5
5
5
by
VINU B. KRISHNAN
B.Tech. University of Kerala, 1996
M.S. Syracuse University, 2001
M.S. University of Central Florida, 2004
Fall Term
2007
ii
ABSTRACT
Shape memory alloys are incorporated as actuator elements due to their inherent ability to
sense a change in temperature and actuate against external loads by undergoing a shape change
trigonal so-called R-phase transformation in NiTiFe shape memory alloys offers a practical
temperature range for actuator operation at low temperatures, as it exhibits a narrow temperature-
hysteresis with a desirable fatigue response. Overall, this work is an investigation of selected
science and engineering aspects of low temperature NiTiFe shape memory alloys.
The scientific study was performed using in situ neutron diffraction measurements at the
newly developed low temperature loading capability on the Spectrometer for Materials Research
at Temperature and Stress (SMARTS) at Los Alamos National Laboratory and encompasses
three aspects of the behavior of Ni46.8Ti50Fe3.2 at 92 K (the lowest steady state temperature
attainable with the capability). First, in order to study deformation mechanisms in the R-phase in
Second, with the objective of examining NiTiFe in one-time, high-stroke, actuator applications
(such as in safety valves), a NiTiFe sample was strained to approximately 5% (the R-phase was
transformed to B19' phase in the process) at 92 K and subsequently heated to full strain recovery
under a load. Third, with the objective of examining NiTiFe in cyclic, low-stroke, actuator
K and subsequently heated to full strain recovery under load. Neutron diffraction spectra were
recorded at selected time and stress intervals during these experiments. The spectra were
iii
subsequently used to obtain quantitative information related to the phase-specific strain, texture
SMARTS provided considerable insight into the mechanisms of phase transformation and
pseudoelasticity phenomena. Three phases (R, B19' and B33 phases) were found to coexist at 92
K in the unloaded condition (nominal holding stress of 8 MPa). For the first time the elastic
modulus of R-phase was reported from neutron diffraction experiments. Furthermore, for the
first time a base-centered orthorhombic (B33) martensitic phase was identified experimentally in
a NiTi-based shape memory alloy. The orthorhombic B33 phase has been theoretically predicted
in NiTi from density function theory (DFT) calculations but hitherto has never been observed
experimentally. The orthorhombic B33 phase was observed while observing shifting of a peak
(identified to be {021}B33) between the {111}R and {100}B19' peaks in the diffraction spectra
collected during loading. Given the existing ambiguity in the published literature as to whether
the trigonal R-phase belongs to the P3 or P 3 space groups, Rietveld analyses were separately
carried out incorporating the symmetries associated with both space groups and the impact of
this choice evaluated. The constrained recovery of the B19' phase to the R-phase recorded
approximately 4% strain recovery between 150 K and 170 K, with half of that recovery
occurring between 160 K and 162 K. Additionally, the aforementioned research methodology
developed for Ni46.8Ti50Fe3.2 shape memory alloys was applied to experiments performed on a
iv
The engineering aspect focused on the development of (i) a NiTiFe based thermal
conduction switch that minimized the heat gradient across the shape memory actuator element,
(ii) a NiTiFe based thermal conduction switch that incorporated the actuator element in the form
of helical springs, and (iii) a NiTi based release mechanism. Patents are being filed for all the
This work was supported by grants from SRI, NASA (NAG3-2751) and NSF (CAREER
DMR-0239512) to UCF. Additionally, this work benefited from the use of the Lujan Center at
the Los Alamos Neutron Science Center, funded by the United States Department of Energy,
v
Dedicated
to
My Parents and Late Uncle
vi
ACKNOWLEDGMENTS
First and foremost I would like to express gratitude to my advisor Dr. Raj Vaidyanathan
whose constant persuasion compelled me to pursue PhD in Materials Science and Engineering. I
appreciate him for involving me in the NSF funded project for the in situ neutron diffraction
studies of NiTiFe shape memory alloys at cryogenic temperatures and for his invaluable advice.
Additionally, I would like to thank Dr. C. Suryanarayana, Dr. Eric L. Petersen, Dr. Helge
Heinrich, Dr. Kevin R. Coffey and Dr. William U. Notardonato for serving on my PhD
dissertation defense committee and for their valuable suggestions. I would like to express sincere
gratitude to Dr. Bjørn Clausen and Dr. Donald Brown of Los Alamos Neutron Science Center
(LANSCE) for their experimental help and numerous valuable suggestions. Also, I would like to
Special thanks to Mr. Richard E. Zotti of CREOL for machining parts required for the
engineering project and his helpful tips. I would like to express sincere gratitude to Ms. Cynthia
Harle, Ms. Karen Glidewell and Ms. Kari Stiles of AMPAC for their help in arranging my visits
to Los Alamos and various conferences. I would like to thank all my colleagues in Dr. Raj’s
group, especially Tim Woodruff, R. Mahadevan Manjeri, Shipeng Qiu and Catherine Bewerse.
Moreover, I would like to thank all my friends and well-wishers, for their constant support.
Above all I would also like to thank my parents for their constant support, love, and
encouragement. Furthermore, I want to pay respect to my maternal uncle Mr. M.S. Jayaraj who
had influenced me in many of my career choices, in his tragic demise in August 2005.
vii
TABLE OF CONTENTS
viii
3.3 Results and Discussion .......................................................................................23
3.3.1 Constrained Recovery Experiments....................................................................27
3.3.2 Elastic Modulus of the R-phase ..........................................................................28
3.3.3 Observation of Orthorhombic B33 Phase ...........................................................29
3.4 Conclusions.........................................................................................................31
ix
5.5 Conclusions.........................................................................................................68
x
8.3 Conclusions.......................................................................................................105
REFERENCES ...........................................................................................................................115
xi
LIST OF TABLES
xii
LIST OF FIGURES
Figure 2.1: Transformation temperatures associated with the shape memory effect [17]...............7
Figure 2.2: Types of SMA actuators (a) one-way actuator, (b) biased two-way actuator, and (c)
differential two-way actuator [17]. ..............................................................................10
Figure 2.3: Isotropic scattering of neutron beam by a nucleus. .....................................................14
Figure 3.1: Schematic of the SMARTS diffractometer showing the incident and diffracted
neutron beam during the application of compressive stress on the sample. ................20
Figure 3.2: Schematic of the SMARTS diffractometer with the cryogenic setup (top view) [8]..21
Figure 3.3: Comparison of the diffraction patterns obtained at room temperature and at 92 K, in
the no-load condition (nominal holding stress of 8 MPa)............................................24
Figure 3.4: Splitting of {112}R and {300}R peaks of diffraction patterns observed at 216 K and at
92 K. .............................................................................................................................24
Figure 3.5: Macroscopic stress-strain curve obtained ex situ at 92 K. ..........................................25
Figure 3.6: Stress-induced transformation from R-phase to B19' phase above a stress of 68 MPa,
during loading at 92 K. ................................................................................................26
Figure 3.7: Limited twinning in the R-phase and strain redistribution above 68 MPa due to the
formation of stress-induced B19' phase during loading at 92 K..................................26
Figure 3.8: Transformation hysteresis between B19' phase and R-phase, typical of shape memory
alloys. The dotted lines qualitatively represent the region where the experiments were
conducted. ....................................................................................................................27
Figure 3.9: Constrained recovery of the R-phase from the B19' phase during the heating of the
5% strained sample from 150 K to 170 K....................................................................28
Figure 3.10: Tilting of {021}B33 during loading at 92 K, observed as a peak movement between
{111}R and {100}B19'. ..................................................................................................30
Figure 3.11: Crystallographic correspondence between the cubic B2 and the orthorhombic B33
structures. .....................................................................................................................30
Figure 3.12: Left – the monoclinic B19' unit cell with γ = 97°. Right – the monoclinic B19' unit
cell with γ = 107° and the orthorhombic B33 unit cell. The green rectangle shows the
B33 {021} plane. (Figures modified from Huang et. al., Nature Materials, 2003 [57]).30
Figure 4.1: Unit cell of R-phase using P3 space group (left) and P 3 space group (right). ..........40
Figure 4.2: A typical GSAS Rietveld refinement output with the R-phase refined as belonging to
the P3 space group (for spectrum obtained from the sample under a compressive
stress of 44 MPa at 237 K) for diffracting lattice planes whose normals are parallel to
the loading axis. The measured data are indicated by cross-marks and the calculated
xiii
profile is indicated by the solid-line curve. The line-marks below the profile pattern
indicate the positions of all possible Bragg reflections. The lower graph shows the
difference between the measured and calculated profile patterns. ..............................40
Figure 4.3: A typical GSAS Rietveld refinement output with the R-phase refined as belonging to
the P 3 space group (for spectrum obtained from the sample under a compressive
stress of 44 MPa at 237 K) for diffracting lattice planes whose normals are parallel to
the loading axis. The measured data are indicated by cross-marks and the calculated
profile is indicated by the solid-line curve. The line-marks below the profile pattern
indicate the positions of all possible Bragg reflections. The lower graph shows the
difference between the measured and calculated profile patterns. ..............................41
Figure 4.4: Section of normalized neutron diffraction spectra corresponding to R-phase {111}
lattice plane reflections (most intense) for the various experimental conditions. The
reflections shown here are from diffracting lattice planes whose normals are parallel
to the loading axis. .......................................................................................................45
Figure 4.5: Section of normalized neutron diffraction spectra corresponding to R-phase {411}
and { 511 } lattice plane reflections (second most intense) for the various experimental
conditions. The reflections shown here are from diffracting lattice planes whose
normals are parallel to the loading axis. ......................................................................46
Figure 4.6: R-phase ( 111 ) axial distribution plots (refined as belonging to the P3 space group)
for the various experimental conditions. φ is the angle between the ( 111 ) plane
normal and the loading axis. ........................................................................................47
Figure 4.7: R-phase ( 111 ) axial distribution plots (refined as belonging to the P 3 space group),
for the various experimental conditions. φ is the angle between the ( 111 ) plane
normal and the loading axis. ........................................................................................48
Figure 5.1: Macroscopic stress-strain curve at 92 K. The symbols indicate the stresses at which
neutron diffraction spectra were obtained and analyzed..............................................54
Figure 5.2: Section of normalized neutron diffraction spectra observed in both banks during
loading and unloading at 92 K. (a) The diffracting lattice planes are parallel to the
loading axis (bank 1). (b) The diffracting lattice planes are perpendicular to the
loading axis (bank 2)....................................................................................................55
Figure 5.3: A typical GSAS Rietveld refinement output with the R, B19' and B33 phases refined
for diffracting lattice planes whose normals are parallel to the loading axis. The
measured data are indicated by cross-marks and the calculated profile is indicated by
the solid-line curve. The line-marks below the profile pattern indicate the positions of
all possible Bragg reflections. The lower graph shows the difference between the
measured and calculated profile patterns.....................................................................56
Figure 5.4: Volume fraction of various phases obtained by Rietveld refinement as a function of
applied compressive load, (a) during loading and (b) during unloading. The typical
error associated with the volume fraction determination is ±3%. ...............................57
xiv
Figure 5.5: Texture evolution in the R and B19' phases represented by the texture index during
(a) loading and (b) unloading.......................................................................................57
Figure 5.6: R-phase ( 111 ) axial distribution plots during (a) loading and (b) unloading. φ is the
angle between the ( 111 ) plane normal and the loading axis.......................................58
Figure 5.7: B19' (100) axial distribution plots during (a) loading and (b) unloading. φ is the
angle between the (100) plane normal and the loading axis.......................................59
Figure 5.8: Lattice strains obtained from single peak fitting and Rietveld analysis for (a) R-phase
{111} and (b) B19' {100} planes, during loading/unloading. The arrows indicate the
loading and unloading direction. The typical errors associated with determination of
d-spacing and lattice parameters (Rietveld) are ±0.0003Å and ±0.0002Å
respectively. .................................................................................................................61
Figure 5.9: Anisotropic strains in B33 {021} planes with respect to {010} and {001} planes,
during distortion of the B33 phase. The typical errors associated with determination
of strains are ±0.0002Å................................................................................................62
Figure 6.1: Constrained recovery of the R-phase from B19' phase during the heating of the 5%
strained NiTiFe sample from 150 K to 170 K. ............................................................72
Figure 6.2: Displacement in the SMARTS load frame as a function of temperature during heating
of the 5% strained NiTiFe sample from 150 K to 170 K.............................................73
Figure 6.3: Phase fraction evolution of the B19' and R phases during constrained recovery of the
5% strained NiTiFe sample from 150 K to 170 K. ......................................................73
Figure 6.4: Texture evolution in the B19' and R phases represented by texture index during
constrained recovery of the 5% strained NiTiFe sample from 150 K to 170 K. .........74
Figure 6.5: (a) B19' (100) and (b) R-phase ( 111 ) axial distribution plots planes during
constrained recovery of the 5% strained NiTiFe sample from 150 K to 170 K. φ is
the angle between the corresponding plane normal and the loading axis...................75
Figure 6.6: Constrained recovery of the B2 phase from the R-phase during heating of the 1%
strained NiTiFe sample from 231 K to 243 K, (a) showing R{411} and R{303}
combining to form B2{210} and (b) showing R{300} and R{112} combining to form
B2{110}. ......................................................................................................................76
Figure 6.7: Phase fraction evolution of the R and B2 phases during constrained recovery of the
1% strained NiTiFe sample from 231 K to 243 K. ......................................................77
Figure 6.8: Texture evolution in the R and B2 phases represented by texture index during
constrained recovery of the 1% strained NiTiFe sample from 231 K to 243 K. .........78
Figure 6.9: (a) R-phase ( 111 ) planes and (b) B2 (100) axial distribution plots during constrained
recovery of the 1% strained NiTiFe sample from 231 K to 243 K. φ is the angle
between the corresponding plane normal and the loading axis. .................................78
Figure 7.1: High temperature furnace on the SMARTS load frame [6]. .......................................84
xv
Figure 7.2: A typical GSAS Rietveld refinement output with B19' structure refined for
diffracting lattice planes whose normals are parallel to the loading axis. The measured
data are indicated by cross-marks and the calculated profile is indicated by the solid-
line curve. The line-marks below the profile pattern indicate the positions of all
possible Bragg reflections. The lower graph shows the difference between the
measured and calculated profile patterns.....................................................................85
Figure 7.3: A typical GSAS Rietveld refinement output with B19 phase refined for diffracting
lattice planes whose normals are parallel to the loading axis. The measured data are
indicated by cross-marks and the calculated profile is indicated by the solid-line
curve. The line-marks below the profile pattern indicate the positions of all possible
Bragg reflections. The lower graph shows the difference between the measured and
calculated profile patterns. ...........................................................................................86
Figure 7.4: Room temperature measurements at the applied load in each cycle, before heating.
These spectra are from the bank 2 detector, where the diffracting lattice planes are
perpendicular to the loading axis. ................................................................................87
Figure 7.5: High temperature measurements for each cycle. These are from the bank 2 detector
where the diffracting lattice planes are perpendicular to the loading axis...................88
Figure 7.6: (a) B19 (100) and (b) B19 (011) axial distribution plots at the maximum temperatures
correspond to cycles 5, 6 and 7. φ is the angle between the corresponding plane
normal and the loading axis. ........................................................................................89
Figure 7.7: Section of normalized neutron diffraction spectra showing strains between the B2
{100} planes for cycles 1 and 7, at the maximum temperature. These spectra are from
the bank 2 detector, where the diffracting lattice planes are perpendicular to the
loading axis. .................................................................................................................91
Figure 7.8: Macroscopic stress-displacement response corresponding to the stress-induced
martensite experiment. The symbols represent the stresses at which neutron
diffraction spectra were recorded.................................................................................92
Figure 7.9: Development of stress-induced martensite during cycling at 498 K. These spectra are
from the bank 2 detector, where the diffracting lattice planes are perpendicular to the
loading axis. .................................................................................................................93
Figure 8.1: Differential scanning calorimeter (DSC) response of the NiTiFe wire used. ...........101
Figure 8.2: Switch performance between 200 K and room temperature. ....................................103
Figure 8.3: The NiTiFe helical spring switch in, (a) the closed (extended) position at a room
temperature of 298 K and (b) the open (contracted) position at 233 K. ....................103
Figure 8.4: The shape memory alloy release mechanism in, (a) before actuation and (b) after
actuation, releasing a 100 lb load...............................................................................105
Figure 9.1: Behavior of Ni46.8Ti50Fe3.2 shape memory alloy between room temperature and 92 K.112
xvi
LIST OF ACRONYMS
xvii
CHAPTER ONE: INTRODUCTION
1.1 Motivation
(pseudoelasticity) that is of the first order and thermoelastic. The strain recovery associated with
the shape recovery is instantaneous (limited to the speed of sound in the material) besides
generating large forces (as high as 500 MPa), resulting in their application as actuators. SMA
actuators have remarkable potential for use in space applications due to the following advantages
[1]: (i) high power to weight and stroke length to weight ratios; (ii) combination of sensor and
actuator elements in a single component; (iii) clean, debris-less, spark-free, silent operation and
(iv) capability of operating in zero gravity environments with small, controlled accelerations. In
NiTi SMAs, the phase transformation usually takes place between a monoclinic, so-called
martensite phase and a cubic, so-called austenite phase. The addition of Fe to the NiTi system
trigonal R-phase.
The prime motivation for this research comes from NASA’s requirement for a cryogenic
thermal conduction switch for thermal management at cryogenic temperatures, and cryogenic
safety mechanisms such as self-healing gaskets, valves and seals. NiTiFe SMA based actuators
1
There is a need for better understanding the microscopic mechanisms concerning the
phenomena for improving the design, the scope of application ranging from cryogenic actuators
techniques for the investigation of microscopic deformation mechanisms coupled with phase
transformations and twinning, which occur in the actuator from temperature changes and
diffraction during loading at varying temperatures is uniquely suited for investigating the texture,
strain and phase fraction evolution in bulk polycrystalline SMA samples [2-5]. Neutron
diffraction measurements have been performed in situ during loading at ambient and high
temperatures, but not at low temperatures. Again, the underlying deformation mechanisms using
neutron diffraction have been studied for shape memory alloys at around room temperature.
lacking. Accordingly, a low temperature loading capability for in situ neutron diffraction
measurements has been implemented on the Spectrometer for Materials Research at Temperature
and Stress (SMARTS) [6] at Los Alamos National Laboratory, with capabilities as low as 90 K
[7].
With the new cryogenic capability, in situ neutron diffraction measurements were
experiments focused on three aspects on the behavior of Ni46.8Ti50Fe3.2 in the context of real
world engineering applications. First, in order to study deformation mechanisms and stress-
constant temperature of 92 K under external loading. Second, with the objective of examining
2
one-time, high-stroke, actuator applications (such as in safety valves), a Ni46.8Ti50Fe3.2 sample
was strained to approximately 5% at 92 K and subsequently heated to full strain recovery under a
constant load. Third, with the objective of examining cyclic, low-stroke, actuator applications
subsequently heated to full strain recovery under a constant load. Neutron diffraction spectra
were recorded at suitable stress or temperature intervals during these experiments that assisted in
the monitoring of the phase-specific strain, texture and volume fraction evolutions.
In theory, the complete strain tensor can be determined by measuring strain in at least six
orientations. Additionally, six banks of detectors would help in determining the preferred
temperature and external loads, limits the SMARTS diffractometer at Los Alamos National
Laboratory where the present study was conducted, to only two banks of detectors [6]. There is a
need to establish a methodology under these constraints, to quantitatively assess the data
collected from NiTiFe systems based on time-of-flight neutron diffraction spectra acquired using
SMARTS, under the influence of external loading and changes in temperature [8]. A variety of
refinement techniques are available to analyze the diffraction data in general and the present
study uses one such method, Rietveld analysis. Of the programs available for Rietveld analysis,
GSAS (General Structure Analysis System) [9] is chosen for the present study in obtaining the
texture, phase fraction, strain and crystal structure parameters. Additionally, the research
methodology for Ni46.8Ti50Fe3.2 shape memory alloys has been applied to experiments performed
3
In addition to the scientific investigation using neutron diffraction measurements and
subsequent analyses, efforts were focused on the engineering development of selected cryogenic
NiTiFe based actuator devices. The engineering aspects were focused on the development of a (i)
NiTiFe based thermal conduction switch that minimized the thermal gradient in the actuator
element, (ii) NiTiFe based thermal conduction switch that incorporated the actuator element in
the form of helical springs, and (iii) a NiTiFe based low-temperature release mechanism.
1.2 Organization
shape memory alloys and briefly introduces the neutron diffraction technique and neutron spectra
analysis. Chapter 3 discusses the cryogenic neutron diffraction setup on SMARTS, overview of
experiments and the qualitative results. Chapter 4 addresses the ambiguity in the structure of the
R-phase through diffraction spectra analyses [8]. Chapter 5 discusses the deformation behavior
performed on NiTiFe compression samples, i.e., from the B19' phase to the R-phase and the R-
phase to the B2 phase. Chapter 7 is an extension of the research methodology presented in the
previous chapters to the experiments performed on Ni29.5Ti50.5Pd20 shape memory alloys using in
situ neutron diffraction. These include load-bias experiments and stress-induced martensite
experiments. Chapter 8 presents engineering aspects and development of NiTi based actuators
4
CHAPTER TWO: LITERATURE REVIEW
This chapter is a selected review of the literature on NiTi/NiTiFe shape memory alloys
and briefly introduces neutron diffraction and neutron spectra analysis. Due to the stand-alone
nature of each chapter in this dissertation, a more detailed review is available in the individual
“Shape memory alloys” (SMAs) can recover from certain changes in shape or
“weaker phase” from their high temperature “stronger phase”. The strain recovery is
instantaneous besides generating large forces (as high as 500 MPa), resulting in their application
as actuators [12-15]. Even though a wide range of alloys are found to exhibit the shape memory
effect or pseudoelasticity, NiTi alloys are of practical interest owing to a superior combination of
material properties coupled with substantial strain recovery (up to 8%). The properties of SMAs
5
2.1.1 Shape Memory Effect
transformation that is responsible for the shape memory effect. A martensitic transformation is
cooperative movement of atoms [12]. Although the transformation involves distortion of the
parent unit cell, there is no diffusion of atoms associated with it. Due to this cooperative
movement (displacive), there exists a one-to-one lattice correspondence between the parent
austenite phase and the resulting martensite phase. In thermoelastic martensitic transformations,
the martensite crystals, once nucleated, will grow or shrink corresponding to a decrease or
with the nucleation and growth of martensite within the parent phase as the temperature
decreases, and an increase in the temperature results in the nucleation and growth of the parent
When an SMA in its high-temperature, austenite phase is cooled (Figure 1.1), the
(martensitic finish temperature). Upon heating (Figure 1.1), the reverse transformation from
6
Ms Af Md
0 100
percentage martensite
percentage austenite
cooling
martensite austenite
heating
100 0
Mf As
temperature
Figure 2.1: Transformation temperatures associated with the shape memory effect [17].
A certain number of symmetry-related variants are possible in the martensitic phase due
morphologies such as lath, plate, wedge or square shapes [18-20]. These morphologies have at
least two variants (including habit plane variants) oriented in a certain fashion. The type of
morphology depends on the crystal system as well as the stress state (both internal and external).
For example, Bhattacharya [20] studied a wedge-like morphology and found that this
helps produce zero macroscopic strain by orienting the variants of resultant martensite in various
morphologies. When external stress is applied on SMAs in the martensitic state, an internal
7
rearrangement (consequence of energy minimization) takes place in such a way that the strain
can be accommodated. This internal rearrangement take place in two ways: (i) the realignment of
differently oriented morphologies in the favor of external stress (martensite reorientation in some
cases [19]), and (ii) the realignment of different variants within a single morphology such that
twinning or variant coalescence (detwinning in some cases [19]) takes place. Strictly speaking,
the realignment of the morphology is not possible unless the crystal structure of the matrix
surrounding the morphology, in addition to that inside the morphology, twins. In other words,
realignment takes care of the shape change without introducing dislocations (that result in
crystal system can form. For example, monoclinic martensite has 12 variants in contrast to 4
variants (excluding habit plane variants) in the trigonal R-phase, resulting in about 8% strain
recovery for monoclinic martensite compared to about 1% in the R-phase in NiTi based alloys.
During the reverse transformation, each variant of martensite transforms back to a single variant
2.1.2 Superelasticity
elasticity. In the presence of applied stress, the parent high symmetry structure undergoes a
stress-induced transformation rather than slip. This happens at relatively lower temperatures
(above the Ms temperature) where the stress-induced transformation is favored rather than slip,
8
caused by the movement of dislocations. The transformed region when compared to the matrix
(parent) has a different crystal structure that is of lower symmetry. Thus the presence of external
stress-induced martensite no longer forms (Figure 1.1). In the case of pseudoelasticity, the
martensite formed from transformation of the parent structure is stable only in the presence of
applied stress, in the absence of it, martensite reverts to the parent phase. However, at
Shape memory alloys (SMAs) have potential for use in actuator elements due to the
following advantages [1,17]: (i) They combine sensory and actuation functions. The SMA
element inherently senses a change in temperature and actuates by undergoing a shape change as
a result of a phase transformation. Consequently, the need for external electronic sensors and
control is eliminated. (ii) They function in a clean, debris-less, spark-free manner. The shape
change that is responsible for the actuator displacement is again an inherent material property. It
is not associated with moving parts that require lubrication or electrical signals with a potential
to spark. (iii) They have high power/weight and stroke length/weight ratios. The operating range
includes strain and stress limits of 8% and 700 MPa, respectively, depending on the number of
required cycles. (iv) They possess the ability to function in zero-gravity environments with
small, controlled accelerations. The displacement strains are a result of a thermally induced
phase transformation which can be controlled by the heat transfer rate. As a result, end-of-
deployment shock loadings (associated with spring deployed structures) can be avoided in SMA-
9
based deployable structures, thus eliminating the overall complexity by avoiding the use of
dampers. (v) They are reliable and offer design flexibility. SMA actuators can function in a
linear or rotary manner (or a combination of the two). Typically, an SMA actuator utilizes the
shape memory effect to generate force and motion. When compared to electronically controlled
processor (e.g., IC circuit) and electric actuator (e.g., motor), SMA actuators have fewer parts,
higher reliability and lower fabrication costs. SMA actuators can be of one-way or two-way type
(Figure 1.2).
force
SMA
elements
(c)
Figure 2.2: Types of SMA actuators (a) one-way actuator, (b) biased two-way actuator, and (c)
differential two-way actuator [17].
One-way SMA actuators (Figure 1.2a) are used where one-time actuation is necessary
such as in safety devices, couplings, release mechanisms etc. In a one-way actuator, the SMA
element acts against some force when the temperature reaches a pre-set temperature. Two-way
10
actuation can occur either in a biasing mode (Figure 1.2b) or a differential mode (Figure 1.2c).
The biasing mode uses a bias spring opposing the SMA spring for actuation in either direction.
When the temperature of the SMA element rises to a pre-set temperature, the SMA element
being stronger as it undergoes a phase transformation, will act against the bias spring force
actuating in one direction. As the temperature of the SMA element drops below a certain
temperature, the bias spring force overcomes the SMA element force, thus acting in the opposite
direction. Biasing mode operation is more commonly employed and offers enhanced flexibility
in design. In the differential mode of operation, the bias spring is substituted by another SMA
spring. Here actuation in either direction can be achieved by appropriately heating or cooling
either of the two SMA elements. Applications requiring precise movement and high accuracy
NiTi based SMAs are a preferred choice for designing actuators due to the following
advantages: (i) able to accommodate large strains (up to 8%), (ii) high corrosion resistance [12,
16,22], (iii) capable of operating in a relatively wide range of temperatures by slightly varying
compared to other SMAs [28] and (v) high fatigue life of ~106 cycles when strains are restrained
to ~ 1% [29].
room temperature, between a monoclinic martensite (B19' structure) and a cubic austenite (B2
structure). NiTi alloys exhibit behavior on the basis of composition (variation in Ni/Ti ratio or
11
addition of a third element replacing either Ni or Ti) and the history of thermo-mechanical
treatments. The behavior can include one or more combinations of a shift in transformation
hysteresis. Accordingly, the shape memory effect and the superelastic response in NiTi alloys are
Typically, there is hysteresis between the forward and reverse transformations in shape
memory alloys. The transformation hysteresis is a result of elastic strain energy dissipation, the
energy associated with frictional resistance to interface motion and similar dissipative processes
[30-32]. Generally, thermal hysteresis ranges from 1.5-2 K for R-phase transformations, 10-15 K
for certain NiTiCu alloys, 20-60 K for binary NiTi alloys and in excess of 100 K for NiTiNb
alloys [22, 30]. Aspects such as alloy composition, thermal cycling, thermo-mechanical
temperature and further introduces an intermediate trigonal R-phase. The cubic austenite to
trigonal R-phase transformation in NiTiFe shape memory alloys offer a practical temperature
12
2.3.3 R-Phase in NiTi Alloys
cubic austenite phase [34-36]. The R-phase transformation exhibits a temperature hysteresis as
low as 1.5 K, useful in designing certain types of actuators that operate in narrow temperature
ranges. The maximum recoverable strain (approximately 1%) during the R-phase transformation
is fairly low compared to the monoclinic B19' phase transformation [37]. However, the fatigue
life of the R-phase is very good compared to the B19' phase [38].
The formation of the R-phase and the subsequent formation of the monoclinic B19' phase
has been attributed to the presence of defects in the form of dislocations, precipitates, and/or
impurity elements. The R-phase in NiTi based SMAs can be achieved by suppressing the
martensitic transformation [23,39], i.e., by: (i) increasing the Ni content, (ii) ageing at lower
temperatures (573-773 K) after solution treatment [40-42] to form Ni4Ti3 precipitates, (iii)
annealing below the recrystallization temperature immediately after cold-working [40,41], (iv)
adding a third element such as Fe, Co or Al [41,43] and (v) thermal cycling [44-46].
13
2.4 Neutron Diffraction
k
r scattered circular
k 2θ wave
x
incident plane
wave
Whereas both neutrons and x-rays (laboratory sources) can be used for materials
characterization, neutrons can provide information on the bulk behavior in materials. The most
obvious difference between neutrons and x-rays is that neutrons interact with atoms via short-
ranged nuclear interactions, as opposed to electron interactions in the case of x-rays. Since the
nucleus is extremely small compared to the electron cloud in an atom, the neutrons can penetrate
a material very deeply before being scattered (Figure 2.1). Crystal structure, residual stress,
phase fraction in a multiphase system and texture are few examples of the information that can
be obtained using this technique [47]. Furthermore, in situ neutron diffraction allows researchers
to study in real time the mode of deformation mechanism in material systems, such as elastic
the study of SMA alloys which twin at low-temperatures to accommodate strains and undergo
14
2.4.1 Rietveld Refinement
The Rietveld refinement procedure was originally introduced for the analysis of constant
wavelength neutron diffraction data [48, 49] and was subsequently used for the refinement of x-
ray diffraction patterns in the late 1970s. In Rietveld analysis, the method of least squares is used
to obtain the best possible fit between the entire observed pattern with the corresponding
calculated pattern. This has many advantages; significant among them are the analysis of elastic
strains owing to the change in lattice plane positions as a result of external loading, an account of
the phase fraction evolution or twinning corresponding to changes in peak intensities and the
Based on the crystal structure of the material being refined, the intensity, Yci , can be
calculated at every point in the spectrum and the position of the observed lattice reflections can
be predicted. The calculated intensity depends on numerous factors and can be determined from
Yci = s ∑
K
LK |FK| 2 φ (2θi - 2θK) PK A+ Ybi (2.1)
where, s is the scaling factor, K is the h, k, and l Miller indices for a Bragg reflection, LK
includes the Lorentz, polarization and multiplicity factors, FK is the structure factor, φ is the
reflection profile function, 2θi is the observed position of the Bragg peak, 2θK is the calculated
position of the Bragg peak corrected for the zero-point error in the detector, PK is the preferred
orientation function, A is an absorption factor and Ybi is the background intensity at the ith step of
the iteration process. Thus in order to obtain a good fit between the experimental and calculated
patterns, a number of atom position, thermal, site occupancy, lattice, background, instrumental
15
and profile parameters are typically refined, among others. The profile function may include
effects from microstrain and crystallite size related peak broadening, specimen and instrument
source and geometry. The presence of preferred orientation necessitates the refinement for
texture. For example in GSAS [9], two different models are used for the refinement of texture.
The first, following the formulation of March and Dollase [51, 52], uses a cylindrical
symmetrical version of an ellipsoidal model to describe texture. The second uses a generalized
studies have shown that the March-Dollase texture formulation is inadequate to account for
this work. The sequence and/or number of parameters being simultaneously refined can
influence whether the least square minimization algorithm settles in a local minimum as opposed
to a global minimum. Thus in order to avoid arriving at such false minima [50], there is a need
to both quantitatively and qualitatively evaluate the accuracy of the refinement. A qualitative
check may include visual inspections as well as ensuring that the model considered has all the
necessary parameters to describe both structural and diffraction effects. A quantitative check
crystallographic residual factors (Rwp, Rp, etc.) and goodness of fit (χ2) that can help judge the
16
2.4.2 Single Peak Fitting
The single peaks corresponding to specific hkl reflections in the diffraction pattern can be
fitted individually for lattice spacing d hkl , to follow strains specific to different grain orientations
with respect to the unloaded state. The strain for a specific hkl plane is then given by:
d hkl − d 0hkl
ε hkl = (2.2)
d 0hkl
where d 0hkl is the corresponding “strain free” spacing. However, a small nominal compressive
stress of 8 MPa (this was needed to keep the compression sample aligned at the SMARTS load
frame) was used as the “strain free” unloaded state, ignoring the pre-existing intergranular
stresses. The RAWPLOT program in GSAS [9] was used to fit the peaks individually.
reflections. The ratio of intensities of peaks with respect to the unloaded state can follow the
texture evolution. The ratio of peak intensity (normalized) for a specific hkl plane is then given
by:
I hkl I 0Spectra
I hkl = × (2.3)
I 0hkl I Spectra
where I hkl is the intensity of a specific hkl plane at a given stress state, I 0hkl is the
corresponding unloaded state, I Spectra is the integrated peak intensity of the entire spectra at the
corresponding stress state and I 0Spectra is the integrated peak intensity of the entire spectra
corresponding to unloaded state. The normalization procedure accounts for the different count
17
CHAPTER THREE: CRYOGENIC NEUTRON DIFFRACTION SETUP
AND MEASUREMENTS AT 92 K
This chapter presents an overview of the cryogenic neutron diffraction setup [7]
implemented on the Spectrometer for Materials Research at Temperature and Stress (SMARTS),
constrained recovery of the sample was studied that simulated the functioning of an SMA as an
actuator under a bias force, by controlled heating of the sample at a constant load of 50 MPa
intervals during these experiments, facilitating the monitoring of the strain, texture and phase
fraction evolution and thus the microscopic deformation mechanisms. The results presented here
identify and outline the fundamental mechanisms and phenomenology, with details presented in
subsequent chapters.
3.1 Introduction
In polycrystalline shape memory alloys, the phase transformation driven by chemical free-energy
change competes with elastic strain energy, frictional resistance to interface motion and similar
dissipative processes, which manifests as thermal hysteresis [31,32]. The unique lattice
18
twinning or variant coalescence (detwinning) are responsible for the phenomena of shape
memory and pseudoelasticity in shape memory alloys [12]. NiTiFe shape memory alloys are
reported to undergo two-stage martensitic transformation, initially from a cubic austenite phase
technique at cryogenic temperatures under the application of stress in NiTiFe alloys, the phase
in situ neutron diffraction measurements, a novel 90 K loading capability [7] was implemented
on the Spectrometer for Materials Research at Temperature and Stress (SMARTS) [6] at Los
Alamos Neutron Science Center. Experiments were performed at 92 K (lowest possible steady
made during these experiments. The quantitative analyses of the experiments are presented in
Chapters 4 through 6.
neutron diffractometer at Los Alamos National Laboratory [6]. SMARTS (Figure 3.1) has two
banks of detectors, bank 1 oriented for lattice planes whose normals are perpendicular to the axis
of loading and bank 2 oriented for lattice planes whose normals are parallel to the axis of
loading. The high temperature vacuum furnace and load frame facilitate research on materials
19
under high loads (250 kN) and high temperatures (1773 K). Recently, a cryogenic loading
capability (Figure 3.2) was incorporated on the SMARTS diffractometer by the UCF research
group [7], principally developed for the study of NiTiFe shape memory alloys. With the new
cryogenic setup, samples can be cooled to temperatures as low as 90 K during loading and
INCIDENT
NEUTRONS
DIFFRACTED NEUTRONS
TO BANK 1
DIFFRACTED NEUTRONS
TO BANK 2
COMPRESSIVE
STRESS
Figure 3.1: Schematic of the SMARTS diffractometer showing the incident and diffracted
neutron beam during the application of compressive stress on the sample.
The cryogenic capability on SMARTS (Figure 3.2) relies on conductively cooling the
sample while in contact with the load-applying compression platens. The capability primarily
consists of an aluminum vacuum chamber that minimizes convective heat losses, and prevents
freezing of moisture and other gases in the atmosphere at low temperatures. It has two identical
flanged cylindrical ports that were exactly opposite to each other that accommodated the
insertion of two compression platens, connected to the push rods of the mechanical loading
20
system on SMARTS. The path for neutrons is provided via four rectangular windows at 90º to
each other and positioned at 45º with respect to the flanged cylindrical ports.
INCIDENT
COLLLIMATOR TRANSLATOR
VACUUM
CHAMBER
LOAD CELL
SAMPLE
ACTUATOR
ABSORPTION
TRAP
BANK 2 OF BANK 1 OF
DETECTORS DETECTORS
Figure 3.2: Schematic of the SMARTS diffractometer with the cryogenic setup (top view) [8].
The cooling system consisted of conductively cooling the test sample via the load-
applying platens that in turn were cooled by circulating liquid nitrogen through the internal
cooling channels. A thermally insulating coupling system isolated the cold compression platens
from the warm push rods (due to heat generated from the hydraulic actuators) of the SMARTS
load frame. The cooling system consisted of a temperature control system that maintained a
constant temperature under variable loading and also varied the temperature at constant loading.
The temperature control system consisted of a custom program developed using LabVIEW
software, a set of thin surface heaters mounted on the conical surface of each platen and
thermocouples mounted at various locations on the test sample, platens, inlet and vent locations
of the liquid nitrogen lines. The controller collected temperature data from the thermocouples
21
and maintained the temperature though surface heaters. The temperature control system allowed
the temperature of the test sample to be stabilized (± 1 K) at any temperature specified by the
operator, at or above 92 K. This further required regulating the liquid nitrogen flow valves
manually.
and vacuum-arc-remelting, was procured from Special Metals, NY. Cylindrical samples of 10
mm diameter by 24 mm length were cut from the billet by electrical discharge machining (EDM)
for the neutron diffraction measurements. The samples were then solutionized at 1023 K for 1
hour in a vacuum furnace and immediately oil quenched to room temperature. Differential
scanning calorimetry (DSC) at a rate of 0.33 Ks-1, under nitrogen cover gas was used to
determine the start and finish of the R-phase transformation from austenite and the
corresponding reverse transformation to austenite from the R-phase. These temperatures were
determined to be 236, 223, 227 and 239 ± 2 K, respectively. The transformation to monoclinic
(B19') phase was below 120 K and was hence outside the operating range of the calorimeter
used.
22
3.2.4 Experimental Temperature and Strain Parameters
ii. With the objective of examining NiTiFe in cyclic, low-stroke, actuator applications (such
iii. With the objective of examining NiTiFe in one-time, high-stroke, actuator applications
Neutron diffraction spectra were recorded at suitable stress and time intervals during
these experiments, which assisted in the monitoring of the phase-specific strain, texture and
While cooling the sample to 92 K, the cubic B2 (austenite) transformed to the trigonal R-
phase. Figure 3.3 shows the comparison of diffraction spectra for the B2 phase (cubic) at room
temperature and the R-phase (trigonal) at 92 K, in the no-load condition (nominal holding stress
of 8 MPa). The R-phase spectra are normalized with respect to the austenite spectra. Certain
regions of the diffraction spectra are magnified for clarity. During the austenite to R-phase
transformation, the {110}B2 peak splits to {112}R and {300}R, respectively (Figure 3.3). This is
23
due to the unit cell elongation in the <111> crystallographic direction of the B2 phase and
interpreted as the manifestation of the R-phase [54]. Similarly, the {210}B2 peak splits to {303}R
110
B2 R phase at 92K R phase at 92K
at holding load
at holding load
2 2.05 2.1 2.15 2.2 2.25 2.3
d-spacing (Å)
B2 phase at RT
normalized intensity (arbitrary units)
B2 phase at RT
411 at holding load
R
R phase at 92K
at holding load 303
R
at holding load 100
B2
B2 phase at RT
at holding load
Figure 3.3: Comparison of the diffraction patterns obtained at room temperature and at 92 K, in
the no-load condition (nominal holding stress of 8 MPa).
normalized intensity (arbitrary units)
300
R
112
R
R phase at 92 K
300
R
112
R
R phase at 216 K
Figure 3.4: Splitting of {112}R and {300}R peaks of diffraction patterns observed at 216 K and at
92 K.
24
The splitting increases with cooling, due to a change in the angle (α) of the trigonal unit
cell. This is evident from comparing the diffraction patterns obtained at 92K with that at 216 K
[55] (Figure 3.4). However, after a certain temperature, the angle (α) becomes constant and so
500
applied compressive stress (MPa)
400
300
200
100
0
0 1 2 3 4 5 6
macroscopic compressive strain (%)
Once the sample was stabilized at 92 K, it was loaded to study the associated deformation
mechanisms. An ex situ stress-strain curve (Figure 3.5) was generated at 92 K using a strain
gauge, which was used to correlate the macroscopic behavior with the microscopic behavior
25
111 011
Figure 3.6: Stress-induced transformation from R-phase to B19' phase above a stress of 68 MPa,
during loading at 92 K.
0.0002 1.1
-0.0002 0.9
strain
-0.0004 0.8
R111
-0.0006 R222 0.7
R411
-0.0008 R322 0.6
R333
-0.001 0.5
0 20 40 60 80 100
compressive stress (Mpa)
Figure 3.7: Limited twinning in the R-phase and strain redistribution above 68 MPa due to the
formation of stress-induced B19' phase during loading at 92 K.
When loaded at 92 K, the emergence of stress-induced B19' phase was noticed at a low
stress of 68 MPa (Figure 3.6) with strain redistribution (not twinning) among lattice planes in the
R-phase (Figure 3.7). This is because the propensity for R-phase to undergo a phase
transformation (due to lower free-energy of monoclinic phase under stress at 92 K) is higher than
26
twinning. Furthermore, the bulk of the R-phase was transformed to B19' phase at approximately
5% strain and the transformed B19' phase was stabilized during unloading at a holding load of 8
MPa (Figure 3.6), unlike a near complete recovery to R-phase observed at 216 K [55]. This
behavior can be attributed to the fact that the B19' phase was stabilized at a temperature lower
than the reverse transformation temperature of B19' phase to R-phase, arising from the thermal
B19' (100%)
B19' finish R start
92 K 216 K
Figure 3.8: Transformation hysteresis between B19' phase and R-phase, typical of shape memory
alloys. The dotted lines qualitatively represent the region where the experiments were
conducted.
strain recovery under a constant load to simulate the working of cyclic, low-stroke actuators
sample was strained to approximately 5% (the R-phase transformed to B19' phase in the process)
at 92 K and subsequently heated to full strain recovery under a constant load to simulate the
working of one-time, high-stroke actuators (such as in a safety valves). The constrained recovery
27
experiments of 5% strained sample showed a significant recovery of 4% strain between 150 K
and 170 K, even more significant was the bulk strain recovery of approximately 2% that took
place between 160 K and 162 K (Figure 3.9). The Rietveld refinements using GSAS [9]
suggested the phase fractions of B19' phase (space group P1121/m) and R-phase (space group
P3) at 160 K to be 83.3 vol. % and 16.7 vol. % respectively, and at 162 K to be 46.3 vol. % and
111 011
R B19'
normalized intensity (a.u.)
170K
168K
166K
164K
162K
160K
158K
156K
154K
152K
150K
2.95 3 3.05 3.1 3.15
d-spacing (Å)
Figure 3.9: Constrained recovery of the R-phase from the B19' phase during the heating of the
5% strained sample from 150 K to 170 K.
The elastic moduli of various planes in the R-phase were calculated from the lattice strain
observed (Figure 3.7), below 68 MPa, during loading at 92 K. The strain for a specific hkl plane
was calculated using equation (2.2). The estimated elastic modulus at 92 K varied between 92.9
GPa for {111} planes to 113.8 GPa for {322} planes. A macroscopic stress-strain curve (Figure
3.4) was generated from the strain gauge readings and load frame data obtained at 92 K. The
macroscopic elastic modulus (determined at stresses below 68 MPa) of the R-phase was found to
be 90.9 GPa, in agreement with that obtained from the neutron measurements. The lack of
28
twinning at 92 K was consistent with the agreement between macroscopic and microscopic
measurements [56].
One of the major contributions of the cryogenic neutron diffraction experiments was the
experimental identification of the orthorhombic B33 phase at 92 K. The orthorhombic B33 phase
has been theoretically predicted in NiTi from density function theory (DFT) calculations [57,
58], but has never been observed experimentally until now. The orthorhombic B33 phase was
identified while observing a peak movement (Figure 3.10) between {111}R and {100}B19' peaks
in the diffraction spectra collected during loading. Huang et al. [57] reported from DFT
calculations that the stress-free martensite in NiTi has a base-centered orthorhombic symmetry
(Cmcm) rather than the widely reported monoclinic symmetry (P21/m or P1121/m). They found
that the angle of the monoclinic structure associated with the minimum energy distortion is not
the experimentally reported γ ≈ 98°, but rather having γ = 107°. However, when γ = 107° it
corresponds to base centered orthorhombic symmetry (B33) (Figures 3.11 and 3.12) rather than
monoclinic symmetry. They proposed that the experimentally reported and observed monoclinic
applied stresses.
29
100 111
B19' R
111
R
normalized intensity (a.u.)
Figure 3.10: Tilting of {021}B33 during loading at 92 K, observed as a peak movement between
{111}R and {100}B19'.
B33 (Cmcm)
8 atom unit cell
B2 (Pm3m) 4 atoms/lattice point
2 atom unit cell
2 atoms/lattice point
Figure 3.11: Crystallographic correspondence between the cubic B2 and the orthorhombic B33
structures.
c b c b
a a
γ = 97 γ = 107
Figure 3.12: Left – the monoclinic B19' unit cell with γ = 97°. Right – the monoclinic B19' unit
cell with γ = 107° and the orthorhombic B33 unit cell. The green rectangle shows the
B33 {021} plane. (Figures modified from Huang et. al., Nature Materials, 2003 [57]).
30
During cooling to 92 K in the no-load condition (nominal holding stress of 8 MPa),
phase and approximately 12 vol. % orthorhombic phase. These values were estimated from
Rietveld refinements using GSAS [9] of the acquired neutron diffraction pattern. We infer that
the orthorhombic martensite was formed in stress-free areas, while monoclinic martensite in
areas that experience internal stresses [57]. Upon loading, the R-phase undergoes a stress-
induced transformation to monoclinic B19' phase. Furthermore, since the orthorhombic B33 is
unstable under stress (shear or hydrostatic) [57] certain planes of B33 phase gradually distort at
higher applied stresses. This distortion can be viewed as a change in monoclinic angle from γ =
107° to γ = 97° (Figure 3.12). Figure 3.10 shows the tilting of {021}B33 planes associated with
the distortion, in the diffraction spectra. Similar tilting was also observed in the {042}B33 planes.
The tilting of these planes starts at approximately 114 MPa and completes at approximately 210
MPa. Using Rietveld refinements, the lattice parameters for B33 at a holding stress of 8 MPa
3.4 Conclusions
Research at Temperature and Stress (SMARTS) at Los Alamos National Laboratory. The
i. The cubic B2 phase transformed to trigonal R-phase during cooling. During the
transformation, the {110}B2 peak split into {112}R and {300}R, respectively. This was
31
due to the unit cell elongation in the <111> crystallographic direction of the B2 phase
associated with the formation of R-phase. A very similar splitting was observed for the
{210}B2 peak to {303}R and {411}R, respectively. The splitting increased with cooling
and was evident from comparing the diffraction patterns obtained at 92K with that at 216
K.
ii. When loaded at 92 K, the emergence of a stress-induced B19' phase was noticed at a low
stress of 68 MPa, with strain redistribution among lattice planes in the R-phase. Bulk of
the R-phase was transformed to the B19' phase at approximately 5% strain and the
transformed B19' phase was stabilized during unloading to a holding load of 8 MPa. This
behavior was attributed to the fact that the B19’ phase was be stabilized at a temperature
lower than the reverse transformation temperature of the B19' phase to the R-phase,
iii. The constrained recovery experiments of the 5% strained sample showed a significant
recovery of 4 % strain between 150 K and 170 K, and with approximately half of that
iv. The estimated elastic modulus of lattice planes varied between 92.9 GPa for {111} planes
to 113.8 GPa for {322} planes. This was in excellent agreement with the macroscopic
elastic modulus of 90.9 GPa. The lack of twinning in the R-phase below 68 MPa was
v. A base-centered orthorhombic B33 phase was experimentally identified for the first time
in NiTi based alloys. The orthorhombic B33 phase was observed while identifying a peak
movement between {111}R and {100}B19' peaks in the diffraction spectra collected during
32
loading. During cooling to 92 K in the no-load condition (nominal holding stress of 8
monoclinic phase and approximately 12 vol. % orthorhombic phase. It was inferred that
areas that experienced internal stresses. Upon loading, the R-phase underwent a stress-
induced transformation to monoclinic B19' phase. Since the orthorhombic B33 was
unstable under stress, certain planes of the B33 phase gradually distorted at higher
applied stresses. This distortion was viewed as a change in the monoclinic angle from γ =
107° to γ = 97°. The tilting of {021}B33 planes was observed in the diffraction spectra,
which started at approximately 114 MPa and completed at approximately 210 MPa.
33
CHAPTER FOUR: CHOICE OF R-PHASE SPACE GROUP
This chapter describes a methodology to quantify the textures, strains and phase
fractions using the General Structure Analysis System (GSAS) [9] for Rietveld refinement of
neutron diffraction spectra [8]. Emphasis is placed on evaluating the choice of P3 and P 3
space groups for the R-phase in the refinements and the impact of this choice on the quantitative
analyses of spectra. Furthermore, this chapter sets the stage for using Rietveld refinement of R-
phase in NiTiFe during mechanical loading at low temperatures and the same methodology has
4.1 Introduction
Shape memory alloys are incorporated as actuator elements due to their inherent ability to
sense a change in temperature and actuate against external loads by undergoing a shape change,
NiTi-based shape memory alloys are more often used due to an advantageous combination of
material properties and shape memory behavior [12,22]. However, the B19' monoclinic
phenomenology in binary NiTi alloys exhibits hysteresis (usually between 1.5-145 K) and
typically occurs around 50-70 K above or below ambient temperature. The addition of Fe to the
NiTi system suppresses the martensitic transformation temperature and stabilizes an intermediate
trigonal R-phase. The R-phase transformation in NiTiFe alloys offers a useful window for low
34
temperature actuator operation while exhibiting reduced transformation hysteresis and a
favorable fatigue response [33]. In order to better utilize the R-phase transformation in actuator
design there is a need to understand fundamental deformation mechanisms associated with it.
characterization techniques.
temperature is uniquely suited to following the texture, strain and phase fraction evolution in a
bulk polycrystalline shape memory alloy sample [2-5]. Whereas both neutrons and X-rays (from
conventional laboratory sources) can be used for materials characterization, neutrons can
penetrate further (several mm) in to the sample and can thus provide insights that are
representative of bulk behavior. Crystal structures, residual strains, phase fractions (in multi-
phase systems) and textures are among the breadth of information that can be obtained using this
technique.
Neutron diffraction measurements have routinely been performed during in situ loading
at ambient and high temperatures. However, there have previously been no measurements
deformation in the R-phase, a capability for in situ neutron diffraction during loading at
cryogenic temperatures was implemented at Los Alamos National Laboratory for the first time
and is outlined in chapter 3. In theory, the accuracy and completeness of texture, strain and phase
fraction measurements increase with the number of measurement orientations and/or detector
area coverage. However, in practice, difficulties arise in accurately orienting large samples with
auxiliary equipment (e.g., load frame and vacuum chamber), from “shadows” cast from such
35
equipment and from errors associated with sample shifts and changes in the diffraction sampling
volume. As a result of the aforementioned considerations, the present study was conducted on
the SMARTS diffractometer at Los Alamos National Laboratory utilizing a polychromatic beam
from a spallation neutron source with detectors in two fixed orientations. There is then a need to
establish a methodology to quantitatively follow the strain, texture and phase fraction evolution
with this limited detector area coverage. The objective of this chapter is to establish such a
methodology using Rietveld refinement of neutron spectra (in the General Structure Analysis
System (GSAS) [9]) that is not merely focused on structure identification (like, e.g., a
transmission electron microscopy investigation would) but uniquely includes the combined
analyses of strain, texture and phase fraction information representative of the bulk in the
sample. Here we note that such a methodology has previously been established for the cubic
phase in NiTi [3] but is now done here for the trigonal phase.
experiment. The sample fabrication was similar to that discussed in section 3.2.3.
The neutron diffraction measurements were conducted on the SMARTS [6], using the
newly developed cryogenic loading capability [7], principally developed for the study of NiTiFe
36
shape memory alloys. Neutron diffraction spectra were collected under three experimental
The accumulated count time in each of the above experiments was 19 minutes or more
(23, 19 and 27 minutes, respectively) at a nominal beam current of 100 μA in order to obtain
As justified in a subsequent section, the initial structural parameters for the refinement
were taken from Ref. [60, 63] and the following general methodology implemented in GSAS in
i. Lattice parameters were refined following the introduction of space groups, histograms
and atom parameters (initial atom positions, fractional occupancies and thermal factors).
specimen, a constraint (sum adding to unity) was imposed for the refinement of phase
fractions.
ii. The profile coefficients, including the Gaussian variance, isotropic strain and anisotropic
strain fitting parameters were then refined. The profile function that fitted the data best
37
was a convolution of two back-to-back exponentials with a Gaussian. The profile
with the lattice parameter. In the presence of strain (i.e., the externally loaded condition),
the strain fitting parameters were refined first and then the variance. For cases where the
quantitative determination of strain was the objective, the isotropic and anisotropic strain
fitting parameters were not refined, but instead only the lattice parameter and the variance
were refined.
iii. In the next step, the presence of texture was accounted for by the use of a generalized
with, the sample orientation relative to the incident neutron and detector banks was
incremental manner such that the order used did not statistically improve the refinement
iv. Following the introduction of texture, the profile parameters were refined again, which
v. Finally, zero point error/ sample height problems were corrected if necessary.
Based on the DSC results, as expected, the phase transformation to the R-phase was
complete under the first two experimental conditions at 216 K. The stresses of 30 and 313 MPa
resulted in elastic deformation of the R-phase and not a stress-induced transformation of the R-
phase to B19' martensite. From the DSC measurements, the reverse transformation from the R-
38
phase to austenite was expected to start at 227 K and finish at 239 K. However, under a
compressive stress of 44 MPa at 237 K (the third experimental condition), the transformation is
expected to finish at a temperature higher than 239 K, which results in a substantial amount of
the R-phase remaining at 237 K as it transforms to the B2 austenite. A summary of the phases
cubic austenite phase [34-36]. However, the structure of the R-phase has been under debate for
some time. A considerable amount of research has been done in understanding the crystal
(XRD), transmission electron microscopy (TEM), synchrotron diffraction and neutron diffraction
[54, 59-65]. Hara et. al. [61] suggested the space group to be P3, employing various techniques
such as convergent beam electron diffraction (CBED), electron diffraction considering dynamic
scattering, and x-ray diffraction using Rietveld method. Schryvers and Potapov [62] utilizing
nanoprobe electron diffraction patterns recommended the structure P 3 rather than P3 (Figure
4.1), even though they obtained a better refinement with P3, considering the interatomic
distances of constituent atoms. They claim that the variation in the interatomic distances is
smaller by choosing P 3 than P3. More recently, Sitepu [64] and Khalil-Allafi [65] obtained
39
better results with P 3 , by using synchrotron and neutron diffraction data that were refined by the
Rietveld method.
Figure 4.1: Unit cell of R-phase using P3 space group (left) and P 3 space group (right).
30.0
normalized intensity
10.0 20.0
Peak positions
Difference curve
1.0 2.0 3.0
lattice plane spacing (Å)
Figure 4.2: A typical GSAS Rietveld refinement output with the R-phase refined as belonging to
the P3 space group (for spectrum obtained from the sample under a compressive
stress of 44 MPa at 237 K) for diffracting lattice planes whose normals are parallel to
the loading axis. The measured data are indicated by cross-marks and the calculated
profile is indicated by the solid-line curve. The line-marks below the profile pattern
indicate the positions of all possible Bragg reflections. The lower graph shows the
difference between the measured and calculated profile patterns.
40
30.0
normalized intensity
10.0 20.0
Peak positions
Difference curve
1.0 2.0 3.0
lattice plane spacing (Å)
Figure 4.3: A typical GSAS Rietveld refinement output with the R-phase refined as belonging to
the P 3 space group (for spectrum obtained from the sample under a compressive
stress of 44 MPa at 237 K) for diffracting lattice planes whose normals are parallel to
the loading axis. The measured data are indicated by cross-marks and the calculated
profile is indicated by the solid-line curve. The line-marks below the profile pattern
indicate the positions of all possible Bragg reflections. The lower graph shows the
difference between the measured and calculated profile patterns.
Given the ambiguity, both space groups P 3 and P3 were considered in Rietveld
refinement of spectra obtained in this study under the various experimental conditions. Figure
4.2 shows a representative GSAS Rietveld refinement output (for spectrum obtained from the
sample under a compressive stress of 44 MPa at 237 K) using the space group P3 for the R-phase
in the presence of austenite. Figure 4.3 is the corresponding output for the R-phase with the
space group P 3 . The atom positions were also refined for each of the three experimental
conditions. Tables 4.2 and 4.3 show results from refining atom positions using space groups P3
and P 3 , respectively. The comparative results of evaluating the refinement for space groups P3
41
and P 3 in terms of refinement quality of fit parameters ( χ 2 , Rwp and Rp) are given in Table 4.4.
The refinement quality of fit parameters for all three experimental conditions do not show
Table 4.2:R-phase atom positions for the three experimental conditions, refined with R-phase
belonging to the P3 space group
Parameters
Atom positions 216 K, 30 MPa 216 K, 313 MPa 237 K, 44 MPa
Refined
Ti 1(0, 0, 0) - - - -
Ti 2(1/3, 2/3, z) z 0.0792 0.0817 0.0740
Ti 3(2/3, 1/3, z) z 0.121 0.0227 0.0065
Ti 4(x, y, z) x 0.3328 0.3345 0.3375
y -0.0056 -0.0046 -0.0020
z 0.3618 0.3611 0.3545
Ti 5(x, y, z) x 0.6753 0.6784 0.6748
y 0.0094 0.0094 0.0128
z 0.6838 0.6868 0.6825
Ni 1(0, 0, z) z 0.4710 0.4685 0.4690
Ni 2(1/3, 2/3, z) z 0.5510 0.5483 0.5421
Ni 3(2/3, 1/3, z) z 0.4204 0.4183 0.4274
Ni 4(x, y, z) x 0.3186 0.3191 0.3199
y -0.0211 -0.0207 -0.0227
z 0.8580 0.8582 0.8572
Ni 5(x, y, z) x 0.6858 0.6861 0.6850
y 0.0194 0.0200 0.0189
z 0.1593 0.1605 0.1582
42
Table 4.3:R-phase atom positions for the three experimental conditions, refined with R-phase
belonging to the P 3 space group
Refined
Atom positions 216 K, 30 MPa 216 K, 313 MPa 237 K, 44 MPa
parameters
Ti 1(0, 0, 0) - - - -
Ti 2(1/3, 2/3, z) z 0.0280 0.0290 0.0236
Ti 3(x, y, z) x 0.3321 0.3313 0.3312
y -0.0070 -0.0083 -0.0069
z 0.3417 0.3416 0.3441
Ni 1(0, 0, 1/2) - - - -
Ni 2(1/3, 2/3, z) z 0.5643 0.5622 0.5589
Ni 3(x, y, z) x 0.3225 0.3228 0.3219
y -0.0194 -0.0188 -0.0199
z 0.8527 0.8539 0.8556
Table 4.4:Quality of fit parameters [9] for the two sets of Rietveld refinements (P3 and P 3 space
groups)
Experimental P3 P3
2 2
Condition χ Rwp Rp χ Rwp Rp
216 K, 30 MPa 4.495 4.88 3.61 4.586 4.93 3.63
216 K, 313 MPa 4.566 5.48 4.07 4.645 5.53 4.07
237 K, 44 MPa 5.318 5.36 3.67 5.640 5.52 3.86
used to account for the evolving texture in the R-phase and the processing induced texture in the
parent B2 austenite phase. It should be noted that refining for texture despite limited detector
area coverage may in some cases merely serve to improve the quality of the refinement rather
than represent crystallographic phenomena. Incorporating data from both detectors and
accounting for cylindrical symmetry in the sample, a 24th order spherical-harmonic description
was used. This order was arrived at by iterating from lower orders and
43
qualitatively/quantitatively examining the refinement, until increasing the order of the harmonic
The quantification of texture can be defined by a single parameter, the texture index J
[66]:
J = ∫ [ f ( g )] 2 dg (4.1)
where f ( g ) is the orientation distribution function that maps the probability of each of
the possible grain orientations with respect to the external sample dimensions and the integration
is over all orientation space. J lies between unity and infinity, corresponding to random
orientations and to a single crystal, respectively. The texture index of the R-phase was 19.4 using
the space group P3 and 25.2 with the space group P 3 for the measurements at 216 K, under a
compressive stress of 30 MPa. Correspondingly, the texture index was 35.7 using the space
group P3 and 42.1 with the space group P 3 for the measurements at 216 K, under a compressive
stress of 313 MPa. However, for the measurement at 237 K (where both R-phase and austenite
co-exist), the space group P 3 resulted in an unrealistic texture index when compared to
The space group P3 has the lowest possible symmetry in the trigonal system, while space
group P 3 has a center of symmetry. By using electron diffraction data from small single
crystalline R-phase regions, Schryvers and Potapov [62] demonstrated that atoms in the R-phase
are located in centrosymetric positions thus belonging to the space group P 3 rather than the
space group P3. This may be true considering that the diffraction measurements represented
single crystalline regions. However, for the case of neutron diffraction, the measurements are
44
representative of much larger polycrystalline areas. Following preferential phase transformation
(i.e., some variants of R-phase transforming earlier to austenite when compared to others) in the
bulk, a heterogeneous strain distribution may result that causes a loss of the center of symmetry.
This has previously been reported for ferroelectric metals [67, 68]. The space group P3 rather
than the space group P 3 better captures this loss of center of symmetry. This may explain the
unrealistically high texture index for a fit with the space group P 3 .
normalized intensity (arbitrary units)
44 MPa, 237 K
30 MPa, 216 K
313 MPa, 216 K
Figure 4.4: Section of normalized neutron diffraction spectra corresponding to R-phase {111}
lattice plane reflections (most intense) for the various experimental conditions. The
reflections shown here are from diffracting lattice planes whose normals are parallel
to the loading axis.
45
normalized intensity (arbitrary units)
313 MPa, 216 K
30 MPa, 216 K
44 MPa, 237 K
Figure 4.5: Section of normalized neutron diffraction spectra corresponding to R-phase {411}
and { 511 } lattice plane reflections (second most intense) for the various experimental
conditions. The reflections shown here are from diffracting lattice planes whose
normals are parallel to the loading axis.
Figures 4.4 and 4.5 represent the normalized intensities of the two most intense peak
reflections for the three experimental conditions. The most intense peak corresponds to the
{111} planes, while the less intense peak corresponds to the {411} and { 511 } planes. It can be
seen that the intensity of the first peak decreases while that of the second peak increases with
transformation at 237 K, the changes in intensities are relatively higher showing substantial
texture evolution. The texture evolution in the sample can be represented by the texture index as
well as axial distribution plots. While the texture index is merely a number that represents trends
in the texture evolution, a more detailed representation is given by axial distribution plots. In an
axial distribution plot, the y-axis is a measure of the number of grains that are at an angle φ
between the normal to the chosen plane and the loading axis (in this case), compared to a
46
randomly oriented polycrystal. Thus a random polycrystalline sample would be represented by a
horizontal line at unity. There is no physical significance associated with values less than zero,
which in these plots is an outcome of the global normalization procedure. Figures 4.6 and 4.7
show the axial distribution plots of the ( 111 ) plane for all the three experimental conditions,
using both space groups P3 and P 3 , respectively. Again here, the choice of the space group P 3
unrealistically overestimates the strength of the texture when correlated with individual peak
44 MPa, 237 K
25
20
times random
30 MPa, 216 K
313 MPa, 216 K
15
10
0 20 40 60 80
angle (φ)
Figure 4.6: R-phase ( 111 ) axial distribution plots (refined as belonging to the P3 space group) for
the various experimental conditions. φ is the angle between the ( 111 ) plane normal
and the loading axis.
47
44 MPa, 237 K
40
30
times random
20
30 MPa, 216 K
313 MPa, 216 K
10
0 20 40 60 80
angle (φ)
Figure 4.7: R-phase ( 111 ) axial distribution plots (refined as belonging to the P 3 space group),
for the various experimental conditions. φ is the angle between the ( 111 ) plane
normal and the loading axis.
The shift in peak positions represents (elastic) lattice strains and can be measured from
either single peak fitting or lattice parameter refinement from the Rietveld method. As mentioned
previously in chapter 2, the single peak method relies on fitting each hkl plane individually for
the lattice spacing, d hkl and the strain for a specific hkl plane is given by equation (2.2). The
three most intense peaks were used for single peak fitting, corresponding to the {111}; {411}
and { 511 }; and {222} planes, respectively. As previously described, the Rietveld method refines
lattice parameters ( a and c in this case) that inherently average all the lattice planes that diffract
in a particular bank of detectors. For strain determination, the histogram from the Bank 2
detector (representing measurements from lattice planes whose normals are parallel to the axis of
loading) was considered and only the variance (among profile coefficients) was refined with the
48
lattice parameter. The lattice parameter obtained from Rietveld refinement was used to obtain the
1 4 ⎡ h 2 + hk + k 2 ⎤ l 2
= ⎥ + c2 (4.2)
( d hkl )2 3 ⎢⎣ a2 ⎦
The strain for a specific hkl plane was then obtained using equation (2.2). This strain,
together with the increase in the applied stress, σapplied, was used to determine a plane-specific
σ applied
Ehkl = (4.3)
ε hkl
lattice planes dhkl at 30 MPa (Å) dhkl at 313 MPa (Å) E (GPa)
{111} 3.0092 2.9969 68
{411} and { 511 } 1.3409 1.3366 87
{222} 1.5043 1.4981 70
The typical error associated with d-spacing determination is ±0.0003.
Table 4.6:Strain evaluation using Rietveld approach for the two sets of refinements at 216 K (P3
and P 3 space groups)
P3 P3
hkl
d at dhkl at dhkl at dhkl at
lattice planes 30 MPa (Å) 313 MPa (Å) 30 MPa (Å) 313 MPa (Å)
E (GPa) E (GPa)
a = 7.3370 Å a = 7.3134 Å a = 7.3371 Å a = 7.3140 Å
c = 5.2616 Å c = 5.2302 Å c = 5.2611 Å c = 5.2291 Å
{111} 3.0092 2.9969 68 3.0092 2.9969 68
{411} and { 511 } 1.3407 1.3362 83 1.3408 1.3363 85
{222} 1.5046 1.4984 69 1.5046 1.4984 69
The typical error associated with lattice parameter determination is ±0.0002.
49
Tables 4.5 and 4.6 show the elastic moduli determined in the aforementioned manner for
the measurements at 216 K using both single peak and Rietveld analyses, with space groups P3
and P 3 , respectively. The results are the same for the lattice reflections considered. Here we
note that we have reported moduli from two strain measurements in order to highlight the
agreement between single peak and Rietveld approaches for strain determination. These values
were checked with additional measurements (not reported here) and remained within error.
In the presence of additional phases, the refinement for phase fraction involved setting of
phase and histogram constraints. During the phase transformation at 237 K, the refinements with
P3 showed a weight fraction of 44 vol. % for austenite (remaining R-phase), while that with P 3
was 38.50 vol. %. These values are within the estimated errors of ±3 vol. %. A qualitative
inspection of the diffraction spectra along with the thermal history suggested that the presence of
austenite was in the range of 30 to 50 vol. % at 237 K. Again, the Rietveld analyses for phase
4.4 Conclusions
Given an existing ambiguity in the published literature as to whether the trigonal R-phase
belongs to the P3 or P 3 space groups, Rietveld analyses were separately carried out
incorporating the symmetries associated with both space groups and the impact of this choice
evaluated. No statistical differences in the refinement quality were noted for structure, strain and
50
phase fraction analyses between the choice of P3 or P 3 space groups. The accuracy of the
refinement for strain determination was confirmed by comparing single peak fitting and Rietveld
approaches which resulted in comparable results for the three sets of reflections considered. The
accuracy of the texture and phase fraction determination was confirmed from visual inspection of
raw spectra. For texture analyses, the choice of the P 3 space group resulted in unrealistically
high texture indices (when compared to qualitative analyses). This was attributed to the
heterogeneous strain in the diffracting volume resulting in a loss of the center of symmetry. In
summary, this work has thus set the stage for using Rietveld refinement to quantitatively follow
the strain, texture and phase fraction evolution in the R-phase of NiTiFe during mechanical
loading at low temperatures on the SMARTS neutron diffractometer at Los Alamos National
Laboratory.
51
CHAPTER FIVE: DEFORMATION BEHAVIOR OF SHAPE MEMORY
NITIFE DURING MECHANICAL LOADING AT 92 K
This chapter discusses the deformation behavior observed during uniaxial compressive
intervals during this experiment, facilitating the quantitative monitoring of the strain, texture and
5.1 Introduction
NiTiFe based shape memory alloys are promising candidates for certain actuator
transformation from the cubic (B2) and the monoclinic (B19') phases typically observed in
equiatomic NiTi. In doing so, the monoclinic (B19') phase transformation temperatures are
lowered to the low temperature/cryogenic regime. However, little is known about the
temperatures. In order to study the deformation behavior of the R-phase, experiments were
performed at 92 K in Ni46.8Ti50Fe3.2 under external loading with the objective of following the
texture, strain and phase fraction evolution. A qualitative overview of the deformation behavior
provides a quantitative analysis of the strain, texture and phase fraction evolution based on
Rietveld refinements.
52
5.2 Experimental Procedure
experiment. The sample fabrication was similar to that discussed in section 3.2.3.
The experimental setup is explained in section 3.2.2. A compression test (rather than a tension
test, limited by the cryogenic capability) was performed at 92 K, while collecting neutron
diffraction spectra at selected stress intervals. An effective compressive holding load of 8 MPa
was applied to sample after sustaining vacuum inside the chamber. The general experimental
procedure consists of cooling the sample to 92 K and the sample was loaded as the temperature
stabilized. The sample was loaded at a rate of 0.167 MPa/s in load control mode during elastic
deformation and was switched to position control at the onset of twinning (plateau in the stress-
strain curve). The accumulated count time during each stress interval was 27 minutes at a
nominal beam current of 100 μA in order to obtain adequate intensity from a diffraction volume
of about 1 cm3. Figure 5.1 shows the stress-strain response of the NiTiFe sample, generated ex
situ under identical conditions at 92 K. The symbols on the stress-strain curve indicate the
stresses at which the in situ neutron diffraction spectra were obtained and analyzed.
53
500
300
200
100
0
0 1 2 3 4 5 6
macroscopic compressive strain (%)
Figure 5.1: Macroscopic stress-strain curve at 92 K. The symbols indicate the stresses at which
neutron diffraction spectra were obtained and analyzed.
stress of 8 MPa) revealed that the alloy was mainly R-phase, with approximately 17 vol. %
monoclinic B19' phase and approximately 12 vol. % orthorhombic B33 phase [Chapter 3].
Orthorhombic (B33) martensite was predicted as the structure possessing the lowest energy in
NiTi alloys from first-principle density function theory (DFT) calculations [57,58]. We reported
direct observation of the orthorhombic (B33) martensite, identified experimentally for the first
time [Chapter 3]. The orthorhombic (B33) martensite that distorts at higher loads can be viewed
as resulting from a change in the angle of the monoclinic unit cell from γ = 107° to 97°.
The structural parameters for the R-phase (P3 space group) were taken from Ref. [8, 60],
for the B19' phase (P1121/m space group) were obtained from Ref. [69, 70] and for the B33
phase (Cmcm space group) were obtained from Ref. [57]. The space group of the R-phase was
54
5.3 Results
B19' {100}
B19' {011}
B19' {011}
B19' {100}
R {111}
R {111}
-8 MPa -8 MPa
unloading
unloading
-125 MPa -125 MPa
normalized intensity (a.u.)
loading
loading
-160 MPa -160 MPa
-130 MPa -130 MPa
-114 MPa -114 MPa
-100 MPa -100 MPa
-82 MPa -82 MPa
-68 MPa -68 MPa
-35 MPa -35 MPa
-8 MPa -8 MPa
2.8 2.9 3 3.1 3.2 2.8 2.9 3 3.1 3.2
d-spacing (Å) d-spacing (Å)
(a) (b)
Figure 5.2: Section of normalized neutron diffraction spectra observed in both banks during
loading and unloading at 92 K. (a) The diffracting lattice planes are parallel to the
loading axis (bank 1). (b) The diffracting lattice planes are perpendicular to the
loading axis (bank 2).
Figure 5.2 shows the normalized neutron diffraction spectra collected during loading and
unloading, from both banks of detectors on SMARTS. The spectra are normalized with respect to
the area under the curve. It contrasts the preference of certain planes in the B19' phase in bank 1
and bank 2 during loading. By using neutron diffraction data from the two banks of detectors
and with a generalized spherical harmonic texture formulation in the Rietveld refinement
procedure, the volume fraction of various phases were quantified at various stress levels as
indicated in Figure 5.2. The orthorhombic (B33) martensite was taken into consideration for
Rietveld refinements until a load of 210 MPa, where the distortion was expected to be completed
[Chapter 3]. Figure 5.3 shows a typical Rietveld refinement output using GSAS. The volume
fraction of the three phases (R-phase, B19' phase and B33 phase) are presented as a function of
55
applied load in Figure 5.4. The results obtained from Rietveld refinement during the loading part
40.0 30.0
normalized intensity
10.0 20.0
Peak positions
Difference curve
1.0 2.0 3.0
lattice plane spacing (Å)
Figure 5.3: A typical GSAS Rietveld refinement output with the R, B19' and B33 phases refined
for diffracting lattice planes whose normals are parallel to the loading axis. The
measured data are indicated by cross-marks and the calculated profile is indicated by
the solid-line curve. The line-marks below the profile pattern indicate the positions of
all possible Bragg reflections. The lower graph shows the difference between the
measured and calculated profile patterns.
56
R-phase R-phase
B19' phase B19' phase
B33 phase B33 phase
100 100
80 80
volume fraction (%)
40 40
20 20
0 0
-8 -35 -68 -82 -100 -114 -130 -160 -175 -210 -227 -286 -350 -425 -425 -350 -225 -125 -8
load (MPa) load (MPa)
(a) (b)
Figure 5.4: Volume fraction of various phases obtained by Rietveld refinement as a function of
applied compressive load, (a) during loading and (b) during unloading. The typical
error associated with the volume fraction determination is ±3%.
1000 1000
loading unloading
800 800
R-phase R-phase
B19' phase B19' phase
texture index
texture index
600 600
400 400
200 200
0 0
0 -50 -100 -150 -200 -250 -300 -350 -400 -400 -350 -300 -250 -200 -150 -100 -50 0
load (MPa) load (MPa)
(a) (b)
Figure 5.5: Texture evolution in the R and B19' phases represented by the texture index during
(a) loading and (b) unloading.
57
As mentioned previously in chapter 4, the quantification of texture can be defined by a
single parameter, the texture index J [equation (4.1)]. Figure 5.5 shows the texture evolution in
the R and B19' phases quantified in terms of texture index during loading and unloading. While
the texture index is merely a number that represents trends in the texture evolution, a more
20 20
R-phase loading R-phase unloading
15 -8 MPa 15
-82 MPa
times random
5 5
-350 MPa
0 0
-425 MPa
-5 -5
0 20 40 60 80 0 20 40 60 80
angle (φ) angle (φ)
(a) (b)
Figure 5.6: R-phase ( 111 ) axial distribution plots during (a) loading and (b) unloading. φ is the
angle between the ( 111 ) plane normal and the loading axis.
58
30 30
-425 MPa
B19' phase loading -425 MPa B19' phase unloading
-350 MPa -8 MPa
25
20 -175 MPa
-82 MPa 20
times random
times random
10
15
-8 MPa
10
0
-10
0
-20 -5
0 20 40 60 80 0 20 40 60 80
angle (φ) angle (φ)
(a) (b)
Figure 5.7: B19' (100) axial distribution plots during (a) loading and (b) unloading. φ is the
angle between the (100) plane normal and the loading axis.
Figure 5.6 shows axial distribution plots for the R-phase ( 111 ) plane in NiTiFe during
loading and unloading. Similarly, Figure 5.7 shows axial distribution plots for B19' phase (100)
plane in NiTiFe during loading and unloading. These two planes were prominent and present
during the entire load-unload sequence in the bank 2 detector, where the diffracting lattice planes
are perpendicular to the loading axis. Measurements at intermediate stresses were left out in
Figures 5.6 and 5.7, for clarity. Texture evolution in the B33 phase is not reported due to the
presence of a shearing type mechanism associated with distortion in the γ angle (linked to a
change in symmetry from orthorhombic to monoclinic), rather than twinning [chapter 3].
The shift in peak positions represents lattice strains and can be measured from either
single peak fitting or lattice parameter refinement from the Rietveld method. As previously
mentioned in chapter 2, the single peak method relies on fitting each hkl plane individually for
the lattice spacing, d hkl . The strain for a specific hkl plane is then given by equation (2.2).
Strains from individual lattice plane reflections using single peak fitting are reported only for
59
separated peaks (i.e., non-overlapping) in the R and B19' phases. These correspond to {111}
planes in the R-phase and {100} planes in the B19' phase, from the bank 2 detector (representing
measurements from the diffracting lattice planes that are perpendicular to the loading axis). The
Rietveld method refines lattice parameters that inherently average all the lattice planes that
diffract in a particular bank of detectors. For strain determination, only the histogram from the
bank 2 detector was considered. The lattice parameters obtained from Rietveld refinement were
used to obtain the lattice spacing, d hkl of the aforementioned planes using [71, 72] equation (4.2)
1 1 ⎡ h2 k 2 l 2 Sin2γ 2hkCosγ ⎤
= + + − for the B19' phase (5.1)
( d hkl ) 2 Sin2γ ⎢⎣ a 2 b 2 c2 ab ⎥⎦
It should be noted that γ (instead of β) is the monoclinic angle, since the space group is P1121/m.
Figure 5.8 (a) shows the lattice strains obtained from single peak fitting and Rietveld
analysis of R-phase {111} planes during loading and unloading. Similarly, Figure 5.8 (b) shows
the lattice strains obtained from single peak fitting (SPF) and Rietveld analysis of B19' {100}
60
-500 -500
R 111 (SPF) B19' 100 (SPF)
R 111 (Rietveld) B19' 100 (Rietveld)
-400 -400
Load (MPa)
Load (MPa)
-300 -300
-200 -200
-100 b
-100 b
a
a
0 0
0 -0.0002 -0.0004 -0.0006 -0.0008 -0.001 -0.0012 0 -0.001 -0.002 -0.003 -0.004 -0.005
Strain Strain
(a) (b)
Figure 5.8: Lattice strains obtained from single peak fitting and Rietveld analysis for (a) R-phase
{111} and (b) B19' {100} planes, during loading/unloading. The arrows indicate the
loading and unloading direction. The typical errors associated with determination of
d-spacing and lattice parameters (Rietveld) are ±0.0003Å and ±0.0002Å
respectively.
GSAS [9] allows the lattice strains be evaluated in two ways, (1) by refining the lattice
parameter and (2) by refining three fitting parameters, α , β , and γ (after fixing the lattice
parameters). Since the diffraction peaks of orthorhombic B33 phase were very small compared
to the R-phase and the B19' phase, the lattice strains in the B33 phase were evaluated using the
second method. The strain for a specific hkl plane is then given by:
α β cos φ γAhkl
ε hkl = + + (5.2)
C C C
where, C denotes the diffractometer constant that converts time-of-flight data to d-spacing [9].
The first term in equation (5.2) represents the isotropic contribution of strain, while the second
and third terms represent the anisotropic contribution of strain for non-cubic and cubic
61
symmetries, respectively. The third term is irrelevant for orthorhombic symmetry and the
h1 h2 k1 k2 l1 l2
+ 2 + 2
cos φ = a2 b c (5.3)
⎛ h1 k1 l1 ⎞⎛ h2 k22 l22 ⎞
2 2 2 2
⎜⎜ 2 + 2 + 2 ⎟⎟⎜⎜ 2 + 2 + 2 ⎟⎟
⎝a b c ⎠⎝ a b c ⎠
-220
Anisotropy in (021) w.r.t. (010)
Anisotropy in (021) w.r.t. (001)
-200
-180
load (MPa)
-160
-140
-120
-100
-0.002 -0.004 -0.006 -0.008 -0.01 -0.012 -0.014 -0.016
anisotropic strain
Figure 5.9: Anisotropic strains in B33 {021} planes with respect to {010} and {001} planes,
during distortion of the B33 phase. The typical errors associated with determination
of strains are ±0.0002Å.
Figure 5.9 shows the anisotropic strains in B33 {021} planes during the distortion of the
B33 phase, calculated with respect to the {010} and {001} planes. The isotropic strains remained
62
5.4 Discussion
The macroscopic stress-strain curve generated ex situ (Figure 5.1) shows a plateau at
stresses lower than 100 MPa. Typically, a plateau in the stress-strain response of shape memory
qualitative examination of neutron diffraction spectra collected at different stress levels (Figure
5.2) confirms a stress induced transformation of the trigonal R-phase to the monoclinic B19'
phase. Furthermore, the appearance of different peaks of B19' phase at each bank of detectors
indicates the preference of variants of B19' phase for compression and tension. The diffraction
spectra of lattice planes perpendicular to the loading axis that experience compression are
acquired in bank 2, while those of lattice planes parallel to the loading axis that experience
tension are acquired in bank 1. Figure 5.2 shows that the B19' variants with {100} planes
perpendicular to the loading axis are preferred in bank 2 and B19' variants with {011} planes
parallel to the loading axis are preferred in bank 1. This is in accordance with the published
literature [73, 74] that suggests a preference for B19' {100} planes in compression and B19'
The bulk of the R-phase was transformed to the B19' phase at the maximum load of 425
MPa and the transformed B19' phase was stabilized during unloading at a holding load of 8 MPa.
This stabilization can be attributed to the fact that the B19' phase was retained at a temperature
lower than the reverse transformation temperature (above 140 K) of B19' phase to R-phase,
63
5.4.2 Phase Fraction Evolution
Figure 5.4 shows the volume fraction of the trigonal R-phase, the monoclinic B19' phase
and the orthorhombic B33 phase obtained by Rietveld refinement, plotted at various stresses
during loading and unloading. As mentioned previously, the base centered orthorhombic B33
cell can be viewed as a monoclinic cell with an angle γ = 107° [57,58]. The orthorhombic B33
phase that is inherently unstable at higher loads will distort its monoclinic angle from γ = 107° to
γ = 97° [Chapter 3]. This distortion can be regarded as a shearing type of mechanism (observed
in martensitic transformations) and starts at a load of approximately 114 MPa and completes at a
load of approximately 210 MPa. Consequently, the orthorhombic B33 phase was accounted for
in Rietveld refinements until 210 MPa. The Rietveld refinements predict the volume fraction of
During the initial part of the loading, until 68 MPa, the volume fraction of the three
transformation to monoclinic B19' phase. The maximum volume fraction of monoclinic B19'
The texture evolution in the trigonal R-phase and the monoclinic B19' phase is illustrated
by the texture index in Figure 5.5. Additionally, axial distribution plots of the ( 111 ) plane and the
(100) plane show a detailed evolution of texture respectively in Figures 5.6 and 5.7. φ is the
angle between the respective plane normal and the loading axis. Hence the distributions at φ =
0° and φ = 90° represents banks 2 and 1 respectively. In Figure 5.5, the R-phase is textured at a
64
holding load of 8 MPa. This is believed to be due to the starting texture of parent B2 austenite
phase, arising from the alloy processing. During loading, the texture of the R-phase increases
while that of the B19' decreases. Two factors can contribute to this variation. First, due to the
preferential transformation (i.e., some variants of R-phase transforming earlier when compared
to others) of R-phase to B19' phase in certain areas. If there is no preferential disappearance but
random transformation to B19' phase, the axial distribution of each stress level would overlap
one another. Second, due to the stress assisted twinning (or detwinning). This is addressed in the
next section. The decrease in the texture of the B19' phase could be the need for the variants of
B19' to satisfy strain compatibility (previously also seen in superelastic NiTi [73]).
The lattice strains obtained from single peak fitting and Rietveld refinement (Figure 5.8)
agree with each other, taking into account the errors associated with the strain determination. As
previously described, the Rietveld method refines lattice parameters that inherently average all
the lattice planes that diffract in a particular bank of detectors. Thus the strains obtained through
lattice parameter refinement using the Rietveld method will have the contribution from all the
diffraction peaks associated with a specific phase. At low stress, below 68 MPa, the R-phase
bears most of the strain (represented by ‘oa’) as illustrated in Figures 5.8 (a) and (b). In this
regime, the R-phase is elastic in nature and we reported an elastic modulus that varies between
92 GPa to 114 GPa from individual lattice planes in Chapter 3. These values are in good
agreement with the elastic modulus of 91 GPa obtained from the linear portion of the
macroscopic stress-strain curve (Figure 5.1). Upon further loading, the stress-induced
65
transformation starts and the R-phase sheds a portion of the load to the B19' phase (represented
by ‘ab’). The rest of the loading curve is comparatively linear for B19' {100} planes and bears
most of the load subsequently. Although the intensity of R-phase {111} planes is diminishing
with load, it bears certain amount of strain as seen in Figure 5.8 (a). The change in sign of the
slope is puzzling and more work has to be done to understand this behavior.
The maximum microscopic strain obtained in B19' {100} plane using single peak fitting
and the Rietveld method is approximately 0.5% at the maximum load of 425 MPa. However, the
macroscopic strain (Figure 5.1) is 10 times larger than the microscopic strain. Considering load
partitioning between the B19' phase (84% volume fraction) and the R-phase (16% volume
fraction), this disparity would be even larger. This is because the microscopic strain is elastic
where as the macroscopic strain is a mixture of elastic and inelastic strains. The inelasticity of
the macroscopic strain can be explained by means of the internal rearrangement (consequence of
energy minimization) of martensite in shape memory alloys. This internal rearrangement can
take place in two ways: (i) the realignment of differently oriented morphologies of martensite
(B19' phase and R-phase) in the favor of external stress (martensite reorientation in some cases
[19]), and (ii) the realignment of different variants within a single morphology such that
twinning or variant coalescence (detwinning in some cases [19]) takes place. Strictly speaking,
the realignment of the morphology is not possible unless the crystal structure of the matrix
surrounding the morphology, in addition to that inside the morphology, twins. In other terms,
realignment takes care of the difference (between macroscopic and microscopic strains) without
introducing dislocations (that result in permanent deformation). This correlates with the twinning
(or detwinning) observed through texture index and axial distribution plots. It has been recently
66
known that twinning decreases the macroscopic elastic modulus of the shape memory material in
the martensitic state [56]. This further suggests the disparity between macroscopic and
microscopic strains. It should be noted that twinning in B19' phase is more pronounced than that
of R-phase, due to significantly larger number of variants possible in the B19' phase [20].
Furthermore, the propensity for twinning increases with decrease in temperature (at 92 K).
During unloading there was a large strain relaxation observed in R-phase between 425 MPa and
350 MPa. Concomitantly, a surge in texture (Figure 5.5b) was observed suggesting detwinning in
the R-phase.
The inelastic part of the macroscopic strain (Figure 5.1) was not recovered during
unloading. In shape memory alloys, the residual macroscopic strain imposed on the martensite
phase can be recovered by heating. Justifying that, the macroscopic strains recovered during
heating. Bulk of that recovery took place during the B19' to R-phase transformation and the
Figure 5.9 shows the anisotropic strains associated with {021} planes in the B33
orthorhombic phase, during its distortion (between 114 MPa and 210 MPa). Tilting of {021} and
{042} planes was observed in the diffraction spectra collected at bank 2, between {111} R-phase
and {100} B19' phase planes, while loading [Chapter 3]. The intense anisotropic strains (with
isotropic strains remaining constant) suggest the shearing type mechanism associated with
67
5.5 Conclusions
The SMARTS spectrometer was used to acquire neutron diffraction spectra in situ during
neutron diffraction spectra, representative of bulk measurements, were subjected to the Rietveld
method using GSAS for quantitative determination of phase fractions, textures, and strains of the
respective phases. The neutron diffraction spectra collected under a holding load of 8 MPa
revealed that the trigonal R-phase was the dominant phase, with approximately 17 vol. %
monoclinic B19' phase and approximately 12 vol. % orthorhombic B33 phase. Upon loading, at
a stress lower than 100 MPa, the trigonal R-phase underwent a stress-induced transformation to
the monoclinic B19' phase. The volume fraction of the orthorhombic B33 phase remained
constant. However, the orthorhombic B33 phase that is unstable at higher loads distorted to the
monoclinic B19' phase. This distortion was viewed as a change in the monoclinic angle from γ =
107° to γ = 97°. The bulk of the R-phase transformed to the B19' phase at the maximum load of
425 MPa and the transformed B19' phase was stabilized during unloading to a holding load of 8
MPa. During loading, the texture of the R-phase increased while that of the B19' decreased. This
was attributed to the preferential transformation and stress assisted twinning (or detwinning).
The lattice strains obtained through single peak fitting and the Rietveld method suggested that
the R-phase carried the load until 68 MPa. Upon further loading the bulk of the load was
transferred to the evolving B19' phase. The discrepancy between macroscopic and microscopic
strains was attributed to detwinning. Additionally, the intense anisotropic strains (with isotropic
strains remaining constant) in B33 orthorhombic phase suggested a shearing type mechanism
68
CHAPTER SIX: CONSTRAINED RECOVERY EXPERIMENTS
This chapter presents a quantitative analysis of the two constrained recovery experiments
performed on Ni46.8Ti50Fe3.2 shape memory alloy samples, i.e., from the B19' phase to the R-
6.1 Introduction
Shape memory alloys combine the functions of a sensor and an actuator in a single
element, by sensing a change in temperature and actuating against external loads as a result of a
temperature-induced phase transformation [1]. In situ neutron diffraction is uniquely suited for
to phase transformations (from temperature changes) and external stresses (from bias forces).
Accordingly, bulk NiTiFe shape memory alloy samples were subjected to two constrained
recovery (recovery of macroscopic strain under a constant stress) experiments in the context of
real world engineering actuator applications. First, with the objective of examining NiTiFe in
approximately 5% (the R-phase was transformed to B19' phase in the process) at 92 K and
subsequently heated to full strain recovery under a load. Second, with the objective of
sample was strained to 1% at 92 K and subsequently heated to full strain recovery under a load.
Neutron diffraction spectra were recorded at suitable temperature intervals during these
69
experiments that assisted in the monitoring of the phase-specific texture and volume fraction
chapter 3. This chapter provides a quantitative analysis of the phase fraction and texture
constrained recovery experiments. The sample fabrication was similar to that discussed in
section 3.2.3.
K. The neutron diffraction spectra collected under no-load condition (nominal holding stress of 8
monoclinic B19' phase and approximately 12 vol. % orthorhombic B33 phase [Chapter 3].
During the application of stress (up to a macroscopic strain of 5%), the trigonal R-phase
was unloaded and held at a constant stress of 50 MPa. The stress-induced monoclinic B19' phase
along with macroscopic strain (deformation) was retained during unloading. Thereafter the
70
sample was heated (at a constant stress of 50 MPa) through the B19' phase to R-phase
collected at selected temperature intervals. The accumulated count time during each temperature
interval was 27 minutes at a nominal beam current of 100 μA in order to obtain adequate
In the second experiment, a Ni46.8Ti50Fe3.2 shape memory alloy compression sample was
constant load of 50 MPa. Thereafter the sample was heated (at a constant load of 50 MPa)
through the R-phase to B2 phase transformation range (231–243 K) in a controlled manner, with
neutron diffraction spectra collected at selected temperature intervals. As with the previous
experiment, the accumulated count time during each temperature interval was 27 minutes at a
nominal beam current of 100 μA in order to obtain adequate intensity from a diffraction volume
of about 1 cm3.
The structural parameters for the R-phase (P3 space group) were taken from Ref. [8, 60]
and the B19' phase (P1121/m space group) were taken from Ref. [69, 70]. The space group of the
71
6.3 Results and Discussion
Figure 6.1 shows the recovery of the R-phase from the B19' phase under a constant load
of 50 MPa. Here a section of the neutron diffraction spectra (normalized with respect to area) are
offset with respect to increasing temperature. The intensities of {011} B19' phase planes
diminish while that of {111} R-phase planes intensify, during heating. The displacement from
the SMARTS load frame (Figure 6.2) displayed a 1 mm deflection (for 24 mm sample length)
during heating from 150 K to 170 K. This corresponds to approximately 4% bulk recovery in
strain (the machine compliance need not be taken into account, since the load remains the same
and CTE is negligible). Even more significant was the bulk strain recovery of approximately 2%
170K
168K
166K
164K
162K
160K
158K
156K
154K
152K
150K
2.9 2.95 3 3.05 3.1 3.15 3.2
d-spacing (Å)
Figure 6.1: Constrained recovery of the R-phase from B19' phase during the heating of the 5%
strained NiTiFe sample from 150 K to 170 K.
72
63.2
63
62.8
displacement (mm)
62.6
62.4
62.2
62
61.8
150 155 160 165 170
temperature (K)
Figure 6.2: Displacement in the SMARTS load frame as a function of temperature during heating
of the 5% strained NiTiFe sample from 150 K to 170 K.
B19' phase
R-phase
100
80
volume fraction (%)
60
40
20
0
150 152 154 156 158 160 162 164 166 168 170
temperature (K)
Figure 6.3: Phase fraction evolution of the B19' and R phases during constrained recovery of the
5% strained NiTiFe sample from 150 K to 170 K.
Figure 6.3 shows the phase fraction evolution of the B19' phase and the R-phase during
the constrained recovery from 150 K to 170 K obtained through Rietveld refinements using
GSAS [9]. The phase fraction is quantified in terms of volume fraction of the phases. The
73
Rietveld refinements suggest that the volume fractions of the B19' phase and the R-phase at 160
K to be 83.3 vol. % and 16.7 vol. %, respectively and at 162 K to be 46.3 vol. % and 53.7 vol. %,
The texture evolution in the B19' and R phases, represented by texture index is shown in
Figure 6.4. It shows that the nascent texture of the R-phase is high and decreases with increasing
phase fraction. This is due to the preferential transformation (i.e., some variants of the B19'
phase transforming earlier when compared to others) of the B19' phase to the R-phase in certain
areas within the sample. At the same time, the texture of the B19' phase is lower initially and
increases with decreasing phase fraction. Again, as the transformation is preferential certain
variants of the B19' phase get transformed to the R-phase when compared to others. A more
detailed evolution in texture is portrayed by axial distribution plots, in terms of sample space.
Figure 6.5 show the axial distribution plots of (a) B19' (100) plane and (b) R-phase ( 111 ) plane,
350
B19'
300 R-phase
texture (times random)
250
200
150
100
50
0
150 155 160 165 170
temperature (K)
Figure 6.4: Texture evolution in the B19' and R phases represented by texture index during
constrained recovery of the 5% strained NiTiFe sample from 150 K to 170 K.
74
15 7
B19' phase 160 K R-phase
6
10 150 K 156 K
156 K
5 162 K
160 K
5 162 K 166 K
4
times random
times random
166 K 170 K
170 K
3
0
2
-5
1
0 150 K
-10
-1
-15
0 20 40 60 80 0 20 40 60 80
angle (φ) angle (φ)
(a) (b)
Figure 6.5: (a) B19' (100) and (b) R-phase ( 111 ) axial distribution plots planes during
constrained recovery of the 5% strained NiTiFe sample from 150 K to 170 K. φ is
the angle between the corresponding plane normal and the loading axis.
This constrained recovery of the R-phase from the B19' phase (especially between 160 K
and 162 K) validates the use of NiTiFe in high-stroke, actuator applications such as in safety
valves. In such devices, the actuation required is one-time (for most cases) with sufficient stroke
as the temperature approaches a pre-determined value. This value depends on the composition of
the shape memory material, along with the history of thermo-mechanical treatments.
75
6.3.2 Constrained Recovery from R-Phase to B2 Phase
B2 {210}
B2 {110}
normalized intensity (a.u.)
normalized intensity (a.u.)
243K
243K
241K
241K
239K
239K
237K 237K
235K 235K
233K
R {411}
R {303}
233K
R {112}
R {300}
231K 231K
1.32 1.33 1.34 1.35 1.36 1.37 1.38 2.08 2.1 2.12 2.14 2.16 2.18
d-spacing (Å) d-spacing (Å)
(a) (b)
Figure 6.6: Constrained recovery of the B2 phase from the R-phase during heating of the 1%
strained NiTiFe sample from 231 K to 243 K, (a) showing R{411} and R{303}
combining to form B2{210} and (b) showing R{300} and R{112} combining to form
B2{110}.
Figure 6.6 shows the recovery of the B2 phase from the R-phase under a constant load of
50 MPa. Here two sections of the neutron diffraction spectra (normalized with respect to area)
are offset with respect to increase in temperature. In Figure 6.6(a), the R-phase {303} planes and
the R-phase {411} planes combine to form the B2 {210} planes. Similarly in Figure 6.6(b), the
R-phase {112} planes and the R-phase {300} planes combine to form the B2 {110} planes.
Figure 6.7 shows the phase fraction evolution of the R-phase and the B2 phase during
constrained recovery from 231 K to 243 K obtained through Rietveld refinements using GSAS
[9].
76
R-phase
B2 phase
100
80
40
20
0
231 233 235 237 239 241 243
temperature (K)
Figure 6.7: Phase fraction evolution of the R and B2 phases during constrained recovery of the
1% strained NiTiFe sample from 231 K to 243 K.
The texture evolution in the R and B2 phases, represented by the texture index is shown
in Figure 6.8. The texture of the R-phase is lower initially and increases with decreasing phase
fraction. This is due to the preferential transformation (i.e., some variants of the R-phase
transforming earlier when compared to others) of the R-phase to the B2 phase in certain areas
within the sample. Given that the B2 phase has only one variant when compared to the R-phase,
the texture in the B2 phase is fairly constant during the recovery. A more detailed evolution in
texture is portrayed by the axial distribution plots, in terms of sample space. Figure 6.9 show the
texture evolution represented by axial distribution plots of (a) R-phase ( 111 ) plane and (b) B2
(100) plane, during constrained recovery of the sample from 231 K to 243 K.
77
1600
1400 R-phase
Austenite
1200
800
600
400
200
0
230 232 234 236 238 240 242 244 246
temperature (K)
Figure 6.8: Texture evolution in the R and B2 phases represented by texture index during
constrained recovery of the 1% strained NiTiFe sample from 231 K to 243 K.
60 25
243 K 231 K
R-phase 235 K B2 phase
50
20
239 K
40 239 K
15 243 K
30
times random
times random
235 K
231 K
20 10
10
5
0
-10
-20 -5
0 20 40 60 80 0 20 40 60 80
angle (φ) angle (φ)
(a) (b)
Figure 6.9: (a) R-phase ( 111 ) planes and (b) B2 (100) axial distribution plots during constrained
recovery of the 1% strained NiTiFe sample from 231 K to 243 K. φ is the angle
between the corresponding plane normal and the loading axis.
78
This constrained recovery of the B2 phase from the R-phase simulates the functioning of
the SMA in an actuator against a bias force. Further it validates the use of NiTiFe in cyclic, low-
stroke, actuator application such as a thermal conduction switch. In such devices, the make or
6.4 Conclusions
The SMARTS spectrometer was used to acquire neutron diffraction spectra in situ during
two constrained recovery experiments performed on Ni46.8Ti50Fe3.2 shape memory alloy samples,
i.e., from the B19' phase to the R-phase and the R-phase to the B2 phase. First, with the objective
strain recovery under a load. Second, with the objective of examining NiTiFe in cyclic, low-
and subsequently heated to full strain recovery under a load. The neutron diffraction spectra,
representative of bulk measurements, were subjected to Rietveld method using GSAS for
quantitative determination of phase fractions and textures of the respective phases. The bulk
recovery in strain for the first experiment recorded approximately 4% strain recovery between
150 K and 170 K, with a significant strain recovery of approximately 2% that took place between
160 K and 162 K. The Rietveld refinements suggested the volume fractions of the B19' phase
and the R-phase at 160 K to be 83.3 vol. % and 16.7 vol. %, respectively and at 162 K to be 46.3
vol. % and 53.7 vol. %, respectively. The texture evolution exhibit a preferential transformation
of the B19' to the R-phase. The bulk recovery in strain for the second experiment recorded
79
approximately 1% between 231 K and 243 K, with a preferential transformation from the R to
the B2 phase.
80
CHAPTER SEVEN: EXTENSION OF RESEARCH METHODOLOGY TO
NITIPD SHAPE MEMORY ALLOYS
chapters to experiments performed on Ni29.5Ti50.5Pd20 shape memory alloys using in situ neutron
Neutron diffraction spectra were recorded at appropriate intervals during these experiments,
facilitating the monitoring of the strain, texture and phase fraction evolution and thus the
7.1 Introduction
Although shape memory alloys (SMAs) have potential in room temperature and low
temperature applications, the potential for high temperature applications is enormous when
appropriate alloys are engineered. This is especially true for the automobile, aerospace and
turbo-machinery (power generation) related applications. Recently there have been substantial
efforts for the development of high temperature shape memory alloys (HTSMAs) by ternary and
quaternary additions of elements such as Au, Pt, Pd, Hf, or Zr to the binary NiTi alloy system.
The integration of HTSMAs into the devices operating at high temperatures can improve the
performance and efficacy of such devices. For example, with the help of HTSMAs, aerospace
engines can be designed to operate at improved efficiencies under a variety of flight conditions.
While most HTSMAs exhibit shape memory behavior under stress-free conditions,
relatively fewer have the capability to produce work output (for functioning as actuators) under
81
the influence of external load. These include bulk NiTiPt [75] and NiTiPd [76] alloys, and
NiTiHf thin films [77,78]. Furthermore, the useful work output in these alloys decreases with an
increase in transformation temperatures above 573 K. This reduction in work output has been
attributed to the deterioration of mechanical properties in the austenite phase [79]. The austenitic
yield strength becomes lower than the martensitic yield strength with higher transformation
temperatures, resulting in “shape setting” the material at the new stress level. It should be noted
that “shape setting” in shape memory alloys is performed by heating the confined shape memory
element (under stress) to a temperature where the applied stress is greater than the yield strength
that cause deterioration in mechanical behavior. Such an understanding can help in fabricating
HTSMAs that can function satisfactorily at temperatures above 573 K. Amidst various in situ
uniquely suited to following the texture, strain and phase fraction evolution in a bulk
polycrystalline shape memory alloy sample [2-5]. The high temperature loading capability at
For the first time Ni29.5Ti50.5Pd20 high temperature shape memory alloys were subjected to
Los Alamos National Laboratory. Two sets of experiments were preformed on these alloys. In
the first set, the experiments consisted of applying a constant stress while thermally cycling
between room and elevated temperatures. Two cycles each were performed at stresses of 100,
200 and 300 MPa, respectively. In the second set, the stress induced martensitic transformation
82
was investigated from the austenite phase by mechanically cycling at elevated temperatures.
Neutron diffraction spectra were recorded at appropriate intervals during these experiments,
facilitating the monitoring of the strain, texture and phase fraction evolution and thus the
diameter by 23.9 mm length were supplied by NASA-GRC for the neutron diffraction
experiments. The alloy was made from high purity elemental constituents by vacuum induction
melting in a graphite crucible. The melt was cast in a cylindrical copper mold of 25.4 mm
diameter by 102 mm length. The ingot was then homogenized in a vacuum furnace at 1323 K for
72 hrs. The ingot was subsequently hot extruded at 1173 K into the required diameter and sliced
The elevated temperature loading capability (Figure 7.1) at SMARTS [6] was used for
the neutron diffraction studies on Ni29.5Ti50.5Pd20. Two thermocouples were mounted on the
samples for temperature measurement. The output from the thermocouples was used to control
the temperature of the sample using LabVIEW software. The accumulated count time during
83
each neutron window was 24 minutes at a nominal beam current of 100 μA in order to obtain
Figure 7.1: High temperature furnace on the SMARTS load frame [6].
Given the lack of prior literature on NiTiPd martensite structure, a decision was made to
separately refine (using the Rietveld method) diffraction spectra from martensite using the
monoclinic B19' structure as well as the orthorhombic B19 structure. The structural parameters
for the B19' (P1121/m space group) were taken from Ref. [66,67] and that of the B19 phase
(Pmcm space group) were taken from Ref. [82]. Figures 7.2 and 7.3 show the Rietveld
refinement outputs using GSAS with the monoclinic B19' structure and the orthorhombic B19
84
structure, respectively. Using the Rietveld refinement procedure, the monoclinic B19' structure
γ=90.42°, against typical values of a=2.89Å, b=4.64Å, c=4.12Å, α=β=90° and γ=97° for a binary
NiTi alloy. The refinement quality of fit parameters were χ 2 =3.504, Rwp=6.94 and Rp=5.06. The
orthorhombic B19 lattice parameters were determined to be a=2.797Å, b=4.687Å, c=4.426Å and
α=β=γ=90°. The refinement quality of fit parameters were χ 2 =3.571, Rwp=6.19 and Rp=4.50. It
should be noted that the monoclinic angle (γ) of the B19' unit cell approached 90°, validating that
the structure is indeed orthorhombic. Furthermore, the crystallographic residual factors (Rwp and
Peak positions
-0.5
Difference curve
Figure 7.2: A typical GSAS Rietveld refinement output with B19' structure refined for
diffracting lattice planes whose normals are parallel to the loading axis. The measured
data are indicated by cross-marks and the calculated profile is indicated by the solid-
line curve. The line-marks below the profile pattern indicate the positions of all
possible Bragg reflections. The lower graph shows the difference between the
measured and calculated profile patterns.
85
1.5
1.0
normalized intensity
Peak positions
-0.5
Difference curve
Figure 7.3: A typical GSAS Rietveld refinement output with B19 phase refined for diffracting
lattice planes whose normals are parallel to the loading axis. The measured data are
indicated by cross-marks and the calculated profile is indicated by the solid-line
curve. The line-marks below the profile pattern indicate the positions of all possible
Bragg reflections. The lower graph shows the difference between the measured and
calculated profile patterns.
The load-bias experiment was performed on the NiTiPd sample by applying selected
loads and then thermally cycling between room temperature and a temperature of (Af + 25) K.
The austenite finish (Af) temperature was qualitatively recognized by identifying the
disappearance of martensite peaks during heating (diffraction spectra were collected at 5 minute
intervals). Two thermal cycles each were performed at compressive stresses of 100, 200 and 300
MPa, respectively in sequence and a last cycle was performed at 100 MPa. Before the start of
each cycle, the sample was unloaded to a holding load (7 MPa) and reloaded again. Neutron
86
diffraction spectra were collected at the holding load and applied load at room temperature, at the
applied load at (Af + 25) K, and at the applied load back to room temperature, in each cycle.
B19 {100}
B19 {011}
normalized intensity (a.u.)
Figure 7.4: Room temperature measurements at the applied load in each cycle, before heating.
These spectra are from the bank 2 detector, where the diffracting lattice planes are
perpendicular to the loading axis.
Figure 7.4 shows the qualitative texture evolution from selected normalized diffraction
spectra as a function of each cycle in the martensitic phase (room temperature), before heating.
The sample has a starting texture where majority of the B19 {011} planes are orientated
perpendicular to the loading axis. This is due to the processing induced texture from hot
extrusion. With cycling as observed in the bank 2 detectors, where the diffracting planes are
perpendicular to the loading axis, the intensity of B19 {100} planes increase at the expense of
B19 {011} planes. This behavior is typical of monoclinic B19' martensite in NiTi based alloys,
where the {100} planes are oriented perpendicular to compression and {011} planes are oriented
87
perpendicular to tension [73,74]. Furthermore, monoclinic B19' and orthorhombic B19 are
B19 {100}
B2 {100}
normalized intensity (a.u.)
Figure 7.5: High temperature measurements for each cycle. These are from the bank 2 detector
where the diffracting lattice planes are perpendicular to the loading axis.
Figure 7.5 shows a section of the normalized diffraction spectra obtained at high
temperature for each cycle. The load and the corresponding temperature at which the
measurements were made are indicated along with the cycle number. A limit of 673 K (400° C)
was put on the maximum temperature to avoid macroscopic strains being permanently set (or
intense peaks corresponding to the B19 {100} planes. With each cycle there is a peak shift as
well as peak broadening. At the same stress there seems to be considerable amount of strain
developing with cycling. Additionally, the temperature for (Af + 25) K is increasing with the
88
As previously mentioned, the texture evolution in a cylindrical sample can be represented
by axial distribution plots. Figure 7.6 shows the axial distribution plots of (a) B19 (100) planes
and (b) B19 (011) planes, corresponding to residual B19 martensite at high temperature for
cycles 5 to 7. The axial distribution plots show that there is slight change in texture in the
(100) (011)
B19 B19
15 12
times random
4
5
0
0
-4
-5
0 20 40 60 80 0 20 40 60 80
angle (φ) angle (φ)
Figure 7.6: (a) B19 (100) and (b) B19 (011) axial distribution plots at the maximum temperatures
correspond to cycles 5, 6 and 7. φ is the angle between the corresponding plane
normal and the loading axis.
As previously mentioned in chapter 2, the shift in peak positions represents lattice strains
and can be measured from either single peak fitting or lattice parameter refinement from the
Rietveld method. The single peak method relies on fitting each hkl plane individually for the
lattice spacing, d hkl and the strain for a specific hkl plane is given by equation (2.2).
The Rietveld method refines lattice parameters that inherently average all the lattice
planes that diffract in a particular bank of detectors. For strain determination, only the histogram
89
from the bank 2 detector was considered. As previously discussed in chapter 5, GSAS [9] allows
the lattice strains be evaluated in two ways, (1) by refining the lattice parameter and (2) by
refining three fitting parameters, α , β , and γ (after fixing the lattice parameters). The strain
The first term in equation (5.2) represents the isotropic contribution of strain, while the
second and third terms represent the anisotropic contribution of strain for non-cubic and cubic
symmetries, respectively. The second term is irrelevant for cubic symmetry and Ahkl in the third
h2 k 2 + h2 l 2 + k 2 l 2
Ahkl = (7.1)
(h 2
+ k2 + l2 )
2
Table 7.1 shows the isotropic strains calculated using the Rietveld method corresponding
to various cycles with respect to cycle 1 for the B2 phase. The contributions from the coefficient
Table 7.1:Isotropic strains in the B2 phase with respect to cycle 1. The contributions of the
coefficient of thermal expansion (CTE) have been subtracted out.
90
Cycle 1
Cycle 7
normalized intensity
Figure 7.7: Section of normalized neutron diffraction spectra showing strains between the B2
{100} planes for cycles 1 and 7, at the maximum temperature. These spectra are from
the bank 2 detector, where the diffracting lattice planes are perpendicular to the
loading axis.
Table 7.1 show high strains developed during the load-bias cycling. A qualitative
representation of the strain can be obtained from Figure 7.7 that shows a peak shift in the B2
{100} planes between cycle 1 and 7. These strains may develop due to the mismatch between the
lattices, under applied load and the corresponding high stresses may explain the deterioration in
mechanical properties such as yield strength in the austenite phase at high temperatures. Further
analysis of strain within various lattice planes that accounts for the anisotropy is shown in Table
7.2. Here the lattice strains corresponding to cycle 5 are determined on the basis of cycle 1.
91
Table 7.2:Lattice strains in the B2 phase for cycle 5 (300 MPa, 659 K) determined with respect
to cycle 1 (100 MPa, 464 K).
samples at 498 K. Figure 7.7 shows the macroscopic response curve for two cycles in the stress-
induced martensite experiment. It shows the first cycle with a plateau and residual macroscopic
-700
-600
Compressive Stress (MPa)
-500
-400
-300
-200
1st Cycle
-100
2nd Cycle
0
32 31.5 31 30.5 30 29.5
Position (mm)
92
B19 {100}
B2 {100}
normalized intensity (a.u.)
7 MPa
Cycle 2
650 MPa
600 MPa
7 MPa
Cycle 1
650 MPa
600 MPa
7 MPa
Figure 7.9: Development of stress-induced martensite during cycling at 498 K. These spectra are
from the bank 2 detector, where the diffracting lattice planes are perpendicular to the
loading axis.
Figure 7.8 shows selected normalized diffraction spectra as a function of cyclic loading.
The first cycle shows stress-induced martensite at higher loads and residual martensite upon
unloading. Furthermore, there is a considerable shift in the B2 {100} planes before and after
loading that signifies residual strain. This residual strain can be expected to stabilize the stress-
induced martensite upon unloading. In the second cycle, the residual martensite is consistent
7.4 Conclusions
The SMARTS spectrometer was used to acquire neutron diffraction spectra in situ during
93
alloys. The neutron diffraction spectra, representative of bulk measurements, were subjected to
the Rietveld method using GSAS for quantitative determination of phase fractions, textures, and
strains of the respective phases. Using the initial spectra collected from the room temperature
martensite, Rietveld refinements were performed separately considering the monoclinic B19'
structure and the orthorhombic B19 structure. The monoclinic angle (γ) with B19' approached
90°, validating that the structure is indeed orthorhombic. Furthermore, the crystallographic
residual factors (Rwp and Rp) were better for the orthorhombic B19 structure. During load-bias
experiments, the intensity of the B19 {100} planes increased at the expense of the B19 {011}
planes for room temperature measurements on bank 2 (the diffracting planes are perpendicular to
the loading axis). Cycles 5 through 7 showed residual martensite at elevated temperatures
indicated by less intense peaks representing B19 {100} planes. Peak shifts as well as peak
broadening were reported with cycling at elevated temperature. The axial distribution plots show
that there is a slight change in texture at various orientations (along φ ) in the residual martensite
from cycles 5 through 7. The strain measurements using Rietveld and single peak fits indicate
development of high strains during the load-bias cycling. These strains may develop due to the
mismatch between the lattices under applied load. The high stresses associated with these strains
may explain the deterioration in mechanical properties such as yield strength in the austenite
Ni29.5Ti50.5Pd20 compression samples at 498 K, showed the first cycle with a plateau and residual
macroscopic strain, and the second cycle with a more linear response. The first cycle showed
stress-induced martensite at higher loads and residual martensite upon unloading. There was a
considerable shift in the B2 {100} planes before and after loading, signifying residual strain.
Residual stress was expected to stabilize the stress-induced martensite upon unloading. The
94
residual martensite from the first cycle was consistent with the macroscopic linear response in
95
CHAPTER EIGHT: ACTUATOR DEVICE DEVELOPMENT
This chapter focuses on the development of NiTi based shape memory alloy actuators.
The NiTi helical springs in the thermal conduction switch described in the author’s master’s
thesis [17] was replaced with NiTiFe springs and tested. Following that work and Lemanski [84],
a design recommendation (patent pending) was made to minimize the heat gradient in the SMA
mechanism was pursued. Such a mechanism can be potentially considered for use in debris-less
8.1.1 Introduction
Shape memory alloys (SMAs) are used as actuator elements due to their intrinsic ability
to sense a change in temperature and actuate against external loads (e.g., a bias spring) by
undergoing a shape change. This ability, called the shape memory effect (SME), is attributed to a
first-order, thermoelastic, martensitic phase transformation that takes place during the
in an associated strain recovery against large stresses (e.g., as high as 500 MPa), making them
superior actuator materials. In NiTi SMAs, this phase transformation usually takes place between
96
a monoclinic, so-called martensite phase and a cubic, so-called austenite phase. The temperature
at which the phase transformation takes place is sensitive to composition and thermomechanical
processing [83]. Furthermore, the addition of a third element in small quantities promotes a
significant shift in the transformation temperatures and the possible introduction of an additional
intermediate phase. The addition of Fe to the NiTi system shifts the martensitic transformation to
Typically, there is hysteresis between the forward and reverse transformations. The
transformation hysteresis is a result of elastic strain energy dissipation, the energy associated
with frictional resistance to interface motion and similar dissipative processes [31,32]. The cubic
to the very low transformation hysteresis, they are useful in designing actuators that operates
between narrow temperature ranges. However, the maximum recoverable strain (1%
(8% approximately) [37]. Furthermore, the fatigue life of the R-phase is superior when compared
The motivation for this work arises from the National Aeronautics and Space
Administration (NASA) Kennedy Space Center’s (KSC) requirement for thermal management at
cryogenic temperatures. The initial objective was focused on NASA’s efforts to support
methane liquefaction for future Mars missions [1,17,33]. However, recently the scope was
extended to address the requirements of future lunar missions. The original objective was to
design, construct and test an SMA thermal conduction switch to facilitate thermal conduction of
97
approximately 8 watts between two liquid reservoirs held at 118 K and 92 K (boiling points of
liquid methane and liquid oxygen, respectively). This switch is expected to control the liquid
methane temperature and pressure in a zero boil-off system by allowing on-demand heat transfer
between two reservoirs kept at separate temperatures, in an efficient and autonomous manner.
The first prototype switch to demonstrate proof-of-concept was developed by Droney et al. in 2003
that used a commercially available NiTi alloy, and operated between an ice-water mixture and hot
water [86]. Subsequently, a low temperature version of the conduction switch that used NiTi helical
springs was developed by Krishnan et al. in 2003 [1,17]. A third version of the switch using NiTiFe
alloy strips for the actuator element was developed and tested by Lemanski et al. in 2005. That
switch operated in the low temperature range and used the R-phase but exhibited limited stroke [33].
The cryogenic range thermal switches currently employed range from gas gap and liquid
gap thermal switches that rely on convective heat transfer between two surfaces to externally
actuated thermal switches. The sensors and active controls in such systems make them more
complicated and expensive, yet less efficient than the proposed switch. Furthermore, gas gap
switches are restricted to long cycle times, tend to exhibit poor thermal isolation in their open
state and have low heat transfer ratios between open and closed states. Other systems using
conduction bands make use of mechanical means to generate sufficient thermal contact and may
not be reliable. SMA thermal switches have the potential to limit these problems.
This work, while addressing the extension to possible use on lunar missions, reports on
modifying the previously reported helical spring switch [1,17] by using NiTiFe as the SMA
element in order that significant stroke coupled with low thermal hysteresis can be achieved.
98
8.1.2 Shape Memory Alloy Actuators
SMAs exhibit significantly lower apparent modulus in their martensite phase compared
to the austenite phase. This apparent lowering of modulus in the martensite phase is not solely
The lower symmetry of martensite, compared to the parent austenite, generates multiple variants
that under go deformation twinning. The design of SMA elements for actuators makes use of this
macroscopic difference in modulus between the austenite and martensite states [21,87].
holds promise as it possesses reduced transformation hysteresis. However, the shape recovery
associated with the R-phase transformation (approximately 1% strain) is significantly lower than
that of the monoclinic phase transformation (approximately up to 8% strain). The use of helical
SMA springs can compensate for this design limitation as springs produce greater stroke when
compared to straight elements such as thin strips and wires [33]. Another advantage of helical
springs is that the stress distribution in the helical spring is more uniform when compared to
strips used in a bending mode (e.g., Ref. [33]) where there is a higher stress concentration in the
middle of the element. A non-uniform stress distribution in the element can reduce the fatigue
life and additionally increase the hysteresis associated with the phase transformation.
Current NASA plans for a crewed lunar outpost are expected to be targeted around the
lunar south pole where direct exposure to the sun occurs about 70% of the time and temperature
extremes are moderate. The temperature varies from approximately 120 K to 160 K during a
99
period of 28 earth days [88]. This region has elevated quantities of hydrogen and possibly water
ice at the bottoms of deep craters. The lunar outpost is expected to incorporate in situ resource
utilization approaches and the switch developed in this work has the potential to be used for
liquefaction of oxygen and zero boil-off control during day/night cycles. Other potential uses
include operations where variable heat transfer is required, including residual propellant
scavenging, chill down of equipment, and long term storage of ascent module propellants.
convective heat transfer, heat leakage on the Moon will be through radiation. Designs for the
switch need to account for this radiative heat exchange to ensure the switch accurately senses the
system temperature. Also, there is a potential issue regarding lunar dust that could increase
friction in mechanisms, degrade seals and add significant thermal resistance between contact
surfaces.
Ni47.07Ti49.66Fe3.27 wire of 0.216 cm diameter was selected for the SMA springs. Following
SMA spring theory [21,87], the NiTiFe wire was set into 3 helical springs of mean diameter 2.6
cm and 4 turns in order to obtain a maximum recovery force of 16 N per spring and a stroke of
0.5 cm, limiting the strain to within 1%. The shape setting was carried out at 773 K for 20
100
heating
cooling
Figure 8.1: Differential scanning calorimeter (DSC) response of the NiTiFe wire used.
were performed on the NiTiFe wire before and after the shape-setting procedure (Figure 7.1).
The samples were tested from 300 K to 230 K using a Perkin-Elmer Diamond DSC at a rate of 0.33
Ks-1 under nitrogen cover gas. The results show the austenite to R-phase transformation during
cooling and the reverse transformation during heating. The first sample, represented by the solid black
line, is the NiTiFe wire (as received) before any heat treatment. The second sample, represented by
the dashed line, received a heat treatment of 773 K for 20 minutes during the shape-setting procedure
and was subsequently ice-water quenched. The start and finish of the austenite to R-phase
transformation and the corresponding reverse transformation from the R-phase to austenite were
determined to be 268, 244, 254 and 277 ± 2 K, respectively, for NiTiFe wire before heat
treatment. The corresponding temperatures for the NiTiFe wire after shape-setting were 265, 247,
101
262 and 272 ± 2 K. The transformation to martensite was below 120 K and was hence outside the
operating range of the calorimeter used. The temperature hysteresis did not change appreciably
during the shape-setting procedure. However, a slightly sharper peak was observed after the heat
treatment for both the austenite to R-phase transformation and the reverse transformation. Even
though the transformation temperatures of the composition currently selected are different from the
exact temperatures required in final application, the goal here is the validation and demonstration of
the use of the R-phase in actuating helical spring based switches. Furthermore, ongoing work at the
University of Central Florida that focuses on tailoring the composition of NiTiFe for specific
cryogenic applications shows that R-phase transformations can be lowered to the cryogenic regime by
modifying the Ni/Ti ratio and Fe content with suitable thermomechanical treatments [89].
In order to evaluate the performance of the switch, testing was conducted using liquid
nitrogen to actuate the NiTiFe SMA springs. The purpose of the experiment was to verify the switch
performance in terms of the temperature hysteresis and to obtain an initial estimate of the magnitude
of the displacement of the switch. In order to determine the hysteresis, data for both the heating and
cooling portions of the thermal cycle were recorded. Temperature vs. displacement data were
obtained and are shown in Figure 7.2. Figure 7.3 shows the switch in the closed (extended) position
at room temperature and in the open (contracted) position at a temperature of 233 K. A contraction of
102
1
displacement (mm)
-1
cooling
heating
-2
-3
-4
-5
180 200 220 240 260 280 300
temperature (K)
(a) (b)
Figure 8.3: The NiTiFe helical spring switch in, (a) the closed (extended) position at a room
temperature of 298 K and (b) the open (contracted) position at 233 K.
displacement reported in Ref. [33] (with NiTiFe strips in a bending mode) was approximately 1
mm [33]. This enhanced displacement will help in achieving superior thermal isolation in the
open state of the switch. Additionally, the contact force can also be better tailored over the
103
displacement range. Furthermore, this substantiates the earlier claim of superior stroke with a
helical element instead of a bending element. The temperature vs. displacement graph also
confirms a small temperature hysteresis despite thermal gradients that develop across the SMA
elements during thermal cycling of the switch that result in a non-uniform phase transformation
The author and Dr. Raj Vaidyanathan supervised a senior design team (2006-7) at UCF,
in developing a NiTi shape memory based release mechanism [11]. This mechanism was
methodologies rely on pyrotechnic charges that break hold-down bolts, and chemical motors that
separate the solid rocket boosters during shuttle launches. In direct comparison to existing
iii. Release of heavy loads by employing mechanical advantage – in the prototype a shape
memory alloy actuator that is in the form of a spring (0.085 inch diameter wire, 5 turns,
mean internal diameter and length of about 0.6”) removes a holding pin that in turn
causes a spring-loaded bearing to rotate and release a hitch that holds as large as 1000 lbs
104
v. Modular and fail-safe design that separates the actuator module from the load-bearing
(a) (b)
Figure 8.4: The shape memory alloy release mechanism in, (a) before actuation and (b) after
actuation, releasing a 100 lb load.
Figure 8.4 shows the shape memory alloy based release mechanism releasing a 100 lb
load. The details of the mechanism can be found in Ref. [11] and a patent has been filed.
8.3 Conclusions
The fabrication and testing of an SMA switch using Ni47.07 Ti49.66 Fe3.27 helical springs
has been presented as part of ongoing efforts to develop and improve upon the design of SMA
thermal conduction switches for thermal management at cryogenic temperatures. Such a switch
can provide on-demand heat transfer between two reservoirs at different temperatures, to meet
NASA’s requirements for advanced spaceport applications. The utilization of a cubic to trigonal
105
Additionally, the utilization of helical NiTiFe SMA springs produces superior stroke (order of 4
mm) compared to straight elements (order of 1 mm) such as thin strips and wires, and provides
improved thermal isolation in the open state. The contact force can also be better tailored over the
displacement range. Furthermore, a more uniform stress distribution in the helical springs when
compared to strips (used in a bending mode) increases the fatigue life and decreases the
hysteresis. As with previous versions of the switch, the thermal gradients that develop across the
SMA elements during thermal cycling of the switch have proven to be problematic. The testing of
another prototype of the switch that minimizes the thermal gradient within the SMA element
incorporating a key configuration change is currently underway. The details of the switch are not
The development of a NiTi based release mechanism was pursued. Such a mechanism
can be potentially considered for use in debris-less separation mechanisms at room temperature
to replace current pyrotechnic based release mechanisms, including booster separation chemical
motors. A prototype was implemented using NiTi, as the author and Dr. Raj Vaidyanathan
supervised a senior design team project at UCF. Patents are being filed for all the three actuator
106
CHAPTER NINE: CONCLUSION
conclusion was included at the end of each chapter. This chapter summarizes all of them
9.1 Conclusion
A novel cryogenic capability was implemented by the UCF research group on the
Spectrometer for Materials Research at Temperature and Stress (SMARTS) at Los Alamos
National Laboratory that can vary temperatures between 300 K and 90 K, while collecting
neutron diffraction spectra in situ during loading. The deformation behaviors of Ni46.8Ti50Fe3.2
shape memory alloys were studied for the first time at 92 K using this capability.
i. The room temperature cubic B2 phase transformed to the trigonal R-phase during
cooling. During the transformation, the {110}B2 peak split into {112}R and {300}R,
respectively. This was due to the unit cell elongation in the <111> crystallographic
direction of the B2 phase associated with the formation of the R-phase. A very similar
splitting was observed for the {210}B2 peak to {303}R and {411}R, respectively. The
splitting increased with cooling and was evident while comparing the diffraction patterns
ii. When loaded at 92 K, the emergence of a stress-induced B19' phase was noticed at a low
stress of 68 MPa, with strain redistribution among lattice planes in the R-phase. Bulk of
the R-phase transformed to the B19' phase at approximately 5% strain and the
107
transformed B19' phase was stabilized during unloading to a holding load of 8 MPa. This
behavior was attributed to the fact that the B19’ phase is stabilized at a temperature lower
than the reverse transformation temperature of the B19' phase to the R-phase, arising
iii. The estimated elastic modulus of lattice planes varied between 92.9 GPa for {111} planes
to 113.8 GPa for {322} planes. This was in excellent agreement with the measured
macroscopic elastic modulus of 90.9 GPa. The lack of twinning in the R-phase below 68
MPa was consistent with this agreement between macroscopic and microscopic
measurements.
iv. The base-centered orthorhombic B33 phase was experimentally identified for the first
time in NiTi based alloys. The orthorhombic B33 phase was observed while identifying a
diffraction peak shifting between the {111}R and {100}B19' peaks in the diffraction
(nominal holding stress of 8 MPa), Ni46.8Ti50Fe3.2 consisted primarily of the R-phase with
phase. It was inferred that orthorhombic martensite was formed in stress-free areas, while
monoclinic martensite formed in areas that experienced internal stresses. Upon loading,
the orthorhombic B33 was unstable under stress, certain planes of B33 phase gradually
distorted at higher applied stresses. This distortion was viewed as a change in the
monoclinic angle from γ = 107° to γ = 97°. The tilting of the {021}B33 planes was
108
observed in the diffraction spectra, which started at approximately 114 MPa and finished
v. Given an existing ambiguity in the published literature as to whether the trigonal R-phase
belongs to the P3 or P 3 space groups, Rietveld analyses were separately carried out
incorporating the symmetries associated with both space groups and the impact of this
choice evaluated. No statistical differences in the refinement quality were noted for
structure, strain and phase fraction analyses between the choice of P3 or P 3 space
groups. The accuracy of the refinement for strain determination was confirmed by
comparing single peak fitting and Rietveld approaches which resulted in comparable
results for the three sets of diffraction patterns considered. The accuracy of the texture
and phase fraction determination was confirmed from visual inspection of raw spectra.
For texture analyses, the choice of the P 3 space group resulted in unrealistically high
texture indices (when compared to qualitative analyses). This was attributed to the
symmetry.
vi. The neutron diffraction spectra acquired in situ during uniaxial compression loading of
Ni46.8Ti50Fe3.2 shape memory alloy at 92 K, were subjected to the Rietveld method using
GSAS for quantitative determination of phase fractions, textures, and strains of the
respective phases. The neutron diffraction spectra collected under the no-load condition
phase with approximately 17 vol. % monoclinic B19' phase and approximately 12 vol. %
orthorhombic B33 phase. Upon loading, at a stress lower than 100 MPa, the trigonal R-
109
phase underwent a stress-induced transformation to the monoclinic B19' phase. The
volume fraction of the orthorhombic B33 phase remained constant, until distortion to
monoclinic B19' phase. The bulk of the R-phase was transformed to the B19' phase at the
maximum load of 425 MPa and the transformed B19' phase was stabilized during
unloading to a holding load of 8 MPa. During loading the texture of the R-phase
increased while that of the B19' decreased. This was attributed to the preferential
transformation and stress assisted detwinning. The lattice strains obtained through single
peak fitting and the Rietveld method suggested that the R-phase carried the load until 68
MPa. Upon further loading, the bulk of the load was transferred to the evolving B19'
phase. The discrepancy between the macroscopic and microscopic strains was attributed
strains remaining constant) in the B33 orthorhombic phase suggested a shearing type
in Figure 9.1.
vii. Neutron diffraction spectra was acquired in situ during two constrained recovery
experiments performed on Ni46.8Ti50Fe3.2 shape memory alloy samples, i.e., from B19'
phase to R-phase and R-phase to B2 phase. First, with the objective of examining NiTiFe
under a load. Second, with the objective of examining NiTiFe in cyclic, low-stroke,
and subsequently heated to full strain recovery under a load. The neutron diffraction
110
spectra, representative of bulk measurements, were subjected to the Rietveld method
using GSAS for quantitative determination of phase fractions and textures of the
respective phases. The bulk recovery in strain for the first experiment recorded
approximately 4% strain recovery between 150 K and 170 K, with a strain recovery of
approximately 2% that took place between 160 K and 162 K. The Rietveld refinements
suggested the volume fractions of the B19' phase and the R-phase at 160 K to be 83.3 vol.
% and 16.7 vol. %, respectively and at 162 K to be 46.3 vol. % and 53.7 vol. %,
respectively. The texture evolution exhibit a preferential transformation of the B19' phase
to the R-phase. The bulk recovery in strain for the second experiment recorded
111
Figure 9.1: Behavior of Ni46.8Ti50Fe3.2 shape memory alloy between room temperature and 92 K.
The research methodology for Ni46.8Ti50Fe3.2 shape memory alloys was applied to
were acquired in situ during load-bias and stress-induced martensite experiments performed on
Ni29.5Ti50.5Pd20 shape memory alloys. The neutron diffraction spectra, representative of bulk
measurements, were subjected to the Rietveld method using GSAS for quantitative determination
of phase fractions, textures, and strains of the respective phases. Using the initial spectra
collected from the room temperature martensite, Rietveld refinement was separately performed
considering the monoclinic B19' structure and the orthorhombic B19 structure. The monoclinic
112
angle (γ) with B19' approached 90°, validating that the structure is indeed orthorhombic.
Furthermore, the crystallographic residual factors (Rwp and Rp) were better for the orthorhombic
B19 structure. During load-bias experiments, the intensity of the B19 {100} planes increased at
the expense of the B19 {011} planes for room temperature measurements in bank 2 (the
diffracting planes are perpendicular to the loading axis). Cycles 5 through 7 showed residual
martensite at elevated temperature indicated by small peaks representing B19 {100} planes. Peak
shifts as well as peak broadening was reported with cycling at elevated temperature. The axial
distribution plots show that there is slight change in the texture in the residual martensite from
cycles 5 through 7. The strain measurements using Rietveld and single peak fits indicate
development of high strains during the load-bias cycling. These strains may develop due to the
mismatch between the lattices under applied load. The high stresses associated with these strains
may explain the deterioration in the mechanical properties such as yield strength in the austenite
Ni29.5Ti50.5Pd20 compression samples at 498 K, showed the first cycle to have a plateau in the
stress-strain curve and residual macroscopic strain, and the second cycle with a more linear
response. The first cycle showed stress-induced martensite at higher loads and residual
martensite upon unloading. There was a considerable shift in the B2 {100} planes before and
after loading, signifying residual strain. The resulting residual stress was expected to stabilize the
stress-induced martensite upon unloading. The residual martensite from the first cycle was
The fabrication and testing of an SMA switch using Ni47.07 Ti49.66 Fe3.27 helical springs
has been presented as part of ongoing efforts to develop and improve upon the design of SMA
thermal conduction switches for thermal management at cryogenic temperatures. Such a switch
113
can provide on-demand heat transfer between two reservoirs at different temperatures, to meet
NASA’s requirements for advanced spaceport applications. The utilization of a cubic to trigonal
Additionally, the utilization of helical NiTiFe SMA springs produces superior stroke (order of 4
mm) compared to straight elements (order of 1 mm) such as thin strips and wires, and provides
improved thermal isolation in the open state. The contact force can also be better tailored over the
displacement range. Furthermore, a more uniform stress distribution in the helical springs when
compared to strips (used in a bending mode) increases the fatigue life and decreases the
hysteresis. As with previous versions of the switch, the thermal gradients that develop across the
SMA elements during thermal cycling of the switch have proven to be problematic. The testing of
another prototype of the switch that minimizes the thermal gradient within the SMA element
incorporating a key configuration change is currently underway. The details of the switch are not
The development of a NiTi based release mechanism was pursued. Such a mechanism
can be potentially considered for use in debris-less separation mechanisms at room temperature
to replace current pyrotechnic based release mechanisms, including booster separation chemical
motors. A prototype was implemented using NiTi, as the author and Dr. Raj Vaidyanathan
supervised a senior design team project at UCF. Patents are being pursued for all the three
114
REFERENCES
[1] V.B. Krishnan, J.D. Singh, T.R. Woodruff, W.U. Notardonato and R. Vaidyanathan, in:
U. Balachandran (Ed.), Adv. in Cryo. Eng., 50 (2004) pp. 26-33.
[2] M.A.M. Bourke, R. Vaidyanathan, and D.C. Dunand, Appl. Phys. Lett., 69 (1996) 2477.
[3] R. Vaidyanathan, M.A.M. Bourke and D.C. Dunand, J. Appl. Phys. 86 (1999) 3020.
[4] R. Vaidyanathan, M.A.M. Bourke and D.C. Dunand, Acta. Mater. 47 (1999) 3353.
[5] R. Vaidyanathan, D.C. Dunand, and U. Ramamurty, Mater. Sci. Eng., A, 289 (2000) 208.
[6] M.A.M. Bourke, D.C. Dunand, and E. Ustundag, Appl. Phys. A, 74 (2002) S1707.
[7] T. Woodruff, V.B. Krishnan, V. Livescu, B. Clausen, D.W. Brown, T. Sisneros, M.A.M.
Bourke, and R. Vaidyanathan, Review of Scientific Instruments (to be submitted).
[8] V.B. Krishnan, R. Mahadevan Manjeri, B. Clausen, D.W. Brown, and R. Vaidyanathan,
Mater. Sci. Eng., A (in press).
[9] A.C. Larson, and R.B. Von Dreele, General Structure Analysis System (GSAS), Los
Alamos National Laboratory LAUR 8-748, 1986.
[10] V.B. Krishnan, C. Bewerse, W.U. Notardonato and R. Vaidyanathan, Adv. in Cryo. Eng.
(2008, in press).
[11] N.S. Lam, K.D. Schoenwald, D.W. Snyder, and D. S. Watson, Senior Design Report,
University of Central Florida (2007).
[12] H. Funakubo, Shape Memory Alloys, Gordon and Breach Science Publishers, New York,
1987.
[13] L.M. Schetky, “Shape Memory Alloy Applications in Space Systems”, in Engineering
Aspects of Shape-Memory Alloys, edited by T. W. Duerig, K. N. Melton, D. Stöckel and
C. M. Wayman, Butterworth-Heinemann, London, 1990, pp.170-178.
115
[14] K. Uchino, “Shape Memory Ceramics”, in Shape Memory Materials, edited by K.
Otsuka, and C. M. Wayman, Cambridge University Press, 1998, pp.184-202.
[17] V.B. Krishnan, Master Thesis, “Design, Fabrication and Testing of a Shape Memory
Alloy based Cryogenic Thermal Conduction Switch”, University of Central Florida
(2004).
[19] K. Gall, J. Tyber, V. Brice, C.P. Frick, H.J. Maier and N. Morgan, Journal of Biomedical
Materials Research, 75A (2005) 810.
[21] I. Ohkata and Y. Suzuki, “The design of shape memory alloy actuators and their
applications”, Shape Memory Materials, Cambridge University Press, Cambridge, UK,
1998.
[22] T.W. Duerig and A.R. Pelton, “Ti-Ni Shape Memory Alloys”, Material Properties
Handbook: Titanium Alloys, ASM International, 1994, pp.1035-1048.
[25] K.N. Melton, “Ni-Ti Based Shape Memory Alloys”, in Engineering Aspects of Shape
Memory Alloys, edited by T.W. Duerig, K.N. Melton, D. Stöckel and C.M. Wayman,
1990, pp.21-35.
[26] J.H. Mulder, P.E. Thoma and J. Beyer, Z. Metalkunde, 84 (1993) 501.
[27] P.E. Thoma, M. Kao, S. Fariabi and D.N. Abyudom, Proceedings of ICOMAT’ 92,
(1993) pp.917-922.
116
[28] D.E. Hodgson, “Using Shape Memory Alloys”, Shape Memory Applications, 1988.
[29] D. Stöckel, Actuator’92, 3rd International Conference on New Actuators, (1992) pp.79-
84.
[30] Y. Suzuki and H. Horikawa, Materials Research Society Symposium Proceedings, Vol.
246 (1991) pp.389-398.
[32] C.R. Rathod, B.C. Clausen, M.A.M. Bourke, and R. Vaidyanathan, Appl. Phys. Lett. 88
(2006) 201919-3.
[33] J.L. Lemanski, V.B. Krishnan, R. Mahadevan Manjeri, W.U. Notardonato and R.
Vaidyanathan, in: U. Balachandran (Ed.), Adv. in Cryo. Eng., 52 (2006) pp. 3-10.
[34] V.N. Khachin, Y.I. Paskal, V. E. Gunter, A.A. Monasevich, and V.P. Sivokha, Phys. Met.
Metallogr. (1978) 46.
[36] M.B. Salamon, M.E. Meichle, and C.M. Wayman, Phys. Rev. B, 31 (1985) 7306.
[41] W.B. Cross, A.H. Kariotis, and F.J. Stimler, NASA CR-1433, September 1969.
[43] M. Matsumoto and T. Honma, JIM International symposium (Kobe) (1976), pp.199-204
[44] C.M. Wayman, L. Cornelis, and K. Shimizu, Scripta Mater. 6 (1972) 115.
117
[45] F.E. Wang, B.F DeSavage, W.J. Buehler, and W.R. Hosler, J. Appl. Phys., 39 (1968),
2166.
[47] M.T. Hutchings, P.J. Withers, T.M. Holden and T. Lorentzen (Eds.), Introduction to the
Characterization of Residual Stress by Neutron Diffraction, Taylor & Francis, 2005.
[50] R.A. Young (Ed.), The Rietveld Method, Oxford University Press, 1995.
[55] S. Shmalo, PhD dissertation (in progress), “In situ Neutron Diffraction Investigation of
NiTiFe Shape Memory Alloys during Mechanical Loading at Cryogenic and Room
Temperatures”, University of Central Florida (2008).
[56] S. Rajagopalan, A.L. Little, M.A.M. Bourke and R. Vaidyanathan, Appl. Phys. Lett. 86,
(2005) 081901.
[57] X. Huang, G.J. Ackland and K.M. Rabe, Nature Mat. 2 (2003) 307.
[58] J.R. Morris, Y. Ye, M. Krcmar and C.L. Fu, Mater. Res. Soc. Symp. Proc. 980 (2007).
[61] T. Hara, T. Ohba, E. Okunishi, and K. Otsuka, Mater. Trans., JIM, 38 (1997) 11.
[62] D. Schryvers and P.L. Potapov, Mater. Trans., JIM, 43 (2002) 774.
118
[64] H. Sitepu, J.P. Wright, T. Hansen, D. Chateigner, H.G. Brokmeier, C. Ritter, and T.
Ohba, Mater. Sci. Forum, 495-497 (2005) 255.
[65] J. Khalil-Allafi, W.W. Schmahl and D.M. Toebbens, Acta. Mater. 54 (2006) 3171.
[66] E.F. Sturcken and J.W. Croach, Trans. Metal. Soc. A.I.M.E, 227 (1963) 934.
[67] P.W. Anderson and E.I. Blount, Phys. Rev. Lett. 14 (1965) 217.
[68] I.A. Sergienko, V. Keppens, M. McGuire, R. Jin, J. He, S.H. Curnoe, B.C. Sales, P.
Blaha, D.J. Singh, K. Schwarz, and D. Mandrus, Phys. Rev. Lett., 92 (2004) 065501-1.
[70] Y. Kudoh, M. Tokohami, S. Miyazaki, and K. Otsuka, Acta Metall., 33 (1985) 2049.
[72] C. Suryanarayana and M.G. Nortan, X-ray Diffraction: A Practical Approach (Plenum
Press, New York, 1998).
[73] R. Vaidyanathan, M.A.M. Bourke and D.C. Dunand, Met. Mat. Trans. A 32A (2001)
777.
[74] C.R. Rathod, A. Little, D. Brown, M.A.M. Bourke, and R. Vaidyanathan, Acta Mater. (to
be submitted).
[75] R. Noebe, D. Gaydosh, S. Paula II, A. Garg, T. Biles, and M. Nathal, SPIE Conf. Proc.
5761 (2005), pp. 364-375.
[76] R. Noebe, S. Paula II, G. Bigelow, O. Rios, A. Garg, and B. Lerch, SPIE Conf. Proc.
6170 (2006), pp. 617010-1-617010-13.
[77] G. K. Rasmussen et al., “Process for Deposition of Sputtered Shape Memory Alloy
Films,” U.S. Patent 6,454,913 (2002).
119
[79] S. Paula II, G. Bigelow, R. Noebe, D. Gaydosh, and A. Garg, Proc. Int. Conf. on Shape
Memory and Superelastic. Technologies 2006 (in press).
[80] G. Bigelow, Master Thesis, “Effects of Palladium Content, Quaternary Alloying, and
Thermomechanical Processing on the Behavior of NiTiPd Shape Memory Alloys for
Actuator Applications”, Colorado School of Mines (2006).
[81] R. Noebe, S. Draper, D. Gaydosh, A. Garg, B. Lerch, N. Penney, G. Bigelow, S. Paula II,
and J. Brown, Proc. Int. Conf. on Shape Memory and Superelastic. Technologies 2006 (in
press).
[84] J. L. Lemanski, Master Thesis, “Cryogenic Shape Memory Alloy Actuators for Spaceport
Technologies: Materials Characterization and Prototype Testing”, University of Central
Florida (2005).
[85] S. Miyazaki, “Thermal and Stress Cycling Effects and Fatigue Properties of Ni-Ti
Alloys”, in Engineering Aspects of Shape-Memory Alloys, edited by T. W. Duerig, K. N.
Melton, D. Stöckel and C. M. Wayman, Butterworth-Heinemann, London, 1990, pp. 394-
413.
[87] T. Waram, “Design Principles For Ni-Ti Actuators”, in Engineering Aspects of Shape-
Memory Alloys, edited by T. W. Duerig, K. N. Melton, D. Stöckel and C. M. Wayman,
Butterworth-Heinemann, London, 1990, pp. 234-244.
[88] P. Eckart, The Lunar Base Handbook, McGraw-Hill Companies Inc., 1999.
[89] R. Mahadevan Manjeri, M.C. Mistretta, and R. Vaidyanathan, Acta Mater. 2007 (to be
submitted).
120