.
17. Churchill, C.B., J.A. Shaw, and M. Iadicola, Tips and tricks for characterizing shape memory
alloy wire: Part 4–thermo-mechanical coupling. 2010.
18. Kumar, P. and D. Lagoudas, Introduction to shape memory alloys, in Shape memory alloys:
modeling and engineering applications. 2008, Springer. p. 1-51.
19. Otsuka, K. and C.M. Wayman, Shape memory materials. 1999: Cambridge university press.
20. ASTM, A.J.U.h.w.a.o.c.-b.r.c., E8M-16a: Standard Test Methods for Tension Testing of
Metallic Materials, Test Standard, ASTM, 2016.
21. F-07, A., Standard test method for tension testing of nickel-titanium superelastic
materials. 2007, ASTM International West Conshohocken, PA, USA.
22. Liu, Y., Y. Liu, and J.J.S.m. Humbeeck, Lüders-like deformation associated with martensite
reorientation in NiTi. 1998. 39(8): p. 1047-1055.
23. Zhang, X., et al., Experimental study on rate dependence of macroscopic domain and
stress hysteresis in NiTi shape memory alloy strips. 2010. 52(12): p. 1660-1670.

96
24. Feng, P., Q.-P.J.J.o.t.M. Sun, and P.o. Solids, Experimental investigation on macroscopic
domain formation and evolution in polycrystalline NiTi microtubing under mechanical
force. 2006. 54(8): p. 1568-1603.
25. Shaw, J.A., S.J.J.o.t.M. Kyriakides, and P.o. Solids, Thermomechanical aspects of NiTi.
1995. 43(8): p. 1243-1281.
26. Chang, B.-C., et al., Thermodynamics of shape memory alloy wire: modeling, experiments,
and application. 2006. 18: p. 83-118.
27. Lexcellent, C., J.J.S.m. Rejzner, and structures, Modeling of the strain rate effect, creep
and relaxation of a Ni-Ti shape memory alloy under tension (compression)-torsional
proportional loading in the pseudoelastic range. 2000. 9(5): p. 613.
28. Oudah, F. and R.J.A.S.J. El-Hacha, Innovative Self-Centering Concrete Beam-Column
Connection Reinforced Using Shape Memory Alloy. 2018. 115(3).
29. Buehler, W.J., J.V. Gilfrich, and R.J.J.o.a.p. Wiley, Effect of low-temperature phase changes
on the mechanical properties of alloys near composition TiNi. 1963. 34(5): p. 1475-1477.
30. Janke, L., et al., Applications of shape memory alloys in civil engineering structures—
Overview, limits and new ideas. 2005. 38: p. 578-592.
31. Alam, M., M. Youssef, and M.J.C.J.o.C.E. Nehdi, Utilizing shape memory alloys to enhance
the performance and safety of civil infrastructure: a review. 2007. 34(9): p. 1075-1086.
32. Andrawes, B., M. Shin, and N.J.J.o.B.E. Wierschem, Active confinement of reinforced
concrete bridge columns using shape memory alloys. 2010. 15(1): p. 81-89.
33. Choi, E., et al., Confining jackets for concrete cylinders using NiTiNb and NiTi shape
memory alloy wires. 2010. 2010(T139): p. 014058.
34. Bhattacharya, K., Microstructure of martensite: why it forms and how it gives rise to the
shape-memory effect. Vol. 2. 2003: Oxford University Press.
35. Leo, P.H., T. Shield, and O.P.J.A.M.e.M. Bruno, Transient heat transfer effects on the
pseudoelastic behavior of shape-memory wires. 1993. 41(8): p. 2477-2485.
36. Hallai, J.F. and S.J.I.J.o.P. Kyriakides, Underlying material response for Lüders-like
instabilities. 2013. 47: p. 1-12.
37. Brinson, L.C., et al., Stress-induced transformation behavior of a polycrystalline NiTi shape
memory alloy: micro and macromechanical investigations via in situ optical microscopy.
2004. 52(7): p. 1549-1571.
38. Daly, S., G. Ravichandran, and K.J.A.M. Bhattacharya, Stress-induced martensitic phase
transformation in thin sheets of Nitinol. 2007. 55(10): p. 3593-3600.
39. Keefe, A.C., G.P.J.S.M. Carman, and Structures, Thermo-mechanical characterization of
shape memory alloy torque tube actuators. 2000. 9(5): p. 665.
40. Liu, Y., Y. Li, and K.J.P.M.A. Ramesh, Rate dependence of deformation mechanisms in a
shape memory alloy. 2002. 82(12).
41. Li, H., Z.-q. Liu, and J.-p.J.E.s. Ou, Experimental study of a simple reinforced concrete beam
temporarily strengthened by SMA wires followed by permanent strengthening with CFRP
plates. 2008. 30(3): p. 716-723.
42. Choi, E., et al., Seismic protection of lap-spliced RC columns using SMA wire jackets. 2012.
64(3): p. 239-252.
43. Shin, M., B.J.S.M. Andrawes, and Structures, Emergency repair of severely damaged
reinforced concrete columns using active confinement with shape memory alloys. 2011.
20(6): p. 065018.
44. Suhail, R., et al., Thermo-mechanical characterisation of NiTi-based shape memory alloy
wires for civil engineering applications. 2021. 32(20): p. 2420-2436.
45. Davis, J.R., Tensile testing. 2004: ASM international.
46. Becker, W., D.J.F.a. McGarry, and prevention, Mechanisms and appearances of ductile
and brittle fracture in metals. 2002. 11: p. 587-626.

97
‫اﻷكاديمية العربية للعلوم والتكنولوجيا والنقل البحري‬
‫كليـــــــة الهندســـــــة والتكنولوجيــــــــــــــــــا‬
‫قسم الهندسة التشييد والبناء‬

‫عدم اليقين في التوصيف الحراري والميكانيكي للسبائك ذاكرة الشكل من النيكل‬


‫تيتانيوم لتطبيقات الهندسة المدنية‬
‫إعداد‬

‫أسامة عﻼء محمد احمد مكاوى‬


‫رسالة مقدمة لﻸكاديمية العربية للعلوم والتكنولوجيا والنقل البحري ﻻستكمال متطلبات نيل درجة‬

‫ماجستير العلوم‬

‫ﻓﻲ‬

‫الهندسة التشييد والبناء‬

‫إشــــــــراف‬
‫أ‪.‬د‪ .‬نبيل اﻻشقر‬ ‫ا‪.‬د‪ .‬عﻼء محمود مرسﻲ‬
‫أستاذ الهندسة اﻹنشائية‬ ‫أستاذ الهندسة اﻹنشائية‬
‫قسم هندسة التشييد والبناء‬
‫كلية الهندسة والتكنولوجيا‬
‫اﻷكاديمية العربية للعلوم والتكنولوجيا والنقل البحري‬
‫اﻹسكندرية‬

‫‪2023‬‬

‫‪2015‬‬

‫أ‬
‫إقرار الباحث‬

‫أقر بأن المادة العلمية الواردة في هذه الرسالة قد تم تحديد مصدرها العلمي وأن محتوى الرسالة غير‬
‫مقدم للحصول علﻰ أي درجة علمية أخرى‪ ،‬وأن مضمون هذه الرسالة يعكس أراء الباحث الخاصة‬
‫وهي ليست بالضرورة اﻵراء التي تتبناها الجهـة المانحة‪.‬‬

‫اﻻسم‪ :‬أسامة عﻼء محمد احمد مكاوى‬

‫التوقيع ‪................................................................................‬‬

‫التاريخ ‪................................................................................‬‬

‫ب‬
‫المستخلص‪:‬‬

‫نظرا لخصائصها الرائعة مثل المرونة الفائقة‬


‫اكتسبت سبائك ذاكرة الشكل اهتما ًما متزايدًا في أبحاث الهندسة المدنية ً‬
‫وتأثير ذاكرة الشكل والليونة العالية ومقاومة التآكل‪ .‬ومع ذلك‪ ،‬تظهر سلو ًكا ماديًا معقدًا يتأثر بتغير اى عامل مثل‬
‫التركيب الكيميائي والمعالجة الحرارية والميكانيكية والمزيد‪ .‬حتى عندما يكون لعينتين من المواد تركيبة متطابقة‬
‫ومعالجة حرارية‪ ،‬يمكن أن تتصرفا بشكل مختلف في ظل اختبارات الشد‪ ،‬مما يؤدي إلى اختﻼف الخواص الميكانيكية‬
‫وأشكال اﻻنهيار واﻻستطالة بعد الكسر‪ .‬تستمر هذه الفوارق حتى عندما يتم إجراء اﻻختبار وفقًا للمعايير الدولية‬
‫المنصوص عليها ‪ ،‬مما يسلط الضوء على هذا التحدي الحاسم‪.‬‬

‫وفي هذه الدراسة‪ ،‬أجرينا تحقيقًا شامﻼً للكشف عن اﻷسباب الكامنة وراء هذه التناقضات‪ .‬كان تركيزنا منصبًا على فهم‬
‫العﻼقة المعقدة بين المعايير المنصوص عليها و المعمول بها وتعقيداتها اﻹجرائية وكيفية مساهمتها في هذه التناقضات‪.‬‬
‫لقد استكشفنا العديد من العوامل الرئيسية‪ ،‬بما في ذلك اختﻼف طول المقياس‪ ،‬وموقع طول المقياس‪ ،‬ومعدﻻت التغير‬
‫فى سرعة اﻻنفعال للماكينة‪ ،‬ومعدﻻت التغير فى سرعة اﻹجهاد للماكينة‪ ،‬وتغير درجة الحرارة لجزء من العينة‪،‬‬
‫وضغط اﻻمساك‪ .‬أظهرت نتائج هذه الدراسة مساهمة كل معامل تمت دراستها في هذا اﻻختﻼف‪ ،‬مما سلط الضوء على‬
‫السلوك المعقد‪.‬‬

‫تكشف إحدى أهم النتائج التي توصلنا إليها أن زيادة ضغط اﻹمساك‪ ،‬وهو عامل لم يتم دراسته مسبقًا في بروتوكوﻻت‬
‫اﻻختبار القياسية‪ ،‬يقلل بشكل كبير من اﻻنزﻻق‪ ،‬ويوفر قيم أكثر دقة لمعامل المرونة‪ ،‬وتحديد دقيق لﻼجهادات الحرجة‪،‬‬
‫وتقليل اﻻختﻼفات بين الخواص الميكانيكية‪ .‬وهذا تم الحصول عليه من قراءات مقياس اﻻجهاد واﻻنفعال‪ .‬عﻼوة على‬
‫ذلك‪ ،‬يؤدي ارتفاع ضغط اﻹمساك إلى مناطق تشوه ذات استطالة كبيرة ممتدة مع تكوين رقبة‪ ،‬مما يشير إلى سلوك‬
‫المرونة‪ ،‬وانخفاض كبير في المساحة )‪ .(٪50-40‬على العكس من ذلك‪ ،‬يؤدي انخفاض ضغط اﻹمساك إلى سلوك أقل‬
‫ليونة مع الحد اﻷدنى من تشوه اللدن بعد الكسر وانخفاضات أصغر في المساحة )‪ .(٪30-20‬تؤكد هذه اﻷفكار على‬
‫الحاجة الملحة لمراعاة العديد من المعامﻼت أثناء اختبارات الشد ‪ ،‬لتأمين تقييمات دقيقة لخصائصها الميكانيكية‪.‬‬

‫ج‬

You might also like