Вы находитесь на странице: 1из 146

Comprehensive Characterization of

Measurement Data Gathered by the Pressure


Tube to Calandria Tube Gap Probe

by

Shaddy Samir Zaki Shokralla

A thesis submitted to the


Department of Physics, Engineering Physics and Astronomy
in conformity with the requirements for
the degree of Doctor of Philosophy

Queen’s University
Kingston, Ontario, Canada
August 2016

Copyright
c Shaddy Samir Zaki Shokralla, 2016
Abstract

Multi-frequency eddy current measurements are employed in estimating pressure tube


(PT) to calandria tube (CT) gap in CANDU fuel channels, a critical inspection ac-
tivity required to ensure fitness for service of fuel channels. In this thesis, a com-
prehensive characterization of eddy current gap data is laid out, in order to extract
further information on fuel channel condition, and to identify generalized applications
for multi-frequency eddy current data. A surface profiling technique, generalizable
to multiple probe and conductive material configurations has been developed. This
technique has allowed for identification of various pressure tube artefacts, has been
independently validated (using ultrasonic measurements), and has been deployed and
commissioned at Ontario Power Generation. Dodd and Deeds solutions to the electro-
magnetic boundary value problem associated with the PT to CT gap probe configu-
ration were experimentally validated for amplitude response to changes in gap. Using
the validated Dodd and Deeds solutions, principal components analysis (PCA) has
been employed to identify independence and redundancies in multi-frequency eddy
current data. This has allowed for an enhanced visualization of factors affecting gap
measurement. Results of the PCA of simulation data are consistent with the skin
depth equation, and are validated against PCA of physical experiments. Finally,
compressed data acquisition has been realized, allowing faster data acquisition for

i
multi-frequency eddy current systems with hardware limitations, and is generalizable
to other applications where real time acquisition of large data sets is prohibitive.

ii
Co-Authorship

The following people contributed to the development of this original thesis:


First author:
Shaddy S. Z. Shokralla, Queen’s University
Supervisors:
Prof. Thomas W. Krause, NDE Development, Royal Military College of Canada
(RMC)
Prof. Jordan Morelli, Queen’s University
This thesis work combines manuscripts of three separate journal publications. All
manuscript preparations, scientific programming, mathematical formulations, data
analysis, and training of analysts in using the tools produced via this thesis work,
were performed by the first author. The following provides a list of all co-authors
and their contributions to respective manuscripts.

Comprehensive Characterization of Measurement Data Gathered by the


Pressure Tube to Calandria Tube Gap Probe

Funded by: Ontario Power Generation


Outcomes:
Manuscript I: S. Shokralla, T. W. Krause, and J. Morelli, “Surface profiling with high

iii
density eddy current non-destructive examination data,” NDT & E International, vol.
62, March 2014, pp. 153-159.

Manuscript II: S. Shokralla, S. Sullivan, J. Morelli, and T. W. Krause, “Modelling


and validation of Eddy current response to changes in factors affecting pressure tube
to calandria tube gap measurement,” NDT & E International, vol. 73, July 2015,
pp. 15-21.

Contributions to Manuscript II:


T. W. Krause: Identification of resistivity - gap probe amplitude relations.
S. Sullivan: Author of “SurfaceCalc” eddy current modeling software.

Manuscript III: S. Shokralla, J. E. Morelli and T. W. Krause, “Principal Components


Analysis of Multifrequency Eddy Current Data Used to Measure Pressure Tube to
Calandria Tube Gap,” IEEE Sensors Journal, vol. 16, no. 9, May 2016, pp. 3147-
3154.

iv
Acknowledgments

I would like to sincerely thank Dr. Thomas W. Krause for his dedication to supporting
this thesis work. Dr. Krause provided steadfast guidance at all critical junctures in
this endeavor and gave both the appropriate academic and industrial context to make
this work impactful in both academic and industrial spheres.
I would also like to thank Dr. Jordan E. Morelli for providing exemplary counseling
and direction to ensure successful completion of doctoral requirements at Queen’s
University. Dr. Morelli’s commitment to rigour ensured manuscripts published as a
product of this thesis were prepared with quality.
Managers of the IMS NDE Projects group at Ontario Power Generation were
instrumental in not only their support of derivation and publication of this work, but
also its field implementation.
My parents, the late Samir Shokralla and Wasima Shokralla instilled an intellec-
tual curiosity in me at a young age, and provided encouragement in pursuit of this
endeavor.
I am most grateful to my loving wife Sally Shehata for her steadfast support and
sacrifice during the period that this work was undertaken. During this period, we
married and were blessed with the birth of our daughter Alexandra. Without Sally’s
continued patience and constancy, this would not have been realized.

v
Contents

Abstract i

Co-Authorship iii

Acknowledgments v

Contents vi

List of Tables ix

List of Figures x

List of Acronyms xiv

Chapter 1: Introduction 1
1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Need for Comprehensive Characterization of Measurement Data Gath-
ered by Gap Probe . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.2.1 Presence of Pressure Tube Artefacts and Relationship to Gap
Probe Lift-off . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.2.2 Identifying Redundancies in Multifrequency Eddy Current Data 12
1.2.3 Employment of an Analytical Model of Responses to Factors
Affecting PT to CT Gap Measurement . . . . . . . . . . . . . 14
1.3 Objective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.4 Thesis Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

Chapter 2: Literature Review 18


2.1 Presence of Pressure Tube Artefacts and Relationship to Gap Probe
Lift-off . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.2 Identifying Redundancies in Multifrequency Eddy Current Data . . . 20
2.3 Employment of an Analytical Model of Eddy Current Responses to
Factors Affecting PT to CT Gap Measurement . . . . . . . . . . . . . 22

vi
Chapter 3: Theory 24
3.1 Mathematical Gradient and Directional Derivative . . . . . . . . . . . 25
3.2 Equivalent Circuit and Impedance Plane Representation . . . . . . . 26
3.3 Maxwell’s Equations and the Skin Depth Equation . . . . . . . . . . 30
3.4 Dodd and Deeds Solutions . . . . . . . . . . . . . . . . . . . . . . . . 32
3.5 Principal Components Analysis . . . . . . . . . . . . . . . . . . . . . 38
3.5.1 Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

Chapter 4: Introduction to Manuscripts 41

Chapter 5: Surface Profiling with High Density Eddy Current Non-


Destructive Examination Data 43
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
5.2 Inspection Requirement . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.3 Measurement Technique . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.4 Lift-off Location Extraction . . . . . . . . . . . . . . . . . . . . . . . 49
5.5 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
5.5.1 Mechanical Wear Marks . . . . . . . . . . . . . . . . . . . . . 51
5.5.2 Pressure Tube Constrictions . . . . . . . . . . . . . . . . . . . 54
5.5.3 Spacers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
5.5.4 Interpretation of Positive and Negative Intensities . . . . . . . 57
5.5.5 Confirmation of Spacer and Constriction Locations using Ul-
trasound Data . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
5.6 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
5.6.1 Relationship to C-Scan Display . . . . . . . . . . . . . . . . . 60
5.6.2 Separability of Lift-off Localization in Rotary and Axial Direc-
tions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
5.6.3 Generalization of Results . . . . . . . . . . . . . . . . . . . . . 62
5.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

Chapter 6: Modelling and Validation of Eddy Current Response to


Changes in Factors Affecting Pressure Tube to Calan-
dria Tube Gap Measurement 65
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
6.2 Measurement Technique . . . . . . . . . . . . . . . . . . . . . . . . . 67
6.2.1 Description of Samples . . . . . . . . . . . . . . . . . . . . . . 71
6.3 Analytic Determination of Eddy Current Responses to Changes in
Physical Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
6.3.1 Modelling Requirements . . . . . . . . . . . . . . . . . . . . . 73
6.4 Validation of Analytically Determined Responses . . . . . . . . . . . 74
6.4.1 Validation of Responses to Changes in Wall Thickness . . . . 74
vii
6.4.2 Validation of Responses to Changes in Resistivity . . . . . . . 78
6.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
6.5.1 Modelling Parameter Continuity . . . . . . . . . . . . . . . . . 81
6.5.2 Parallel Plate Approximation . . . . . . . . . . . . . . . . . . 84
6.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

Chapter 7: Principal Components Analysis of Multi-frequency Eddy


Current Data Used to Measure Pressure Tube to Ca-
landria Tube Gap 87
7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
7.2 Pressure Tube to Calandria Tube Gap Measurement . . . . . . . . . . 90
7.2.1 Analytical Model . . . . . . . . . . . . . . . . . . . . . . . . . 92
7.3 Principal Components Analysis . . . . . . . . . . . . . . . . . . . . . 92
7.4 PCA Applied to Analytical Model Data . . . . . . . . . . . . . . . . 94
7.4.1 PCA Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
7.4.2 Relationship to Skin Depth Equation . . . . . . . . . . . . . . 97
7.5 Compressed Data Acquisition . . . . . . . . . . . . . . . . . . . . . . 97
7.6 Experimental Validation . . . . . . . . . . . . . . . . . . . . . . . . . 102
7.6.1 Wall Thickness Variation . . . . . . . . . . . . . . . . . . . . . 103
7.6.2 Resistivity Variation . . . . . . . . . . . . . . . . . . . . . . . 104
7.6.3 Differences Between Experimental and Analytical PCA . . . . 107
7.7 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
7.7.1 Normalization of Phase and Scale . . . . . . . . . . . . . . . . 107
7.7.2 Practical Implementation . . . . . . . . . . . . . . . . . . . . . 108
7.8 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

Chapter 8: Discussion 111

Chapter 9: Conclusions 114

Chapter 10: Future Work 117

Bibliography 120

viii
List of Tables

6.1 Modelled Pressure Tube and Calandria Tube Parameters. . . . . . . . 74


6.2 Statistical Computations for 4 kHz, 8 kHz, and 16 kHz Frequencies. . 76
6.3 Resistivity and Wall Thickness Parameters for Measured Composite
Resistivity Pressure Tubes. . . . . . . . . . . . . . . . . . . . . . . . . 79
6.4 Statistical comparisons between modeled and measured results. . . . 80

7.1 Modelled Pressure Tube and Calandria Tube Parameters. . . . . . . . 94


7.2 Mean and standard deviation of percentage voltage error. . . . . . . . 102

ix
List of Figures

1.1 CANDU reactor face with array of fuel channel end-fittings [1]. . . . . 2
1.2 Heat transport system with primary components shown [1]. . . . . . . 3
1.3 CANDU reactor assembly with heat transport sub-assemblies [1]. . . 4
1.4 CANDU reactor assembly with fuel channel components magnified [1]. 5
1.5 Fuel channel components - garter spring spacer separates PT and CT
[1]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.6 Principle of eddy current testing. Primary magnetic field induces eddy
currents in conductive surface secondary magnetic field that opposes
primary field. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.7 PT to CT gap measurement configuration showing minimum and max-
imum gap locations and eddy current (EC) probe that traverses the
circumference of the PT ID. . . . . . . . . . . . . . . . . . . . . . . . 8
1.8 Method for generation of estimated pressure tube to calandria tube gap. 9
1.9 Eddy current (EC) probe riding over pressure tube surface protrusion
with probe degrees of motion indicated. . . . . . . . . . . . . . . . . . 11
1.10 Impedance plane response of probe lift off from pressure tube surface.
Direction of ˆl gives direction of impedance plane probe lift-off response
starting from contact. . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

x
3.1 Model of a transmit and receive coil with a test object. . . . . . . . . 27
3.2 Equivalent parallel circuit of coil and test sample represented without
and with equivalent impedance. Zp is the equivalent impedance of the
parallel circuit with Np2 Rs and ωL0 components. . . . . . . . . . . . . 28
3.3 Equivalent series circuit of coil and test sample. . . . . . . . . . . . . 29
3.4 Impedance (reactance vs. resistance) diagram with effects of lift-off,
test-object resistivity (ρ) change and test-object equivalent resistance
(Rs ). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.5 δ-function coil between conductor plates k and k 0 . Conductor plates
are enumerated n = 1, 2, . . . , k − 1, k, 10 , 20 , . . . , k 0 − 1, k 0 . . . . . . . . 34
3.6 Coaxial transmitter and detector coils above a series of conductors. . 37
3.7 PCA transformation, reducing the dimensionality of a data set from 3
to 2 (modified from [2]). . . . . . . . . . . . . . . . . . . . . . . . . . 38

5.1 Configuration of a CANDU Fuel Channel. . . . . . . . . . . . . . . . 46


5.2 Eddy current (EC) probe riding over exaggerated pressure tube pro-
trusion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
5.3 16kHz gx revealing fuel bundle mechanical wear marks with nominal
depth 50µm. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
5.4 16kHz C-Scan, plotting eddy current data orthogonal to direction of
lift-off vector, ˆl. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
5.5 16kHz C-Scan, plotting eddy current data in direction of lift-off vector, ˆl. 54
5.6 16kHz gy revealing locations of local pressure tube constrictions (hori-
zontal stripes extending from 0◦ to 360◦ ). . . . . . . . . . . . . . . . . 55
5.7 8kHz gy identifying the position of a spacer. . . . . . . . . . . . . . . 56

xi
5.8 8kHz gy identifying the positions of three constrictions and one spacer. 60
5.9 a) gy computed from 8kHz eddy current data, averaged at the bottom of
the pressure tube. b) dP T −ID computed from ultrasonic measurement
data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

6.1 Configuration of pressure tube to calandria tube gap measurement. . 68


6.2 Geometries in close proximity to the gap probe. Pressure tube and
Calandria tube modelled as flat plates. . . . . . . . . . . . . . . . . . 68
6.3 Composite pressure tube and associated test rig assembly. . . . . . . 72
6.4 Impedance plane display of measured response to change in gap of 0.6
to 16.3 mm is shown by circles ( ), while scaled modeled response to
change in gap of 0 to 16 mm is shown by squares (). . . . . . . . . . 75
6.5 Effect of wall thickness variation on modeled versus measured eddy
current amplitudes for a) 4kHz, b) 8kHz, and c) 16kHz frequencies. . 77
6.6 Amplitude as a function of resistivity for (a) 4kHz, (b) 8kHz, (c) 16kHz
frequencies for scaled model and experimental data from composite tubes. 82
6.7 Variation of scaled model based on gap probe response amplitude (0.5
to 16 mm gap) with resistivity at 55 µΩ · cm plotted as a function of
frequency. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

7.1 Material components in the vicinity of the gap probe, where pressure
tube and calandria tube are approximated as flat plates. . . . . . . . 91
7.2 PCA Scores 1, 2, and 3, plotted together. Analytically modelled data
is clustered in five groups, corresponding to the different modelled re-
sistivity values as labeled. As indicated in Table 7.1, each cluster has
a 0 to 16 mm gap change for three different wall thicknesses. . . . . . 96
xii
7.3 Score 1 peak-to-peak variation (due 0 to 16 mm change in gap) for
model data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
7.4 Data Acquisition Flowchart. . . . . . . . . . . . . . . . . . . . . . . . 99
7.5 Reconstructed 16 kHz ET data on Lissajous plot. . . . . . . . . . . . 101
7.6 Component 1 score with regression surfaces, for variations in PT wall
thickness and PT to CT gap. Experimental data is shown in blue while
analytical model data is shown in yellow. . . . . . . . . . . . . . . . . 105
7.7 Component 1 score with regression surfaces, for variations in PT resis-
tivity and PT to CT gap. Experimental data is shown in blue while
analytical model data is shown in yellow. . . . . . . . . . . . . . . . . 106

xiii
List of Acronyms

AC - Alternating current
AECL - Atomic Energy Canada Limited
CANDU - Canada deuterium uranium
CT - Calandria tube
EC - Eddy current
ET - Eddy current testing
ID - Inner diameter
IMS - Inspection and Maintenance Services
Nb - Niobium
NDE - Non-destructive evaluation
Ni - Nickel
OD - Outer diameter
OPG - Ontario Power Generation
PCA - Principal components analysis
PT - Pressure Tube
Zr - Zirconium

xiv
1

Chapter 1

Introduction

1.1 Background

CANDU (CANada Deuterium Uranium) reactors are pressurized heavy water nuclear
reactors, developed in Canada for electricity generation. Fission reactions are gener-
ated in the reactor core, where uranium bundles serve as fuel. Figure 1.1 depicts the
reactor face, where a lattice of fuel channel end-fittings are shown. These end fittings
are removed prior to uranium fuel bundle deposit.
Figure 1.2 depicts the heat transport system of a CANDU reactor. Fuel bundles
are pushed through the reactor core, and provide the energy source necessary to
keep heavy water surrounding the fuel at elevated temperatures (greater than 250◦
Celsius).
Fuel channels are a critical component of CANDU power plants. They carry the
fuel bundles as they are transported across the core of the reactor. Positioning of
fuel channels relative to other reactor assembly components is shown in Figure 1.3,
while fuel channel components are shown in Figure 1.4. Each fuel channel is made
up of a six metre long pressure tube (which directly houses twelve 0.5 m long fuel
1.1. BACKGROUND 2

Figure 1.1: CANDU reactor face with array of fuel channel end-fittings [1].

bundles) and a surrounding calandria tube. The pressure tube and calandria tube
are separated by four annulus garter spring spacers, which are manufactured from
a square cross-section (either Zr-alloy or Ni-alloy) wire formed into a tight helix.
Further details of these fuel channel components are shown as a magnification in
Figure 1.4, as well as Figure 1.5.
CANDU reactors contain upwards of 400 fuel channels, enclosed within a calan-
dria, a tank containing heavy water neutron moderator (see Figure 1.3). Each fuel
channel consists of a Zr 2.5% Nb pressure tube (PT) with 104 mm inner diameter (ID)
lying within a larger 129 mm ID Zircaloy-2 calandria tube (CT). Nominal resistivities
1.1. BACKGROUND 3

Figure 1.2: Heat transport system with primary components shown [1].

for pressure tubes and calandria tubes are 52 µΩ · cm and 72 µΩ · cm, respectively.
In addition to the fuel bundles, the pressure tube contains heavy water used for
heat transport. Between the PT and the CT is a gas annulus that insulates the hot
(greater than 250◦ Celsius) pressure tube from the colder (50 ◦ C) moderator cooled
CT. Pressure, heat and irradiation induced creep produce diametral creep (increase of
pressure tube diameter and simultaneous reduction of pressure tube wall thickness)
of the PT ID to a maximum of 111 mm and a gradual change in the nominal 4.2
mm wall thickness to an allowable minimum of 3.7 mm [3]. This results in a PT to
CT gap that starts with an average of 8.3 mm, which then decreases with increasing
reactor life. Contact between pressure tube and calandria tube is initially avoided by
1.1. BACKGROUND 4

Figure 1.3: CANDU reactor assembly with heat transport sub-assemblies [1].

4 garter spring spacers that have either a 5.7 mm (older reactors) or 4.8 mm thick-
ness. However sag of the hotter PT within the cooler CT introduces the potential
for contact between PT and CT. Contact for a prolonged period of time introduces
1.1. BACKGROUND 5

Figure 1.4: CANDU reactor assembly with fuel channel components magnified [1].

a risk of hydride blister formation on the PT outer diameter with a consequent risk
of PT cracking [3]. Therefore, PT to CT contact is to be avoided, and monitoring of
PT to CT gap is required [4]. Eddy current testing is employed in monitoring PT to
CT gap [3].
Eddy current testing of conductive materials employs electromagnetic induction
for identification of material properties, including material classification, detection of
flaws, and characterization of defects [5]. Eddy current testing is typically realized
by voltage excitation of a coil, resulting in a time-varying magnetic field that is
produced around the coil. This coil can act as an eddy current test probe [5]. If the
1.1. BACKGROUND 6

Figure 1.5: Fuel channel components - garter spring spacer separates PT and CT [1].

coil is in the proximity of a conductive material, eddy currents will be induced in


the conductive material via Faraday’s law, due to the presence of the time-varying
magnetic field (see Figure 1.6). The eddy currents induced in the test material in
turn produce a magnetic field, which opposes changes in the primary magnetic field
induced into the test material, according to Lenz’s law [6]. The presence of flaws or
material changes in the proximity to the probe modifies the induced eddy currents,
that may be sensed as a change in the coil’s impedance. In general, any modification
of conducting structure geometry (e.g. flaws, variation of material thickness, edges)
in proximity of the probe will modify measured probe impedance. Eddy current
testing is a comparative technique, where measurements on a test piece are related
to measurements on a known configuration. [5]. Capacitive effects are negligible,
relative to large inductance, at the low frequencies used in this work for application
of eddy current testing for PT to CT gap measurement [5].
Since variation of PT to CT gap may affect measured probe impedance, eddy
current testing can also be used to measure PT to CT gap, from within the PT. The
separation of PT and CT is monitored by using an eddy current based technique that
1.1. BACKGROUND 7

Figure 1.6: Principle of eddy current testing. Primary magnetic field induces eddy
currents in conductive surface secondary magnetic field that opposes pri-
mary field.

measures the gap between PT and CT [3]. An example of the parameters of interest
and the basic measurement configuration are shown in Figure 1.7. The technique
compensates for both wall thickness and PT inner diameter variations using normal
beam ultrasonic measurements. Since eddy current response to change in PT to
CT gap is highly sensitive to PT wall thickness, calibrated eddy current responses,
including changes in PT-CT gap and PT wall thickness are employed. The objective
of the inspection system is to measure the PT-CT gap with sufficient accuracy so that
time-to-contact between PT and CT can first be predicted and second, be avoided
[3].
Figure 1.8 depicts the general format for generating the estimate of gap. Inputs
are the calibrated multi-frequency eddy current responses that account for PT wall
1.2. NEED FOR COMPREHENSIVE CHARACTERIZATION OF
MEASUREMENT DATA GATHERED BY GAP PROBE 8

Figure 1.7: PT to CT gap measurement configuration showing minimum and maxi-


mum gap locations and eddy current (EC) probe that traverses the cir-
cumference of the PT ID.

thickness, PT diameter and wall thickness as measured by gauging systems. Probe


response to PT wall thickness and PT-CT gap variations is established by a calibration
facility, where these parameters are varied over the range expected in the reactor.
The algorithm identified in Figure 1.8 assumes constant resistivity for PT and CT,
constant CT wall thickness and diameter and a circular PT. All of these conditions
are present in the calibration facility.

1.2 Need for Comprehensive Characterization of Measurement Data Gath-


ered by Gap Probe

In this section, motivation for a comprehensive characterization of measurement data


gathered by the gap probe is introduced. Further in-depth motivation for specific top-
ics researched in this thesis is presented in the introductions of individual manuscripts
1.2. NEED FOR COMPREHENSIVE CHARACTERIZATION OF
MEASUREMENT DATA GATHERED BY GAP PROBE 9

Figure 1.8: Method for generation of estimated pressure tube to calandria tube gap.

(Chapters 5, 6, and 7).


Multi-frequency eddy current data is acquired by the gap measurement system
with helical motion of the probe within the PT. Eddy current data is gathered every
degree circumferentially and every 1 mm axially, on the surface of a 6 m long pressure
tube. The acquired data is currently only used in the measurement of pressure tube
to calandria tube gap. Given the high sensitivity of eddy current to lift-off, the
high data density employed in acquisition, and the presence of artefacts that can be
characterized as depressions or protrusions (local changes in PT diameter) on the
pressure tube surface, it is natural to assume that more inspection information can
be extracted by developing a new method for analysis of the acquired eddy current
data (as will be discussed in Section 1.2.1).
The employment of PT to CT gap measurement data towards identification of
1.2. NEED FOR COMPREHENSIVE CHARACTERIZATION OF
MEASUREMENT DATA GATHERED BY GAP PROBE 10

pressure tube artefacts, as well as identification of inherent redundancies in acquired


data will serve to provide for a comprehensive characterization of relevant inspection
data gathered by the gap probe. The following subsections will provide further detail.
Additionally, the use and benefits of an analytical model for determination of effects
of PT wall thickness, and PT to CT gap change on independent frequency responses
will also be discussed (Section 1.2.3).

1.2.1 Presence of Pressure Tube Artefacts and Relationship to Gap Probe


Lift-off

In this section, the mechanics of the gap probe will be introduced, and the relationship
of acquired eddy current signals to pressure tube artefacts (described by depressions
and protrusions on the pressure tube surface) will be discussed.
Although the specifics of the probe geometry and design are proprietary, general
aspects of the probe structure may be considered. The gap probe is secured on a
fuel channel inspection head. Given an inspection head installed in the fuel channel,
transmit and receive coils are aligned in the axial direction. The area of the eddy
current probe body is approximately 50 mm x 25 mm. A spring loads the probe
against the pressure tube surface, such that it maintains contact and its area rides
the pressure tube inner diameter surface. The probe is allowed two degree-of-freedom
motion. The first degree of freedom is away from the pressure tube surface - the
direction of spring force. The second degree of motion is about the inspection head’s
rotary direction of motion. This can be described as forward and backward probe
tilt. See Figure 1.9 for an illustration of gap probe directions of motion.
Interactions with pressure tube artefacts [7, 8, 9], which can be characterized as
1.2. NEED FOR COMPREHENSIVE CHARACTERIZATION OF
MEASUREMENT DATA GATHERED BY GAP PROBE 11

Figure 1.9: Eddy current (EC) probe riding over pressure tube surface protrusion
with probe degrees of motion indicated.

protrusions on the pressure tube surface, will induce lift-off of the gap probe, as
shown in Figure 1.9. The presence of lift-off, can therefore, be used as a marker
of pressure tube artefacts, since a lift-off response will be acquired by the gap mea-
surement system. The identification of coincidence of acquisition of lift-off signals
with the locations of pressure tube artefacts is a prime goal for the application of
a signal processing algorithm to extract the location and features of pressure tube
artefacts. Employment of raw eddy current data gathered by the gap measurement
system, information on probe frequency lift-off response (which is near constant for
a given PT/CT configuration) in addition to locations of acquired eddy current ac-
quisition data can all be used as inputs into an algorithm characterizing pressure
tube artefacts. Figure 1.10 gives the direction of impedance plane probe lift-off re-
sponse. Impedance plane representation of eddy current measurements, including
1.2. NEED FOR COMPREHENSIVE CHARACTERIZATION OF
MEASUREMENT DATA GATHERED BY GAP PROBE 12

lift-off response is discussed in Section 3.2.

Figure 1.10: Impedance plane response of probe lift off from pressure tube surface.
Direction of ˆl gives direction of impedance plane probe lift-off response
starting from contact.

1.2.2 Identifying Redundancies in Multifrequency Eddy Current Data


p
The skin depth relation [5, 10] (δ = 50 ρ/µr f , Equation 3.18, derived in Section 3.3)
gives the depth (δ) in mm at which eddy current density is decreased by 1/e, where
f represents frequency in Hz, ρ represents resistivity in µΩ · cm, and µr represents
relative permeability. It is clear from the skin depth relation that δ is dependent on
coil excitation frequency. Increased eddy current penetration occurs with a decrease
in excitation frequency. Conversely, an increase in coil excitation frequency will result
in increased sensitivity to probe lift-off and defects [5] due to decreased skin depth.
1.2. NEED FOR COMPREHENSIVE CHARACTERIZATION OF
MEASUREMENT DATA GATHERED BY GAP PROBE 13

Given ZrNb pressure tube resistivity of 52 µΩ · cm, for 4 kHz, 8 kHz, and 16 kHz
frequencies, the skin depth is 5.7 mm, 4.1 mm, and 2.9 mm, respectively. Despite
differences in penetration depth and sensitivity, there is nonetheless a significant
intersection of useful penetration depth data and responses to lift-off and defects, for
different inspection frequencies (4 kHz, 8 kHz and 16 kHz) employed for PT to CT
gap inspection at Ontario Power Generation.
With respect to employment of multi-frequency eddy current data for NDE appli-
cations, a characterization of the unique (or non-redundant) data gathered by using
multiple frequencies is not presently available in the literature. As multi-frequency
eddy current is used for estimation of PT to CT gap, principal components analysis
(PCA) was investigated to assess what, if any, independent information is gathered
by the gap probe due to changes in critical parameters: pressure tube wall thickness,
and pressure tube to calandria tube gap. This will form an essential component of de-
veloping a comprehensive understanding of relevant inspection information gathered
by the gap probe, as redundancies in information gathered by multifrequency eddy
current measurements can be identified and excluded from a comprehensive model.
PCA involves employing orthogonal transformations to convert a set of observa-
tions, which are correlated, into values belonging to linearly uncorrelated (orthogonal)
variables, called principal components. The main advantage of PCA is that complex
data sets can be reduced to data sets of lower dimension to reveal unapparent and/or
simplified trends in the data [11, 12].
Multi-frequency responses to changes in gap, PT wall thickness, and PT resistiv-
ity can be represented in PCA space, where a smaller number of variables (compared
to the dimension of the original multi-frequency dataset) can be used to investigate
1.2. NEED FOR COMPREHENSIVE CHARACTERIZATION OF
MEASUREMENT DATA GATHERED BY GAP PROBE 14

relationships between the identified physical variables of interest. Principal compo-


nent’s scores represent signal responses to changes in these variables. Often, a small
number of principal components (which are a linear combination of vectors, whose
entries are the original measurement data series) can be used to describe variation
in data sets when these data sets contain correlations [13]. Since multi-frequency
eddy current responses to PT to CT gap, PT wall thickness and PT resistivity are
correlated, a small number of PCA components is expected to represent the original
multi-frequency responses to variation in these parameters.

1.2.3 Employment of an Analytical Model of Responses to Factors Af-


fecting PT to CT Gap Measurement

In the previous section, the importance of identifying independence and redundancies


in multifrequency gap probe eddy current data was highlighted, especially towards
developing a comprehensive characterization of measurement data gathered by the
gap probe. Experimental data is not ideally employed in a PCA approach to repre-
senting multifrequency gap data, as a number of complications would be introduced,
compromising the efficacy of the PCA process including:

• variable signal-to-noise ratio across the frequencies employed for inspection;

• limited control of test piece geometry, due to limited machining tolerances,


resulting in multiple simultaneous parameter variation;

• variable execution of experiments, including effects of probe tilt/and possible


lift-off; and

• effects of instrumentation and setup, which can vary across different instruments
1.2. NEED FOR COMPREHENSIVE CHARACTERIZATION OF
MEASUREMENT DATA GATHERED BY GAP PROBE 15

(e.g. Olympus MS5800 [14], Zetec TC7700 [15]) and experimental setup (e.g.
cable length) employed.

The limitations to employing only experimental data as input into PCA on gap
eddy current data, suggest potential application of an analytical model, which ac-
curately emulates probe response to experimental conditions. Fortunately, variables
which affect gap probe response such as PT wall thickness, PT to CT gap, and PT
resistivity can be modelled via Dodd and Deeds solutions [16, 17, 18], by approx-
imating the pressure tube and calandria tube, in the vicinity of the gap probe, as
parallel plate conductors (introduced in Section 3.4). Data from an analytical model
can be employed towards a PCA analysis, without the complications and limitations
imposed by experimental data. Such an examination will facilitate the development
of methods for extracting parameters of interest from real data.
Physical measurements involving changes in PT wall thickness, PT resistivity and
PT to CT gap can be used to validate PCA computations made using data sets
derived from the analytical model. This will involve using the linear transformations
derived from the PCA computation, on both the analytical model and physical data,
and comparing the transformed data of both data sets.
Finally, using an analytical model (and validating its approach) has benefit to the
inspection qualification process, which is integral to the nuclear industry’s efforts to
qualify its inspection procedures to the regulator in Canada [19]. Showing that it is
possible to analytically describe the effect of variation of an essential parameter (a
parameter whose change in value can affect an inspection procedure’s ability to meet
its intended objectives [20]) can facilitate determination of an inspection system’s
performance as compared to its intended requirements. This can reduce the reliance
1.3. OBJECTIVE 16

on onerous testing in characterizing an inspection system’s capabilities.

1.3 Objective

The objective of this thesis is to develop a comprehensive characterization of eddy


current gap data in order to extract further information on fuel channel condition
and to identify generalized applications for multi-frequency eddy current data.
Multi-frequency eddy current data will be analyzed for changes in the direction
of the probe lift-off impedance response adjacent to the location of the probe. This
will reveal a powerful and generalizable surface profiling technique; identifying the
location of pressure tube flaws, fuel bundle wear marks, garter spring spacers, as well
as other fuel channel features. In contrast to an analysis of localized changes in eddy
current data, PCA will be employed to identify redundancies (and independence) in
multi-frequency eddy current data sets.
The effectiveness of PCA analysis of multi-frequency eddy current data, will be
realized via a complete description of the effects of parameters affecting gap data,
including PT to CT gap, PT resistivity and PT wall thickness in a low number of
dimensions, compared to the dimension of the original data set. Additionally, a gen-
eralizable methodology for compressed data acquisition is realized. Dodd and Deeds
solutions [16, 17, 18] will be employed to model the effects of variation in physical
parameters on gap eddy current data, and will be validated against experimental
data.

1.4 Thesis Outline

In this section an overview of remaining chapters is provided.


1.4. THESIS OUTLINE 17

In Chapter 2, an introductory literature review is provided for material presented


in subsequent chapters. Chapter 3 presents theoretical background related to work
performed towards this thesis. Topics presented in this chapter provide fundamental
background on electromagnetics related to eddy current inspection, as well as Dodd
and Deeds solutions.
An introduction to manuscripts presented in this thesis is given in Chapter 4.
Manuscripts containing the work done towards this thesis are presented in Chap-
ters 5, 6, and 7. Chapter 5 presents a surface profiling technique used to detect PT
artefacts, which is generalizable to variable probe and test-piece configurations. Vali-
dation of Dodd and Deeds solutions for modelling eddy current responses to variable
PT to CT gap, PT wall thickness, and PT resistivity is provided in Chapter 6. Appli-
cations of PCA of validated model data are presented in Chapter 7. Each manuscript
chapter contains an topic-specific introduction, literature review, results, discussion,
and conclusion.
Summarizing discussions and conclusions are provided in Chapters 8 and 9, re-
spectively. Finally, proposed topics for future work are laid out in Chapter 10.
18

Chapter 2

Literature Review

In this chapter a review of relevant literature related to developing a comprehensive


model of inspection information gathered by the gap probe is laid out. The review of
literature should be considered introductory, as more in-depth reviews are conducted
for specific topics presented in attached manuscripts (Chapters 5, 6, and 7).

2.1 Presence of Pressure Tube Artefacts and Relationship to Gap Probe


Lift-off

First, available works that identify the topology of pressure tube artefacts are dis-
cussed. Localized loading of a pressure tube in the vicinity of spacers can produce
deformation on the pressure tube interior surface at these locations [7, 8, 9] (see
Figure 1.4). Mechanical wear marks are caused by fuel bundles dragging along the
bottom surface of pressure tubes as they are moved along the fuel channel during
refuelling operations. These are typically axially oriented and uniform in depth [7].
Pressure tube diametral creep refers to the increase of pressure tube diameter and
simultaneous reduction of pressure tube wall thickness due to prolonged exposure to
radiation and heat. Pressure tube constrictions located near fuel bundle ends result
2.1. PRESENCE OF PRESSURE TUBE ARTEFACTS AND
RELATIONSHIP TO GAP PROBE LIFT-OFF 19

from relatively less diametral creep (characterized by less PT inner diameter increase
resulting from increased radiation and heat) compared to areas close to fuel bundle
centres. As these pressure tube constrictions have relatively less inner diameter [9],
these locations have a protruded pressure tube inner diameter surface. As shown in
Figure 1.9, the identified artefacts interact with the gap probe so as to cause the gap
probe to tilt about the rotary direction of motion (about the pin) and to retract back
(despite the opposing spring force) towards the inspection head. Surface contact at
two points on the probe body is still maintained but the surface is curved at these
locations resulting in variation of probe lift-off.
One characteristic of acquired gap data is high spatial data density. There have
been many studies, which have employed this feature to extract signals of interest.
The complexity of these techniques has varied. Data can be largely unfiltered [21, 22],
or highly conditioned, employing image and signal processing techniques [23, 24,
25, 26]. A common theme in these works, for highlighting artefacts of interest, is
examination of the data with special consideration of the corresponding physical
phenomena resulting in artefact data markers. Interrogation of gap data should
therefore, involve a search for effects of lift-off, which can mark the presence of pressure
tube artefacts.
There are currently two standard ways eddy current data is examined. The first
is through the Lissajous curve, which identifies resistance and inductive reactance
of acquired eddy current data on an oscilloscope display [5]. The second is through
C-Scan examination. C-Scan data is typically represented by displaying the voltage
component orthogonal to lift-off of 2D eddy current voltage data, at locations for
which eddy current data was collected [27]. An example C-Scan is given in Figure
2.2. IDENTIFYING REDUNDANCIES IN MULTIFREQUENCY
EDDY CURRENT DATA 20

5.4.
Since a desired objective of this work is to correlate acquired eddy current signals
with lift-off indications, it is necessary to identify standard methods for correlating
signals. A common method for examining the correlation between two signals is cross-
correlation [28], while convolution of two signals is another technique [29]. However,
a changing two-dimensional signal (e.g. eddy current represented on a Lissajous plot)
can be evaluated in the direction of lift-off vector using a variant of the directional
derivative (defined in Section 3.1); the dot product is defined for complex valued
signals and can be used to compute the directional derivative for complex valued
signals [30].
In extracting features of interest (e.g. artefact locations), employment of image
processing tools can be utilized. One such tool is the Canny edge detector [31], which
relies on computation of the partial derivative of source data, which has also been
employed by Gonzalez et al. [32]. The partial derivative technique will be examined
as a step in identifying characteristics of probe lift-off, relative to the pressure tube
inner diameter. This is performed in the first manuscript provided in this thesis
(Chapter 5)

2.2 Identifying Redundancies in Multifrequency Eddy Current Data

Principal components analysis has been employed in a number of applications to


identify independence and redundancies in multidimensional data sets.
This technique was first developed by Karl Pearson [33] and later developed in-
dependently by Hotelling [34]. It has wide application, including image processing,
2.2. IDENTIFYING REDUNDANCIES IN MULTIFREQUENCY
EDDY CURRENT DATA 21

particularly facial recognition and satellite image processing algorithms [35, 36]. Prin-
cipal components analysis has been employed on conventional eddy current testing,
to improve reliability of interpretation of eddy current signals for steam generator
tubes [37].
Particular to pulsed eddy current testing (a class of eddy current testing which
examines the transient responses to short voltage spikes applied to eddy current
coils) principal components analysis has been recently employed toward a number
of applications. PCA applied to pulsed eddy current was first shown to provide
enhanced classification of defects [38]. Further applications of PCA to pulsed eddy
current include detection of defects in multilayer aluminum lap joints [39, 40, 41],
steel [42], and aircraft structures [43, 44, 45]. Additionally, PCA was employed for
detection of cracks in multiple layers using pulsed eddy current and giant magneto
resistive sensors [46, 47]. A modified PCA technique was used to detect cracks in an
F/A-18 inner wing spar without wing skin removal [13, 48, 49]. Applied to pulsed
eddy current, PCA has also been combined with wavelet analysis [50] and smooth
nonnegative matrix factorization [51].
Typical employment of principal components analysis for pulsed eddy current ap-
plications considers principal component eigenvectors plotted as a function of time,
in classification of defects [49, 40, 50, 38]. It should be noted however, that conven-
tional eddy current data provides measurement results, while the magnetic circuit is
in steady state, and therefore, conventional eddy current measurement can be con-
sidered time invariant. Principal components analysis can therefore, be employed on
conventional eddy current data, removing the time dependence from interpretation of
results. Also, for conventional eddy current data, principal components analysis can
2.3. EMPLOYMENT OF AN ANALYTICAL MODEL OF EDDY
CURRENT RESPONSES TO FACTORS AFFECTING PT TO CT
GAP MEASUREMENT 22

simplify the data analysis, by relating the reduced score representation of the data to
parameters that produce variations in the signal.
Principal components analysis has not yet been applied to multifrequency eddy
current applications (e.g. [52, 24]), employing conventional eddy current testing. As
will be shown, principal components analysis will be employed to identify indepen-
dence and redundancies in multifrequency eddy current gap data.

2.3 Employment of an Analytical Model of Eddy Current Responses to


Factors Affecting PT to CT Gap Measurement

Analytic solutions to Maxwell’s equations, which model eddy current probe coils
have been developed by Dodd et al. [16, 17, 18]. Employment of these solutions
in determining independence (and redundancies) in gap probe eddy current data is
especially useful in that these models do not exhibit variable probe frequency-gain
and noise, compared to physical systems, and as such, performing PCA with model
data will provide solutions that are repeatable and free of data corruption.
In order to employ solutions to probe coils using Dodd and Deeds analytical
models, these models have to be validated. Examples of validation of analytical
models of eddy current testing methods are available [53, 54]. An important element
of the validation process is defining physical limitations for which the validation is
correct. This should also be considered in validating Dodd and Deeds solutions [16,
17, 18] for modeling pressure tube to calandria tube gap data.
Using an analytical model for predicting effects of physical parameters such as
pressure tube resistivity, pressure tube wall thickness, and pressure tube to calandria
2.3. EMPLOYMENT OF AN ANALYTICAL MODEL OF EDDY
CURRENT RESPONSES TO FACTORS AFFECTING PT TO CT
GAP MEASUREMENT 23

tube gap has positive side effects with respect to the process of inspection qualifi-
cation. Inspection qualification, which is a nuclear operator regulatory requirement,
[19], serves to determine an inspection system’s capabilities, as compared to its in-
spection specification requirements. Physical parameters whose change in value could
affect an inspection system, such that its specified objectives would no longer be met,
are termed essential parameters [55] and are central in the process of inspection qual-
ification. The effectiveness of modeling has been highlighted for characterizing effects
of essential parameter variation, in the context of inspection qualification [20].
24

Chapter 3

Theory

In this chapter a brief theoretical introduction to tools employed towards developing


a comprehensive model of inspection information gathered by the gap probe is laid
out. A full review of the specific theory employed for each paper is available via
review of the related theory sections in attached manuscripts (Chapters 5, 6, and 7).
Eddy current testing is employed for inspection of electrically conductive materi-
als. If an alternating current is used to drive a coil, a time-varying magnetic field will
be generated in and around the coil [5]. According to Faraday’s law, eddy currents will
be induced in the conductive material that is in the proximity of the time-varying
field. Lenz’s law specifies that induced eddy currents will have directions that in-
duce magnetic fields opposing the change in magnetic fields induced by the drive coil
current. Eddy current testing monitors coil impedance and is thereby, sensitive to
parameters that modify coupling (mutual inductance) between probe (made up of a
single combination drive/detector coil, or seperate drive/detector coils) and sample.
These parameters include the presence of flaws, and probe lift-off, as these will have
a direct impact on the impedance measured by an eddy current testing system.
3.1. MATHEMATICAL GRADIENT AND DIRECTIONAL
DERIVATIVE 25

3.1 Mathematical Gradient and Directional Derivative

The concepts presented in this section are further developed in the analysis of eddy
current signals in one of the papers (Manuscript I, Chapter 5) provided in this thesis
and are therefore, introduced only briefly. As laid out in Section 5.4, the directional
derivative is used to compute the projections of change in eddy current signals, in the
direction of the lift-off vector. The axial and rotary components of the directional
derivative are considered separately, in imaging pressure tube artefacts (Section 5.5).
The mathematical gradient is a well defined mathematical operator employed in
many analytical fields of study. The gradient of a function of three variables, f (x, y, z)
where î is the unit vector in the x direction, ĵ is the unit vector in the y direction,
and k̂ is the unit vector in the z direction is defined as [56]

∂f ∂f ∂f
∇f = î + ĵ + k̂. (3.1)
∂x ∂y ∂z

The gradient of a function is a vector, and represents the slope of the tangent of
a function. The direction of the gradient is in the greatest rate of increase of the
function, with respect to variables x, y and z.
The directional derivative of f along a given unit vector û = (ux , uy , uz ) at a point
(x0 , y0 , z0 ) represents the rate of change of f (x0 , y0 , z0 ) with respect to variables x, y, z,
along the direction of û. The directional derivative in the direction û is computed as

Dû f (x0 , y0 , z0 ) = ∇f (x0 , y0 , z0 ) · û (3.2)


∂f (x0 , y0 , z0 ) ∂f (x0 , y0 , z0 ) ∂f (x0 , y0 , z0 )
= ux + uy + uz . (3.3)
∂x ∂y ∂z
3.2. EQUIVALENT CIRCUIT AND IMPEDANCE PLANE
REPRESENTATION 26

3.2 Equivalent Circuit and Impedance Plane Representation

In this section a derivation is provided for an impedance plane representation of


an eddy current coil, in proximity to a conductive test sample. The impedance
plane representation is useful, in that the changing coil impedance relates to the
sensitivity to liftoff (both PT and CT), resistivity and wall thickness. It can also be
used to explain effects due to variation in temperature, which causes changes in coil
impedance (resistance) and sample resistivity. All these factors facilitate a qualitative
interpretation of the gap probe response under various in-reactor conditions.
In the proximity of an electrically conductive test material, a coil used in eddy
current testing can be modeled as a primary winding of a transformer [5, 57]. The
eddy currents induced in the test material, conversely, can be modeled as a single
turn of a secondary winding (which sets Ns , the number of turns in the secondary
winding to 1). In the ‘send-receive’ method of eddy current testing, another coil is
used to monitor variations in received voltage (Figure 3.1). The following analysis in
this section does not consider electromagnetic interactions between the receive coil
and either the transmit coil or conductive test object.
Equivalent circuits can be used to simplify the electromagnetic relationship be-
tween a coil and test sample, and to obtain an impedance diagram, which is central
to the analysis of eddy current measurement data (see Figures 3.2 and 3.3). Note
that this treatment neglects skin effects in the conductor as described by Equation
3.17.
The equivalent parallel circuit has equivalent impedance [5]

Z1 Z2
Zp = (3.4)
Z1 + Z2
3.2. EQUIVALENT CIRCUIT AND IMPEDANCE PLANE
REPRESENTATION 27

Figure 3.1: Model of a transmit and receive coil with a test object.

where Z1 = Np2 Rs and Z2 = jX0 . Np is the number of turns in the transmit coil,
Rs is the variable resistance of the conductive test object, and j is the imaginary
number. Where L0 is the probe inductance, X0 = ωL0 is the inductive reactance of
the coil in proximity to the test object. As Rp is a constant which is not affected by
test object impedance, Rp can be excluded from Zp in deriving an impedance plane
representation of an EC coil in proximity to a test object. Zp can be evaluated as

jNp2 Rs X0
Zp = 2 . (3.5)
Np Rs + jX0

Rationalizing the denominator yields [5, 57]

Np2 Rs X02 (Np2 Rs )2 X0


Zp = + j , (3.6)
(Np2 Rs )2 + (X0 )2 (Np2 Rs )2 + (X0 )2
3.2. EQUIVALENT CIRCUIT AND IMPEDANCE PLANE
REPRESENTATION 28

Figure 3.2: Equivalent parallel circuit of coil and test sample represented without and
with equivalent impedance. Zp is the equivalent impedance of the parallel
circuit with Np2 Rs and ωL0 components.

where RL can be used to represent the real part of the equation, and Xp can be used
to represent the imaginary part of the equation, such that [5, 57]

Zp = RL + jXp . (3.7)

Test samples have corresponding resistivities, which are represented as a parallel


resistance to the test coil inductive reactance, and can be affected by the presence of
defects [5]. Changes in test sample resistivity are reflected as changes to resistance and
reactance changes experienced by the test coil. The operating point in an impedance
3.2. EQUIVALENT CIRCUIT AND IMPEDANCE PLANE
REPRESENTATION 29

Figure 3.3: Equivalent series circuit of coil and test sample.

plane representation of impedance experienced by the coil is moved up the impedance


curve (Figure 3.4 [5]. When mutual coupling between the coil and sample is decreased
(i.e. the coil experiences lift off from the test sample), smaller semi-circles are traced
on the impedance plane [5].

Figure 3.4: Impedance (reactance vs. resistance) diagram with effects of lift-off, test-
object resistivity (ρ) change and test-object equivalent resistance (Rs ).
3.3. MAXWELL’S EQUATIONS AND THE SKIN DEPTH
EQUATION 30

3.3 Maxwell’s Equations and the Skin Depth Equation

In this section a derivation of the skin depth equation is provided, beginning with
Maxwell’s equations. The skin depth equation is discussed in relation to results
presented in Manuscripts I and II (Chapters 5 and 6).
Where E is the electric field, B is the magnetic flux, ρf is the free electric charge
density, µ is the permeability,  is the permittivity, and σ is the conductivity, in a
linear medium. Maxwell’s equations have the form [6]

ρf
∇·E = (3.8)


∇·B =0 (3.9)

∂B
∇×E =− (3.10)
∂t

∂E
∇ × B = µσE + µ (3.11)
∂t

Taking the curl of Equation 3.10, substituting into Equation 3.11, and using the
identities ∇ × ∇ × E = ∇(∇ · E) − ∇2 E and ∇(∇ · E) = 0, we obtain, with some
manipulation,

∂E ∂ 2E
∇2 E = µσ + µ 2 . (3.12)
∂t ∂t

For good conductors, where σ is very large (σ/ω  1 [10]), and for a time
harmonic excitation, this can be written as the Diffusion Equation (for frequencies
3.3. MAXWELL’S EQUATIONS AND THE SKIN DEPTH
EQUATION 31

less than 108 Hz [16]),

∂E
∇2 E = µσ . (3.13)
∂t

By Ohm’s Law, J = σE, the electric field drives the current

∂J
∇2 J = µσ . (3.14)
∂t

If the current varies sinusoidally in time, in phasor notation, J (t) = Re{J0 e−jωt }
where J0 is the vector having magnitude and direction of the current density, and
substituting into the diffusion equation we have [10]

∇2 J0 + α2 J0 = 0, (3.15)

where α2 = jµσω. Where j is the imaginary number, and since the following mathe-

matical identity holds, j = √12 (1 + j),

r
µσω 1+j
α = (1 + j) = , (3.16)
2 δ

where (in MKS units) [10]

r
2
δ= , (3.17)
σµω

or [5]
r
ρ
δ = 50 , (3.18)
µr f

where δ is the skin depth measured in mm, ρ is the resistivity measured in µΩ · cm,
3.4. DODD AND DEEDS SOLUTIONS 32

µr is the relative permeability, and f is the frequency measured in Hz. Equation 3.18
is the Skin Depth Equation, where δ is the depth at which the eddy current density
has decreased to 1/e, or 36.8% of the surface density. In general, penetration depth
decreases with increased frequency, and increases with increased resistivity as per
Equation 3.18. Eddy currents produced at and below the test surface oppose changes
in the primary magnetic field produced by the alternating current in the probe coil,
according to Lenz’s law [6, 10].

3.4 Dodd and Deeds Solutions

The following is a high-level summary of how the eddy current probe response can be
derived analytically, by approximating the test surface(s) as one or more flat plates.
The most significant points, from Ref. [18], are summarized below.
A coordinate system is employed, where the driving current and media are axially
symmetric, reflecting the coil geometry above a conducting plane (Figure 3.5). Due to
the axial symmetry, the current density J and vector potential A have only azimuthal
components, such that

J (x) = J(r, z)êφ (3.19)

and

A(x) = A(r, z)êφ , (3.20)

where êφ is an azimuthal unit vector. The magnitude of the magnetic vector potential
can be expressed as [18]
3.4. DODD AND DEEDS SOLUTIONS 33

ZZ
A(r, z) = G(r, z; r0 , z 0 )J(r0 , z 0 )dr0 dz 0 , (3.21)

where G(r, z; r0 , z 0 ) is the Green’s function for a δ-function current (having infinite
current density, but total current I) at (r0 , z 0 ), as shown in Figure 3.5. G(r, z; r0 , z 0 )
is also the magnetic vector potential at (r, z) of the δ-function current [16]. A(r, z)
is the summation of the magnetic vector potential, due to a number of δ-coils, which
together approximate real physical coils. The Green’s function satisfies the following
equation, assuming a linear, isotropic, homogeneous medium, having time-harmonic
current angular frequency ω [18],

∂2 ∂2
 
1 ∂ 1
2
+ − 2 + 2 − jωµσ + ω µ G(r, z; r0 , z 0 ) = −µδ(r−r0 )δ(z−z 0 ) (3.22)
2
∂r r ∂r r ∂z

where µ, , and σ are the permeability, permittivity, and conductivity of the medium,
respectively.
To find the vector potential, given in Equation 3.21, solutions of Equation 3.22
that satisfy proper boundary conditions are used. From the vector potential, all other
electromagnetic quantities of interest can be calculated [18].
The solution of Equation 3.22 for each region is given by [18]

Z ∞
0 0
(n)
Bn (α)e−αn z + Cn (α)eαn z J1 (αr)dα
 
G (r, z; r , z ) = (3.23)
0

for n = 1, 2, . . . , k −1, k, 10 , 20 , . . . , k 0 −1, k 0 , where J1 (x) is a Bessel function [58] of the


first kind and the first order and αn2 = α2 +jωµn σn −ω 2 µn n , where j is the imaginary
number. The Green’s function is finite when z approaches positive or negative infinity,
3.4. DODD AND DEEDS SOLUTIONS 34

Figure 3.5: δ-function coil between conductor plates k and k 0 . Conductor plates are
enumerated n = 1, 2, . . . , k − 1, k, 10 , 20 , . . . , k 0 − 1, k 0 .

and therefore B1 = 0 when z → −∞ and C10 = 0 when z → +∞. Other coefficients


Bn , and Cn can be determined through considering boundary conditions at each
conductor interface [18]. The following boundary conditions for Green’s functions are
determined by considering tangential components of the electric field and magnetic
3.4. DODD AND DEEDS SOLUTIONS 35

field strength [18]:

[G(n) (r, z; r0 , z 0 )]z=zn = [G(n+1) (r, z; r0 , z 0 )]z=zn , (3.24)

and

1 ∂G(n) (r, z; r0 , z 0 ) 1 ∂G(n) (r, z; r0 , z 0 )


   
= + δn,k δ(r − r0 ). (3.25)
µn ∂z z=zn µn+1 ∂z z=zn

Where δn,k = 1 when n = k, and 0 otherwise, substitution of Equation 3.23 into


Equations 3.24 and 3.25 and employing the Fourier Bessel integral [16], the following
relations between coefficients of neighbouring conductors are provided [18],

1
Bn+1 = eαn+1 zn [(1 + βn+1,n )e−αn zn Bn + (1 − βn+1,n )eαn zn Cn ] (3.26)
2

and

1
Cn+1 = e−αn+1 zn [(1 − βn+1,n )e−αn zn Bn + (1 + βn+1,n )eαn zn Cn ], (3.27)
2

where βn+1,n = βn /βn+1 and βn = αn /µn .


Exhaustive calculations are developed in Ref. [18], to simplify the calculation of
the vector potential A(r, z) for a number of conductor configurations (e.g. coil above
conductors, coil between three-layer conductors, coil above three-layer conductor), as
well as a number of coil regions (e.g. coil region, region above the coil, region below
the coil.)
3.4. DODD AND DEEDS SOLUTIONS 36

Depicted in Figure 3.6, for the case where a coil is above the conductors (e.g. gap
eddy current probe above PT and CT conductors), and in the coil region (i.e. neither
above nor below the coil), the magnetic vector potential is given as [18]

Z ∞
1 1
A(r, z) = nc I J(r2 , r1 )J1 (αr)e−α0 l1 [(γe−α0 l1 + eα0 z )(1 − e−α0 (z−l1 ) )
2 0 β0 α 0 α
+ (γe−α0 z + eα0 z )(1 − e−α0 (l2 −2) )]dα,

(3.28)

where γ is a constant, which captures the positions and conductivities of the conduc-
tors [18], nc is the turn density, I is the current, and nc I is the current density of a
densely and uniformly wound driving coil. l1 and l2 are positions of the upper and
lower surfaces of the driving coil, respectively, and

Z αr2
J(r2 , r1 ) = xJ1 (x)dx, (3.29)
αr1

where r1 and r2 are the inner and outer radii of the coil, as shown in Figure 3.6.
For each of the conductor configurations, the induced voltage and coil impedance
in a detector (receiver) coil located at (r, z) are computed, where the induced voltage
is given [18]

I
V1 (r, z) = − E · dl = jω2πrA(r, z), (3.30)

as obtained by Faraday’s law V = −dφ/dt (where V is voltage, φ is magnetic flux,


and t is time) and with Stoke’s theorem [6] applied to Equation 3.10.
3.4. DODD AND DEEDS SOLUTIONS 37

For uniformly wound, dense conductors, the total induced voltage is equal to

Z l20 Z r20
V = jω2πnd rA(r, z)drdz, (3.31)
l10 r10

where l10 , l20 , r10 , r20 are the upper and lower surfaces and inner and outer radii, respec-
tively, of the detector coil as shown in Figure 3.6, and nd is its turn density.

Figure 3.6: Coaxial transmitter and detector coils above a series of conductors.

For a detector coil located in the coil region, using Equations 3.28 and 3.31, the
total induced voltage is evaulated [18]
3.5. PRINCIPAL COMPONENTS ANALYSIS 38

Z ∞
1
V =jωIπnc nd 2 3
J(r2 , r1 )J(r20 , r10 )
0 β0 α0 α
0 0 0 0
[2α0 (l20 − l10 ) + (1 − e−α0 (l2 −l1 ) )(−e−α0 l2 (γe−α0 l1 + eα0 l2 )+ (3.32)
0
e−α0 l1 (γe−α0 l1 − eα0 l1 )]dα.

3.5 Principal Components Analysis

Principal components analysis (PCA) involves employing orthogonal transformations


to convert a set of observations, which are correlated, into a set of linearly uncor-
related variables, called principal components [12]. The main advantage of PCA is
that complex data sets can be reduced to data sets of lower dimension to reveal un-
apparent and/or simplified trends in the data [11, 12]. This can be observed through
examination of Figure 3.7, where highly correlated three dimensional data is reduced
to two dimensions. Note that in Figure 3.7, PC1 accounts for the largest variance in
the data set and PC2, which is orthogonal to it, the next largest variance.

Figure 3.7: PCA transformation, reducing the dimensionality of a data set from 3 to
2 (modified from [2]).

PCA will be used to determine uniqueness and redundancy across multi-frequency


3.5. PRINCIPAL COMPONENTS ANALYSIS 39

eddy current data sets, and relate effects of physical parameter changes to principal
component scores, for PT to CT gap data.

3.5.1 Process

The objective of a principal components analysis transformation is to find a linear


combination of the original variables X = [x1 , x2 , . . . , xp ] with maximum variance
[12]. X is first normalized such that each variable xi , i ∈ [1, 2, . . . , p] has zero mean
and unit variance.
A new set of vectors

tk(i) = x(i) · w(k) (3.33)

is defined such that individual variables of t inherit (successively) maximum possible


variance from x, with each loading vector w constrained as a unit vector. The first
loading vector w(1) is given as

wT X T Xw
 
w(1) = arg max , (3.34)
wT w

where arg max is defined as the argument of the maximum, which is the set of points
of the given argument for which the given function attains its maximum value. In
this case w is chosen such that wT X T Xw/wT w attains its maximum value. The kth
component can be found by first subtracting the k − 1 principal components from x:

k−1
X
T
X̂k = X − Xw(s) w(s) . (3.35)
s=1

Now,
3.5. PRINCIPAL COMPONENTS ANALYSIS 40

( )
wT X̂kT X̂k w
ŵ(k) = arg max . (3.36)
wT w

The kth principal component of vector x(i) is given as a score tk(i) = x(i) · w(k) , where
the full decomposition of X is given as

T = XW. (3.37)

Equation 3.37 is used in Manuscript III (Chapter 7) to transform eddy current


responses to factors affecting PT to CT gap measurement responses (represented by
horizontal and vertical impedance plane responses for 4 kHz, and 8 kHz, and 16
kHz frequencies), to principal component scores. Scores belonging to the first and
subsequent principal components contain the majority of variance of the original data
set.
41

Chapter 4

Introduction to Manuscripts

The following journal articles, Manuscripts I, II, and III, are presented as Chapters
5, 6, and 7, respectively.

Manuscript I: S. Shokralla, T. W. Krause, and J. Morelli, “Surface profiling with high


density eddy current non-destructive examination data,” NDT & E International, vol.
62, March 2014, pp. 153-159.

Manuscript II: S. Shokralla, S. Sullivan, J. Morelli, and T. W. Krause, “Modelling


and validation of Eddy current response to changes in factors affecting pressure tube
to calandria tube gap measurement,” NDT & E International, vol. 73, July 2015,
pp. 15-21.

Manuscript III: S. Shokralla, J. E. Morelli and T. W. Krause, “Principal Components


Analysis of Multifrequency Eddy Current Data Used to Measure Pressure Tube to
Calandria Tube Gap,” IEEE Sensors Journal, vol. 16, no. 9, May 2016, pp. 3147-
3154.
42

In Manuscript I, a generalizable surface profiling technique is derived, exploiting


the interaction between the PT to CT gap probe and pressure tube inner diameter
surface. Manuscript II provides experimental validation of Dodd and Deeds solutions,
given expected in-reactor variations in physical parameters: PT wall thickness, PT
resistivity, and PT to CT gap. Using validated Dodd and Deeds solutions, PCA of
modelled changes to parameters affecting PT to CT gap measurement is employed
in Manuscript III, to relate physical parameter changes to unique information across
PT to CT gap multifrequency data. Additionally, in Manuscript III, PCA output is
validated against physical experiments, and a compressed data acquisition algorithm,
reliant on PCA fundamentals is derived and emulated using modelled PT to CT
multifrequency eddy current data.
43

Chapter 5

Surface Profiling with High Density Eddy Current

Non-Destructive Examination Data

Shaddy Shokralla, Thomas Krause, and Jordan Morelli

Abstract: The pressure tubes (PT) in CANDU (CANada Deuterium Uranium) re-
actors undergo creep induced deformation due to operating pressure, temperature
and radiation conditions. While global deformation of the tube in the form of elonga-
tion and diametral creep are well characterized and monitored by station inspection
systems, local PT deformation and the presence of inner surface artifacts due to wear
are not as directly monitored, but can still provide additional information of fuel
channel condition. A surface profiling technique for monitoring local deformation
and identification of surface wear, using an eddy current probe mounted in a small
(50 mm x 25 mm) planar probe body is presented. The sensitivity of the eddy current
probe to small lift-off variations combined with high density C-Scan information is
used to extract information on smoothly varying local deformation as well as monitor
more significant wear on the inner surface of pressure tubes. Vector separation of
components permits independent identification of axial and circumferential surface
5.1. INTRODUCTION 44

features. Analysis of this data can be used to characterize local PT deformation due
to constrictions at fuel bundle ends and loaded garter spring spacers, as well as iden-
tify areas where shallow mechanical wear has occurred. Examples of the features that
may be identified are presented.
Keywords: surface profiling, eddy current, lift-off, signal processing, data density

5.1 Introduction

Eddy current inspection is a key non-destructive examination (NDE) process em-


ployed towards characterization of defects in conductive materials. Piping and heat
exchanger (including steam generator) inspections are common applications of eddy
current non-destructive examination of nuclear power plants [27, 59, 60]. Specific to
CANDU (CANada Deuterium Uranium) reactors however, eddy current examination
has been used towards a number of non-standard fuel channel related applications
with regards to the geometry identified in Figure 5.1. These include identification
of the presence and proximity of remote structures to fuel channels [61], detection
of spacers, which maintain gap between pressure and calandria tubes [9] and direct
measurement of gap between pressure tube and calandria tube [3].
There are two standard ways in which eddy current data is interrogated. The first
is by examination of the Lissajous curve, which provides the resistance and inductive
reactance recorded by an eddy current receive probe during data acquisition. The
second is through examination of the C-Scan, in which the component orthogonal to
the direction of lift-off is plotted as a function of measurement location (points on a
two-dimensional grid).
There are many examples of employing high data density towards eddy current
5.1. INTRODUCTION 45

C-Scan imaging. Imaged data can either be minimally conditioned [21, 22], or al-
ternatively, a significant degree of signal and image processing can be employed
[23, 24, 25, 26].
For flaw detection applications, examination of lift-off in eddy current data is
normally avoided and is considered a source of noise [5, 62]. The application of paint
and metal deposition thickness measurement is well known. But its potential for
surface profiling is not as well recognized [63]. For the particular circumstance where
a large planar probe body is present, lift-off may be utilized to provide additional
information that would be available if the probe maintains surface following or surface
riding characteristics.
This study shows how high eddy current data-density, basic knowledge of mechani-
cal interaction between the probe and test-piece, and lift-off response can be employed
towards profiling a test-piece surface to a high degree of precision. The technique to
be described in detail is simple to implement, relying on a computation of partial
derivatives with respect to variables whose direction represent orthogonal image axes
(an image processing building block [32, 31]). Ease of technique implementation is
in contrast to more elaborate techniques for analysis of eddy current data, including
neural network implementation and wavelet transformation [25, 64, 65, 66].
The technique will be compared against another currently employed ultrasound
based method for detection of spacer and pressure tube artefact locations, important
activities in ensuring integrity of CANDU nuclear reactor fuel channels.
5.2. INSPECTION REQUIREMENT 46

5.2 Inspection Requirement

Fuel channels are a critical component of CANDU power plants. They carry the fuel
bundles as they are transported across the core of the reactor, whose configuration
is shown in the top portion of Figure 5.1. Each fuel channel is made up of a six
metre long pressure tube (which directly houses twelve 0.5m long fuel bundles) and a
surrounding calandria tube. The pressure tube and calandria tube are separated by
four spacers, which are manufactured from a square cross-section Zr-alloy wire formed
into a tight helix. Further details of these Fuel Channel components are shown as a
magnification in Figure 5.1.

Figure 5.1: Configuration of a CANDU Fuel Channel.

Fuel channels are periodically inspected in order to characterize the presence and
status of different degradation mechanisms, thereby meeting their safety and licensing
requirements [4]. This is performed by insertion of an inspection head along the
length of the fuel channel during reactor shutdown periods. A number of different
degradation mechanisms and pressure tube artefacts can be identified by various NDE
5.3. MEASUREMENT TECHNIQUE 47

methods (including eddy current). These include the following.

• Localized deformation due to garter spring spacer loading: Garter spring spac-
ers serve the purpose of ensuring separation between the pressure tube and
relatively cool calandria tube. Loading is therefore induced on the pressure
tube by the spacer and supporting calandria tube. This leaves the pressure
tube slightly ovalized in the vicinity of the spacer.

• Constrictions: Locations at fuel bundle ends where radiation induced creep is


less due to lower flux densities relative to the centre of fuel bundle locations.
This leaves the diameter of the pressure tube at a constriction smaller than
for neighbouring regions, while forming a pressure tube wall protrusion on the
order of 0.1mm.

• Mechanical wear marks: Mechanical interaction between fuel bundles moving


across the fuel channel, and the supporting pressure tube can induce wear on
the pressure tube. Mechanical wear marks indicate this repeated interaction.
The wear marks have nominal depth less than 30 µm [67].

5.3 Measurement Technique

Eddy current data is gathered helically by an eddy current probe mounted on an


inspection head that is driven axially and rotationally along the pressure tube (Figure
5.2). The axial pitch of the probe’s helical travel is small (1mm), compared to its
rotational travel, allowing for approximation of the helical trajectory as a series of
circular trajectories separated axially by 1mm.
5.3. MEASUREMENT TECHNIQUE 48

The eddy current probe is mounted on the inspection head, with transmit and re-
ceive coils situated along the axial direction of the pressure tube, when the inspection
head is installed in-channel. The probe is spring loaded such that it rides the surface
of the pressure tube. In the event that the probe makes contact with an obstruction
embedded in the pressure tube interior, as shown in Figure 5.2, it is also permitted
limited motion with two degrees of freedom: about the inspection head rotary direc-
tion of motion (forward and backward tilt), and away from the pressure tube surface
in the direction of spring force.

Figure 5.2: Eddy current (EC) probe riding over exaggerated pressure tube protru-
sion.

The eddy current signals are generated and received via an Olympus NDT Multi-
scan MS5800 eddy current instrument. Custom software is employed to view acquired
signals in real time as well as post-acquisition. Eddy current signals are acquired on
multiple frequencies, which include 8kHz and 16kHz. Unadjusted raw measurements
have units mV.
5.4. LIFT-OFF LOCATION EXTRACTION 49

5.4 Lift-off Location Extraction

We represent the acquired eddy current data set as a complex valued function f (x, y)
mapping rotational and axial values (x and y) to the complex plane, C. We use C
to represent the domain where eddy current signals lie. Orthogonal components of
C represent the resistance and inductive reactance components in the eddy current
impedance plane. It is worthy to note that axially adjacent values of f (x, y) (along
the y-axis) are collected 360 degrees apart, due to the helical eddy current probe
trajectory; data is collected helically at 1 rotary degree (approximately 0.9mm) per
sample.
A retraction signal is sent to the eddy current probe during its acquisition sequence
and a lift-off signal (which can be represented as a horizontally oriented vector on the
Lissajous curve) is obtained. The direction of the lift-off vector is a key component
used in the calculations and is employed to allow for surface profiling using eddy
current data. ˆl is used to denote the complex valued unit vector lying in C with the
same direction as the direction of the lift-off response.
Since values of f (x, y) lie on the complex plane, C, f (x, y) can be written

f (x, y) = u(x, y) + i v(x, y). (5.1)

Partial derivatives of f (x, y) with respect to x and y are defined as

∂f (x, y) ∂u(x, y) ∂v(x, y)


= +i (5.2)
∂x ∂x ∂x

and
5.4. LIFT-OFF LOCATION EXTRACTION 50

∂f (x, y) ∂u(x, y) ∂v(x, y)


= +i . (5.3)
∂y ∂y ∂y

Where the dot product is defined for complex valued vectors [30], the scalars
gx (x, y) and gy (x, y) are defined as the components of ∂f (x, y)/∂x and ∂f (x, y)/∂y
in the direction of lift off, such that

∂f (x, y) ˆ
gx (x, y) = · l, (5.4)
∂x

and

∂f (x, y) ˆ
gy (x, y) = · l. (5.5)
∂y

gx (x, y) and gy (x, y) represent the change in eddy current data, in the rotary and
axial directions, respectively, in the direction of lift-off.
Examination of gx and gy reveal not only areas in the inspection space where
the eddy current probe has lifted off from the pressure tube, but also reveal key
mechanical interactions between the eddy current probe and the inspection surface.
gx represents rotary lift-off experienced by the probe. That is to say, changes in lift-off
experienced by the probe as it traverses in the rotary direction over pressure tube
obstructions extending along the axial direction of the pressure tube.
Consider Figure 5.2. The eddy current probe is not free to rotate about the axial
direction of travel. So, the physical motion of the probe as it traverses over axially
extending pressure tube obstructions is away from the surface of the pressure tube in
the direction of spring force, and is limited to one degree of freedom.
gy on the other hand, represents changes in axial lift off experienced by the probe
5.5. APPLICATIONS 51

as it traverses in the axial direction over pressure tube obstructions away from the
surface of the pressure tube. The obstructions extend in the rotary direction of the
pressure tube. Unlike the case for lift-off of the probe as it traverses in the rotary
component of travel, (again consider Figure 5.2) the probe is free to move both away
from the pressure tube wall as well as rotate (tilt) about the rotary direction of
travel. Probe motion is therefore capable of two degrees of freedom when lifting off
over rotationally extended pressure tube obstructions.

5.5 Applications

In this section a number of pressure tube artefacts are identified using the computa-
tional method introduced in the previous section.

5.5.1 Mechanical Wear Marks

Calculation of gx for a set of eddy current data obtained in-reactor determines probe
lift-off variation in its rotary component of travel. The direction of probe lift-off is
normal to the pressure tube wall.
Vertical stripes are evident in Figure 5.3, revealing a systematic phenomenon,
which possesses a lift-off component in the rotary direction of travel at the bottom of
the pressure tube (near 180 degrees) for a significant axial length along the pressure
tube (800mm). The vertical stripes represent pressure tube mechanical wear marks
that have been caused by fuel bundles sliding along the bottom of the pressure tube.
The depth of these mechanical wear marks is typically less than 30µm [67]; as such,
computation of gx for the 16kHz frequency has been used to identify features in the
pressure tube surface not evident in standard eddy current Lissajous representations.
5.5. APPLICATIONS 52

Figure 5.3: 16kHz gx revealing fuel bundle mechanical wear marks with nominal depth
50µm.

Figure 5.4 shows a conventional C-Scan for the same eddy current data, repre-
sented in Figure 5.3. Vertical stripes representing positions of mechanical wear marks
are evident in the top and middle of the figure (representing lower axial positions
and rotary positions at the bottom of the pressure tube). However, it is clear that
mechanical wear marks, given in the Figure 5.4 C-Scan, are much less obvious than
those given by gx in Figure 5.3. Furthermore, it is difficult using the C-Scan alone to
differentiate between vertical stripes in the middle of the figure, representing location
of mechanical wear marks, versus vertical stripes at the sides of the figure (corre-
sponding with the top of the pressure tube), where no mechanical wear marks are
present.
5.5. APPLICATIONS 53

Figure 5.4: 16kHz C-Scan, plotting eddy current data orthogonal to direction of lift-
off vector, ˆl.

Figure 5.5 provides 16kHz C-Scan eddy current data, in the direction of lift-
off. Mechanical wear marks are more evident in this figure compared to Figure 5.4.
gx given in Figure 5.3 however still provides increased discrimination of wear mark
features. An increased number of vertical stripes revealing locations of wear marks
are apparent in Figure 5.3 over Figure 5.5. Furthermore, whereas mechanical wear
marks are not evident in axial positions between 5950mm and 6000mm in Figure 5.5,
these wear marks are visible for the same axial positions in Figure 5.3.
The SNR (signal-to-noise ratio) of mechanical wear mark indications is used to
quantitatively compare imaging capability of gx , against C-Scan representations. The
SNR for this application is defined as the rotational peak-to-peak mechanical wear
mark signal amplitude, averaged over all axial positions, divided by the background
5.5. APPLICATIONS 54

Figure 5.5: 16kHz C-Scan, plotting eddy current data in direction of lift-off vector, ˆl.

noise, which is computed as the standard deviation of noise pixel intensities. The
SNR of mechanical wear mark indications imaged via computation of gx , shown in
Figure 5.3, is computed as 5.8, while the SNRs of mechanical wear mark indications
shown in C-Scan Figures 5.4 and 5.5 are computed as 1.5 and 1.6, respectively. This
demonstrates increased wear-mark imaging capability of gx , over C-Scan representa-
tions.

5.5.2 Pressure Tube Constrictions

Figure 5.6 represents the computation of gy for the same eddy current data set em-
ployed to calculate gx , plotted in Figure 5.3.
Large horizontal bands, which extend 360 degrees circumferentially around the
pressure tube (evident at axial positions 5500mm and 6050mm) are shown in Figure
5.5. APPLICATIONS 55

5.6. These large horizontal bands represent the locations of fuel channel constrictions.
The rotary extent (360 degrees around the pressure tube) of the bands is consistent
with the circular formation of constrictions. Furthermore, the bands are approxi-
mately 0.5m apart, consistent with the length of fuel bundles where constrictions
form (Figure 5.1).
It is worthy to note that Figures 5.4 and 5.5 give C-Scan displays for the same
eddy current data revealing constrictions in Figure 5.6. Indications of constrictions
in Figures 5.4 and 5.5 coincide, and are inseparable from indications of mechanical
wear marks, contrary to indications of constrictions in Figure 5.6.

Figure 5.6: 16kHz gy revealing locations of local pressure tube constrictions (horizon-
tal stripes extending from 0◦ to 360◦ ).
5.5. APPLICATIONS 56

5.5.3 Spacers

Identification of spacer positioning can also be performed through identification of


the axial component of probe lift-off. Figure 5.7 shows clearly the position of the
spacer, where the spacer has locally induced a load on the pressure tube, forming an
inner-diameter pressure tube protrusion, as described in Section 5.3. As expected, the
formed protrusion is centred at the bottom of the pressure tube (near 180 degrees).
The extent of contact between pressure tube and spacers can also be inferred from
Figure 5.7. It is identified by the rotary arc for which the spacer response is made
visible, which can be approximated as 60 degrees rotationally.

Figure 5.7: 8kHz gy identifying the position of a spacer.


5.5. APPLICATIONS 57

5.5.4 Interpretation of Positive and Negative Intensities

Intensity values in images formed by the computation of gx and gy have physical


meaning. Consider Figure 5.7, which reveals the location of lift-off due to pressure
tube inner-diameter wall protrusion. Examination of the figure reveals that the spacer
location is identified by two horizontal bands having opposite intensity values. Recall-
ing Equations 5.4 and 5.5, positive intensity values correspond to probe lift-off action,
while negative intensity values correspond to return of the probe to the surface of the
pressure tube. The two horizontal bands having opposite intensity, evident in Fig-
ure 5.7 therefore correspond to locations of increase and decrease in lift-off, where
the probe first lifts off when encountering the pressure tube wall protrusion caused
by spacer loading (as in Figure 5.2), is at minimum lift-off where the protrusion is
maximized, and then extends to a surface riding configuration against the pressure
tube wall after traversing over the protrusion (in the direction of opposite tilt when
compared to initial lift-off). The ordering of positive, then negative intensities with
increased axial position is consistent with direction of probe travel (increasing in axial
position).

5.5.5 Confirmation of Spacer and Constriction Locations using Ultra-


sound Data

The inspection head used to collect eddy current data also collects ultrasonic non-
destructive examination (NDE) data. Ultrasonic NDE data can be used to confirm
position of spacers and constrictions identified through computation of gy .
Two ultrasonic probes diametrically opposed, and positioned on the inspection
5.5. APPLICATIONS 58

head, rotate at the same pitch as the eddy current probe. This allows for measure-
ments to be taken at 0 degree and 180 degree rotary positions, for all axial positions
along the pressure tube. The distance, d, between a probe face and the pressure tube
wall can be determined by the simple relationship

ct
d= , (5.6)
2

where c is the speed of sound in water, and t is the travel time between ultrasound
pulse emission from, and return to, the ultrasonic probe. The pressure tube inner
diameter along the 0 degree and 180 degree vertical line can then be computed

dP T −ID = d0 + dl + d180 , (5.7)

where d0 is the distance between the probe positioned at the 0 degree rotary position
and the pressure tube inner diameter wall, d180 is the distance between the probe
positioned at the 180 degree rotary position and the pressure tube inner diameter
wall, and dl is the distance between diametrically opposed probe faces.
gy for 8kHz eddy current data is shown in Figure 5.8, revealing three constriction
locations and a single spacer location. To confirm the position of these artefacts,
a comparison between locations identified through examination of gy , and locations
identified through examination of ultrasonic measurements (dP T −ID ) is made. A
correction factor is introduced for axial positions of dP T −ID . This accounts for the
difference between axial position of ultrasonic probe centres and axial position centred
between transmitting and receiving eddy current coils.
Figure 5.9 a) gives gy along axial positions between 3600mm and 5000mm, but
5.5. APPLICATIONS 59

averaged for points centred at the bottom of the channel; the average is calculated for
values of gy between rotary positions 160 degrees and 200 degrees. The points where
gy crosses the x-axis in the vicinity of visible signal perturbations (each characterized
by a large positive, then negative fluctuation) marks the position of spacers and
constrictions. These points lie on x-axis positions: 3695mm, 4195mm, 4478mm, and
4698mm.
Figure 5.9 b) gives ultrasonic measurements (dP T −ID ), along pressure tube axial
positions between 3600mm and 5000mm. Background variation in pressure tube inner
diameter has been removed. Positions of constrictions and spacers can be identified
by localized reductions in dP T −ID , plotted as a function of axial position. For localized
reductions, the pressure tube inner diameter is relatively smaller as compared to the
immediately surrounding regions, reflecting the characteristic protrusions at spacers
and constrictions.
The positions of lowest points in localized reductions can be used as markers for
the axial locations of spacers and constrictions, where axial locations are identified
as 3694mm, 4190mm, 4474mm, and 4702mm. Localized diameter changes can also
be estimated from Figure 5.9 b), where the first, second, and third constrictions have
associated diameter reductions 60µm, 70µm, and 90µm, respectively, while the spacer
has an associated diameter reduction of 170µm.
Comparing spacer and constriction positions determined through examination of
gy , and through ultrasonic measurement reveals consistency in artefact localization
between the two methods; measurements agree within +/- 5mm.
5.6. DISCUSSION 60

Figure 5.8: 8kHz gy identifying the positions of three constrictions and one spacer.

5.6 Discussion

5.6.1 Relationship to C-Scan Display

Eddy current data is collected rotationally every 1 degree (∼ 0.9mm) and axially
every 1mm. High data-density (typically output in C-Scan display for eddy current
applications), has been exploited in the surface profiling process.
It is worthy to note that images formed by the computation of gx (and gy ) differ
from C-Scan displays due to isolation of change in eddy current data, in the direction
of lift-off. C-Scan displays reveal the component of densely collected eddy current
data in one vector direction, usually orthogonal to lift-off. Images formed through
computation of gx and gy not only differ from C-Scan data by calculation of the
component of eddy current data parallel (not orthogonal) to the lift-off direction, but
5.6. DISCUSSION 61

make use of a maximum rate of change calculation, of eddy current signals in the
direction of lift-off (Equations 5.4 and 5.5). This key feature allows the visualization
of change in acquired eddy current data in the direction of lift-off. Without this,
features such as those visible in Figure 5.6, identifying position of constrictions would
be difficult to identify. Indeed, the position of constrictions cannot be identified from
C-Scan images (Figures 5.4 and 5.5) where the same eddy current data is employed
in generation of Figure 5.6.

5.6.2 Separability of Lift-off Localization in Rotary and Axial Directions

Although eddy current data is gathered in a helical trajectory by the eddy current
probe mounted on the fuel channel inspection head, changes in probe lift-off due to
mechanical interaction between the eddy current probe and pressure tube protrusions
are decomposed into rotary and axial directions. Indications visible in figures gen-
erated via computation of gx and gy represent probe lift-off in the rotary and axial
directions, respectively. gx can therefore be used to detect axially extending obstruc-
tions, and is robust against noise due to lift-off in the axial direction of probe motion.
Conversely, gy can be be used to detect rotationally extending obstructions and is ro-
bust against noise due to lift-off in the rotary direction of probe motion. As is shown
in Figures 5.3 and 5.6, rotationally extending obstructions (e.g. constrictions) do not
appear in the figure generated through computation of gx , while axially extending
obstructions (e.g. mechanical wear marks) do not appear in the figure generated by
gy . This demonstrates the high degree of separability between localization of probe
lift-off in rotary and axial directions, through computation of gx and gy .
5.7. CONCLUSIONS 62

5.6.3 Generalization of Results

Application of the techniques described in this study have focused on localization of


CANDU fuel channel component features. However, the surface profiling methods
provided can be generalized to other applications where the following requirements
are met: eddy current data with high data density is collected; direction of the lift-off
vector can be identified; the surface area of the eddy current probe is larger than the
dimensions of surface protrusions to be imaged; and, mechanical interaction between
the probe body and profiled surface is understood.
Separability of lift-off localization, allowing identification of protrusion orientation
relative to direction of probe motion is dependent on the number of degrees of freedom
of the probe. In the case of the eddy current probe used towards results presented in
this study, two degrees of freedom are allowed, and therefore, separability of lift-off
localization is available in two directions (axial and rotary directions of probe travel).
An increased number of degrees of freedom (e.g. tilt about the axial direction of probe
travel) will allow for increased characterization of protrusion orientation relative to
direction of probe motion.

5.7 Conclusions

Surface profiling of CANDU fuel channel pressure tube artefacts has been realized
through the methods laid out in this study. Mechanical wear marks, constrictions,
and local pressure tube deformation due to spacer loading can be identified through a
simple computational manipulation of high density eddy current data. These artefacts
range in size; mechanical wear marks have depressions ranging from 50µm to 100µm,
constrictions have protrusions in the same range, while garter spring spacers have
5.7. CONCLUSIONS 63

protrusions on the order of 200µm.


Detection capability of mechanical wear mark and constriction locations using
the presented surface profiling technique was shown to increase as compared to em-
ployment of C-Scan imaging. Mechanical wear marks were made clearly discernible
from background noise (SNR increased by over a factor of 3), while both mechani-
cal wear marks and constrictions could be independently localized, despite remaining
inseparable using C-Scan imaging.
Ultrasonic inspection data was used to independently verify the axial location
of garter spring spacers and pressure tube constrictions, against their location as
determined from the surface profile technique provided in this study.
Generalization of the presented technique towards profiling of surfaces other than
CANDU fuel channels has been considered. A base number of requirements for exten-
sion of the current set of technique applications have been identified. These are: high
data density of acquired eddy current inspection data; identification of lift-off vector
direction; greater surface area of the eddy current probe body, compared to dimen-
sions of surface protrusions to be imaged; and well understood mechanical interaction
between the probe body and the profiled surface.

Acknowledgements

The authors would like to thank, from Ontario Power Generation: Tulchand Harduwar
for supporting this work; as well as Andrew Hong and Jerry Piskorski for facilitating
access to source material.
5.7. CONCLUSIONS 64

Figure 5.9: a) gy computed from 8kHz eddy current data, averaged at the bottom
of the pressure tube. b) dP T −ID computed from ultrasonic measurement
data.
65

Chapter 6

Modelling and Validation of Eddy Current

Response to Changes in Factors Affecting Pressure

Tube to Calandria Tube Gap Measurement

Shaddy Shokralla, Sean Sullivan, Jordan Morelli, and Thomas W. Krause

Abstract: Procedures employed to non-destructively examine nuclear power plants


must undergo inspection qualification to ensure that they meet their respective in-
spection specification requirements. Modelling is a powerful tool that can be exploited
in the inspection qualification process. The gap between pressure tubes (PTs) and
calandria tubes (CTs) in CANDU (CANada Deuterium Uranium) fuel channels is
periodically measured, as contact can result in localized cooling and potential crack-
ing. This work shows how an analytical model can be employed to characterize the
effects of PT wall thickness and resistivity variation on gap measurement, and details
its validation against physical experiments.
Keywords: inspection qualification, multi-frequency, eddy current, resistivity
6.1. INTRODUCTION 66

6.1 Introduction

CANDU reactors contain over 400 fuel channels, each consisting of a 6 m long Zr
2.5% Nb pressure tube (PT) having 104 mm inner diameter (ID) concentric within a
129 mm ID Zircaloy-2 calandria tube (CT). Fuel bundles and heavy water required
for heat transport are contained within the PT. An annulus gas lies between the PT
and CT and insulates the hot ( 300 ◦ C) pressure tube from the ( 50 ◦ C) moderator
cooled CT. Diametral creep of the PT ID is induced by pressure, heat, and irradiation
induced creep, to a maximum ID of 111 mm, and a progressive decrease from nominal
4.2 mm wall thickness to an allowable minimum of 3.7 mm. PT to CT gap therefore
starts with an average of 8.3 mm and decreases with aging of the reactor. Potential for
contact between PT and CT is introduced by sag of the hotter PT within the cooler
CT. Hydride blister formation on the outer PT diameter (with risk of PT cracking) is
introduced with prolonged contact and PT to CT contact is therefore, to be avoided.
An eddy current based technique (complemented by ultrasonic PT wall thickness
measurements), which measures PT to CT gap, is used to monitor separation of PT
to CT. Figure 6.1 shows parameters of interest as well as the simplified measurement
configuration.
Qualification of an inspection system is an important step in determination of
the system’s capabilities against its inspection specification requirements, and has
become a nuclear operator regulator requirement [19]. Integral to the qualification
process is the study and determination of the effects of variation of an inspection
system’s essential parameters, those parameters whose change in value could affect
an inspection system such that its specified objectives would no longer be met [55].
Modelling is an important tool that has been promoted in the study of effects of
6.2. MEASUREMENT TECHNIQUE 67

essential parameter variation [20]. Modelling can be used to evaluate the effect of
specific parameter variations on gap probe response and therefore, is a tool for the
evaluation of gap measurement accuracy. However, models need to be validated in
order to provide assurance that they are providing the correct results. Examples doc-
umenting the validation of mathematical models for eddy current testing techniques
can be found in the literature [53, 54].
Mathematical models for realistic eddy current probe coils based on analytic so-
lutions to Maxwell’s equations were developed by Dodd, Deeds and other colleagues
[18, 16, 17]. A software tool which emulates Dodd and Deeds equations (expressing
analytical solutions to eddy current problems containing stratified conductors) has
been employed to determine the eddy current response to changes in essential param-
eters affecting accuracy of the PT to CT gap system. In this model the pressure tube
and calandria tube approximate parallel plate conductors for the eddy current probe
near the surface.
This paper documents validation performed to quantify the agreement between
modeled and physical responses to variation of essential parameters. Validation of
modelling effects of essential parameter variation on eddy current response was im-
plemented by comparing model output against physical measurements where PT to
CT gap, PT resistivity, and PT wall thickness were varied. Validation of the model
was critical to legitimizing its use in the inspection qualification process.

6.2 Measurement Technique

A transmit-receive eddy current probe acquires eddy current signals in a helical tra-
jectory. This probe is mounted on an inspection head, which is driven axially and
6.2. MEASUREMENT TECHNIQUE 68

Figure 6.1: Configuration of pressure tube to calandria tube gap measurement.

Figure 6.2: Geometries in close proximity to the gap probe. Pressure tube and Ca-
landria tube modelled as flat plates.
6.2. MEASUREMENT TECHNIQUE 69

rotationally along the pressure tube. The probe’s helical travel is approximated by a
series of circular trajectories, due to its axial pitch being small (1 mm), compared to
its rotational travel.
Transmit and receive coils belonging to the eddy current probe are situated along
the axial direction of the pressure tube when the inspection head is installed in-
channel. The probe is spring loaded such that it rides the surface of the pressure tube
[68]. The surface riding characteristic of the probe minimizes lift-off variations, which
would otherwise introduce additional variability in the probe response [5]. In addition,
deviations of lift-off from the ideal surface-riding condition, which could compromise
calibrated response of the probe to changes in gap, can be monitored using a gradient
response analysis as outlined in Ref. [68], helping to identify conditions where lift-off
variations may have compromised calibrated gap probe response. The center-to-center
spacing of the coil pair of 11 mm is much smaller than the nominal 330 mm inner
circumference of the PT. A schematic representation of coil positioning in relation to
other components is provided in Figure 6.2.
An Olympus NDT Multi-scan MS5800 eddy current instrument is used to induce
eddy currents and measure pickup coil response, while acquired signals are viewed in
real time. Multiple frequencies, including 4kHz, 8kHz, and 16kHz are used to acquire
eddy current signals. Unadjusted raw measurements have units mV .
Measurement of pressure tube to calandria tube gap is primarily a calandria tube
lift-off measurement, obfuscated by the presence of the pressure tube [69, 70]. Vari-
ations in eddy current probe response are monotonic with variation in the distance
between the probe and the calandria tube.
Calibration data is generated from measurement of multiple concentric pressure
6.2. MEASUREMENT TECHNIQUE 70

tube to calandria tube segments, where for each pressure tube to calandria tube com-
bination, pressure tube to calandria tube gap is varied from the expected minimum
to maximum. The concentric segments differ in pressure tube wall thickness, where
beginning of life (larger) to end of life (smaller) wall thicknesses are employed. Cal-
ibration data, along with ultrasonic measurement of in-channel pressure tube wall
thickness, are used to map eddy current data collected in-channel for each frequency
to in-channel pressure tube to calandria tube gap values [70].
Responses are attenuated for larger pressure tube wall thicknesses. The basic skin
depth equation describes shielding of a uniform electromagnetic field source near a
thick conductor, where the conducting domain is a homogeneous half-space and the
uniform magnetic flux density is directed tangentially to the surface of the material
[71, 5]:


J
= exp − x .
n o
J0 (6.1)
δ

Here, J is the eddy current density at distance x into the conductor, J0 is the surface
eddy current density and δ is the electromagnetic skin depth of penetration into the
conductor given by

r
ρ
δ = 50 . (6.2)
f µr

At δ, the eddy current density has been reduced to 1/e (36.8%) of the surface eddy
current density, J0 [5]. Equation 6.2 reveals that the effect of pressure tube resistivity
variation on eddy current voltage response for each frequency employed in pressure
6.2. MEASUREMENT TECHNIQUE 71

tube to calandria tube gap measurement is nonlinear.

6.2.1 Description of Samples

Carefully produced samples were used to measure responses to change in gap, wall
thickness, and resistivity. One key sample consisted of a singular pressure tube, fixed
within a calandria tube, where PT to CT gap between the pressure tube and calandria
tube varied between 0.8mm and 16mm. This sample also allowed for measurement of
responses to pressure tube wall thickness changes, as two of the pressure tube sections
were machined from their inner diameters, reducing their respective wall thicknesses.
The resulting pressure tube is made up of three sections with wall thickness 3.48 mm,
3.84 mm, and 4.36 mm (with a constant resistivity of ∼53.8 µΩ · cm at 20◦ C).
A set of samples was also produced to measure the response to changes in re-
sistivity and gap. These samples consisted of two composite pressure tubes. Each
composite pressure tube was made up of two pressure tube sections that had been cut,
machined, and welded together so that both could be fixed within the calandria tube.
Variations in composite tube wall thickness were small (a few microns) such that the
voltage response due to resistivity could be isolated, from potential wall thickness ef-
fects. The first composite pressure tube was made up of medium and high resistivity
sections (55.0 µΩ · cm and 57.4 µΩ · cm, respectively) while the second composite
pressure tube was made up of low and high resistivity sections (52.7 µΩ · cm and 57.2
µΩ · cm, respectively). The first composite pressure tube is shown in Figure 6.3.
6.3. ANALYTIC DETERMINATION OF EDDY CURRENT
RESPONSES TO CHANGES IN PHYSICAL PARAMETERS 72

Figure 6.3: Composite pressure tube and associated test rig assembly.

6.3 Analytic Determination of Eddy Current Responses to Changes in


Physical Parameters

Closed form solutions to Maxwell’s electromagnetic equations had been developed


by Dodd and Deeds for pancake coils near flat plates [16]. As previously applied
elsewhere [72, 73], the model uses the axisymmetric Dodd and Deeds method to
calculate the difference in signal induced between two coaxial receive coils. One coil
is larger, with a diameter that extends to the outer edge of the gap probe receive
coil. The second, smaller axisymmetric coil, extends to the inner edge of the gap
probe receive coil. The resulting differential signal produces the time rate of change
of magnetic flux in the annulus between the two receive coils. Since the resultant
magnetic flux is axisymmetric, with respect to the central transmit coil, the response
from the two differential coaxial detectors is considered approximately equivalent to
6.3. ANALYTIC DETERMINATION OF EDDY CURRENT
RESPONSES TO CHANGES IN PHYSICAL PARAMETERS 73

that from the receive coil of the real transmit-receive gap probe, but reduced by the
appropriate ratio of areas, namely that of the gap probe receive coil to that of the
annulus between the differential detectors.
A 2D model was assumed to be sufficient, since changes in gap are in the radial
direction with respect to the PT, and axial and circumferential directions can be
treated as planar to a first approximation, due to the relative localization of induced
eddy currents. As indicated in the introduction, the nominal ID for as-installed PT
is 104 mm and at end-of-life ID is 111 mm. The inner circumference of the PT is
therefore, 327 to 349 mm, as compared to the 12 mm x 11 mm axially oriented probe
sensing area of the transmit-receive coil pair, thereby justifying first-order treatment
by a planar geometry. It also neglects the potential effects of PT and CT curvature,
not considered a factor due to field spread on a 52 mm PT ID radius, compared
to the probe, which only has an 11 mm spacing. Furthermore, the mathematical
model assumes that the electromagnetic properties (electrical resistivity and magnetic
permeability) of the materials (PT, CT and Shield) are linear, homogeneous and
isotropic. Effects of any anisotropy in conductive material properties are ignored.

6.3.1 Modelling Requirements

The analytical model based on Dodd and Deeds equations was employed to model
eddy current response to a number of varying PT and CT conditions. The modelling
program was applied to determine the voltage developed in the receive coil of the
probe at frequencies 4, 8, and 16 kHz. Modelling conditions that were employed in the
simulation and subsequent validation are given in Table 6.1, where all permutations
of given parameters were modeled. The resistivities (varying between 45 µΩ · cm and
6.4. VALIDATION OF ANALYTICALLY DETERMINED
RESPONSES 74

60 µΩ · cm) bound resistivities for non-irradiated PTs [5, 74]. Similarly, the range of
wall thickness values (3.48 mm and 4.38 mm) employed in modelling was chosen to
emulate the range of wall thicknesses of in-service pressure tubes.

Table 6.1: Modelled Pressure Tube and Calandria Tube Parameters.

Parameter Values
0, 0.5, 1.0, 1.5, 2.0
PT to CT Gap (mm) 2.5, 3.0, 4.0, 5.0, 6.0
7.0, 8.0, 9.0, 11.0, 16.0
PT Wall Thickness (mm) 3.48, 3.76, 4.36
PT Resistivity (µΩ · cm) 45, 49, 52.4, 56.9, 60

6.4 Validation of Analytically Determined Responses

6.4.1 Validation of Responses to Changes in Wall Thickness

Validation of model responses to changes in wall thickness was obtained by compari-


son of modeled responses to change in both wall thickness and gap, against physical
measurements, where both wall thickness and gap were varied. Physical measure-
ments of eddy current responses to changes in wall thickness and gap employed in
the comparison made use of gap values (listed in Table 6.1) of 1.0 mm and greater,
as well as wall thickness values of 3.48 mm and 4.36 mm.
Impedance plane display of measured and modeled responses to variation in gap
(at 4kHz) where wall thickness is equal to 3.48 mm is presented in Figure 6.4. Model
results () from 0 to 16 mm gap are scaled to measured variation ( ) from 0.6 mm
to 16.3 mm gap. A spline curve follows the modeled points.
For the 4kHz, 8kHz, and 16kHz frequencies used, model outputs were linearly
scaled to match physical measurement output at 1 mm and 16 mm gap, for 3.48
6.4. VALIDATION OF ANALYTICALLY DETERMINED
RESPONSES 75

Figure 6.4: Impedance plane display of measured response to change in gap of 0.6
to 16.3 mm is shown by circles ( ), while scaled modeled response to
change in gap of 0 to 16 mm is shown by squares ().

mm wall thickness. A comparison of the modeled and measured amplitudes for the
different frequency signals, and an increase in wall thickness is shown in Figure 6.5.
Evident in examination of Figure 6.5 are the monotonic relationships between eddy
current voltage, wall thickness and gap; eddy current voltage magnitude increases
with gap, while it decreases with wall thickness.
To correct for a small mismatch in PT wall thickness between modelled responses
at 3.76 mm and measured responses at 3.84 mm, modelled responses at 3.84 mm
were generated via interpolation. For each gap value (listed in Table 6.1) of 1.0 mm
and greater, a cubic spline interpolation [75] was used to output voltage response at
pressure tube wall thickness of 3.84 mm using modeled responses at 3.48 mm, 3.76
6.4. VALIDATION OF ANALYTICALLY DETERMINED
RESPONSES 76

mm, and 4.36 mm as inputs.


Excellent agreement on the effects of wall thickness increase on eddy current re-
sponse is exhibited across all frequencies of interest (4kHz, 8kHz, and 16kHz). This is
obvious from examination of Table 6.2, where the average and standard deviation of
the agreement (defined in Equation 6.3 for particular gap and wall thickness values)
between modeled and measured voltage response has been computed.

|Vmeasured |
Agreement = × 100%. (6.3)
|Vmodelled |

The maximum effect of wall thickness increase from 3.48mm to 4.36mm on eddy
current amplitude (at 16mm gap), the maximum measured response (%), for the
different frequencies is also identified in Table 6.2.
Table 6.2: Statistical Computations for 4 kHz, 8 kHz, and 16 kHz Frequencies.

Parameter 4kHz 8kHz 16kHz


Average Agreement (%) 96.5 95.3 96.5
Std. Dev. of Agreement (%) 7.9 11.8 18.5
Max. Measured Response (%) -27.1 -34.8 -44.4

The effect of frequency on relative voltage response to wall thickness changes,


compared to response to change in PT to CT gap is evident in examination of Figure
6.5. Relative response to wall thickness changes increase with frequency; the 16kHz
signal responds the most to wall thickness changes, while the 4kHz signal responds
the least, in agreement with the skin depth expression, Equation 6.2. The inverse
relationship between frequency and eddy current penetration results in greater sen-
sitivity to wall thickness change with increased frequency. Conversely, a decrease in
frequency represents an increase in sensitivity to gap.
The relationship between frequency and response to wall thickness changes is not
6.4. VALIDATION OF ANALYTICALLY DETERMINED
RESPONSES 77

Figure 6.5: Effect of wall thickness variation on modeled versus measured eddy cur-
rent amplitudes for a) 4kHz, b) 8kHz, and c) 16kHz frequencies.
6.4. VALIDATION OF ANALYTICALLY DETERMINED
RESPONSES 78

only evident in comparing responses to bulk wall thickness changes (e.g. responses
between 3.48 mm, 3.84 mm, and 4.36 mm sections), but also responses to wall thick-
ness changes within individual sections where wall thickness was machined to a near
constant value. Deviations from a smooth monotonic curve representing the mea-
sured relationship between absolute voltage and gap are a result of small changes in
wall thickness about the circumference of the pressure tube. These deviations are
identified at the same gap values across Figures 6.5 a) to c), but are more prominent
for higher frequency responses, particularly at 16kHz.

6.4.2 Validation of Responses to Changes in Resistivity

Validation of modeled resistivity effects were performed by examining the change of


amplitude under minimum-to-maximum gap variation conditions. The presence of
even small variations in PT wall thickness could cause deviations in gap response and
therefore, only locations in the composite tubes (see Table 6.3) where wall thickness
at minimum and maximum gap could be matched, were chosen. Axial positions in the
composite tube samples were selected where wall thickness at minimum gap matched
wall thickness at maximum gap within 0.02 mm. As well, sections of pressure tube
were selected where nominal wall thickness was similar with other sections in the
sample set. In this case wall thicknesses ranged between 4.00 mm to 4.06 mm. Table
6.3 summarizes the measurements performed for composite resistivity pressure tube
samples employed in the study. Gap was varied between 0.4 mm and 16.3 mm for
these measurements.
Figure 6.4 shows the impedance plane display of the near gap probe response at
4 kHz. The amplitude between minimum and maximum gap, indicated by the solid
6.4. VALIDATION OF ANALYTICALLY DETERMINED
RESPONSES 79

Table 6.3: Resistivity and Wall Thickness Parameters for Measured Composite Re-
sistivity Pressure Tubes.

PT Change in
Number of Average Wall
Resistivity at Wall
Pressure Tube Measure- Thickness
20 ◦ C Thickness at
Section ments (mm)
(µΩ · cm) 0 ◦ and 180 ◦
4.03, 4.0, 0.0, 0.0, 0.0,
Medium ρ in 4 55.0
4.01, 4.015 -0.01
Med-High Tube
4.08, 4.08,
High ρ in Med- 4 57.4 0.0, 0.0, 0.02
4.08
High Tube
Low ρ in Low- 2 52.7 4.055, 4.05 0.01, 0.01
High Tube
High ρ in Low- 2 57.2 4.056, 4.055 0.0, -0.01
High Tube

line, is obtained from

p
Amplitude = (x0 − x1 )2 + (y0 − y1 )2 (6.4)

where 0.6mm gap occurs at (x0 , y0 ) and 16.3mm gap at (x1 , y1 ). The measured am-
plitude is used to scale the modeled gap probe response using one standard resistivity
pressure tube. Recorded horizontal and vertical voltages are in units of mV .
Figure 6.6 shows the measured amplitudes for 0.4mm to 16.3 mm change in gap,
obtained from the probe at 4 kHz for the various resistivity sections of the Low-
High and Med-High composite tubes, detailed in Table 6.3. Modeled results for an
equivalent change in gap at the same frequency, coil spacing and wall thickness have
been scaled to the experimental results at 52.7 µΩ · cm, while model results at other
resistivities have been scaled by the same amount. Table 6.4 summarizes standard
deviation, σ, between scaled model and experimental results and σ normalized by
gap amplitude, presented as a percentage. Excellent agreement between the two sets
6.4. VALIDATION OF ANALYTICALLY DETERMINED
RESPONSES 80

of measured amplitudes and model results is observed. At 4kHz a standard deviation


of 29.5 mV between experimental and modeled results for the 11 measurements is
observed, where the standard deviation between the scaled model and experimental
results is given as:

v
u N
uX (f (ρi ) − Ai )2
σ=t (6.5)
i=1
N −1

where ρi is the ith resistivity and Ai is the corresponding amplitude as obtained


from a quadratic polynomial best fit to the scaled model results, Equation 6.4. The
relative standard deviation, obtained by scaling with the amplitude at 52.7 µΩ · cm,
is 0.9%. Small variations in experimental amplitude are attributed to differences in
wall thickness in the composite tube sections as indicated in Table 6.3. Similar results
are observed at 8 kHz and 16 kHz. However, the relative standard deviation between
model and measurements has increased slightly to 1.1% and 2.6%, respectively. The
increasing σ with frequency is attributed to the effect of wall thickness variations on
gap response, which increase with frequency, particularly at 16 kHz.

Table 6.4: Statistical comparisons between modeled and measured results.

Ratio of
Composite Amp. vs ρ Amp.
Frequency σ σ/A Amplitude rate at 54 variation
(kHz) (mV) (%) Scaled to 4 µΩ · cm with ρ to wall
KHz (mV /µΩ · cm) thickness
(µΩ·cm/mm)
4 29.5 0.9 1 51.7 25
8 28.5 1.1 1.2 85.0 17
16 26.4 2.6 3.3 121 5.7
6.5. DISCUSSION 81

Figure 6.7 shows the gap probe response rate of change with resistivity as a func-
tion of frequency, obtained from the slope at 55 µΩ · cm in the scaled model curve
in Figure 6.6 as well as slopes for 8kHz and 16kHz frequencies. Slope values are
shown in the second last column of Table 6.4. These results indicate that the effects
of resistivity on gap response are greater at higher frequencies, again consistent with
Equation 6.2.
The last column in Table 6.4 shows the ratio of rate of change of gap amplitude
with resistivity relative to that of wall thickness as obtained from model results. The
values show that the effect of resistivity, relative to wall thickness, on gap measure-
ment decreases with increasing frequency. The largest ratio arises at 4 kHz frequency
(25 µΩ · cm/mm) indicating that this frequency is the most impacted by changes in
resistivity when compared with response due to PT wall thickness variations. The
next largest, with similar order of magnitude (17 µΩ · cm/mm), is the 8 kHz ampli-
tude response. These results are consistent with the relative % error, presented in
the second column of Table 6.4. The relative % error shows a progressive increase
in % standard deviation with increasing frequency that results in increased scatter,
attributed to the small wall thickness variations in this tube.

6.5 Discussion

6.5.1 Modelling Parameter Continuity

Although experimental results have been employed to show agreement with a discrete
number of modeled results, an advantage to modelling effect of parameter variation
is the ability to model all values within a range of an essential parameter (or ranges
of essential parameters). This is obvious, considering voltage responses are through
6.5. DISCUSSION 82

Figure 6.6: Amplitude as a function of resistivity for (a) 4kHz, (b) 8kHz, (c) 16kHz
frequencies for scaled model and experimental data from composite tubes.
6.5. DISCUSSION 83

Figure 6.7: Variation of scaled model based on gap probe response amplitude (0.5
to 16 mm gap) with resistivity at 55 µΩ · cm plotted as a function of
frequency.

Dodds and Deeds equations [18, 16, 17], which are analytic and closed form. However,
this presents clear advantages to examining effects of essential parameters in the
context of inspection qualification. Effects of essential parameter variation can be
observed to an arbitrary level of specificity, which can be performed rapidly and
inexpensively, compared to performing a large number of experiments.
In the model results presented here, the relative sensitivity of the eddy current
based gap response to bounding pressure tube wall thickness and resistivity variations
has been studied. The model results have assisted in the identification of frequencies
most sensitive to either the effects of wall thickness or resistivity, relative to the desired
PT to CT gap response. This information can be applied to select the most robust
frequency, that which is most independent of the essential parameter variations, for
in-reactor gap measurement.
6.6. CONCLUSIONS 84

6.5.2 Parallel Plate Approximation

Voltage responses to wall thickness and gap have been modeled by Dodds and Deeds
equations [18, 16, 17], where the pressure tube and calandria tube geometries (relative
to the eddy current probe) have been approximated as parallel conductive plates. For
the probe configuration employed in this study, this has shown not to be a significant
source of error. However, if select probe parameters are modified, the parallel plate
approximation may cease to approximate pressure tube to calandria tube geometries
to sufficient accuracy. If, for example, frequency is lowered well below 4kHz, there is
potential for increased eddy current penetration for larger PT and CT surface areas,
which, relative to the probe surface, no longer approximate parallel plates. Similarly,
if the probe was situated away from the surface of the pressure tube, curvature of PT
and CT components relative to the probe surface would discount their approximation
as paraellel plates.

6.6 Conclusions

Validation of a model based on Dodds and Deeds equations [18, 16, 17] used to es-
timate eddy current voltage response to PT wall thickness and resistivity changes
(with varying PT to CT gap) has been performed via comparison with physical mea-
surements.
For wall thickness values between 3.48mm and 4.36mm, and gap values between
1mm and 16mm, average agreement exceeded 95 % for all frequencies used, despite
small machining deviations (microns) in the wall thickness of test pieces, which re-
sulted in an exaggeration of disagreement between measured and modeled results.
Standard deviation of the agreement ranged between 7.9 % (4kHz signal) and 18.5
6.6. CONCLUSIONS 85

% (16kHz signal). Increased standard deviation with frequency is a result of larger


sensitivity to wall thickness variations with higher freuquency. This follows from an
inverse relationship between eddy current skin depth and frequency.
Gap measurements performed on composite pressure tube (PT) sections with dif-
ferent resistivities (52.7, 55.0, 57.2 and 57.4 µΩ·cm) were compared with 2D modelling
results obtained with equivalent gap probe parameters. Gap signal amplitudes for a
0.4 mm minimum to 16.3 mm maximum gap were evaluated at axial positions in
composite PT sample sections where pressure tube wall thickness was in the range of
4.0 to 4.1 mm and differences in wall thickness between 0.4 and 16.3 mm gap were
at a minimum (0.01 mm). The minimum wall thickness variation criterion was main-
tained for the probe, with its sensing area of 12mm x 11mm, giving good agreement
between measured and scaled model results. Results demonstrated consistent agree-
ment between modeled and laboratory measured dependence of gap signal amplitude
variation with resistivity over all frequencies. The observed agreement validates the
2D analytical models for the range of resistivities tested at a nominal wall thickness
of 4 mm.
Validated employment of an analytical model to measure the effect of essential
parameter variation on inspection system output has significant implications for the
inspection qualification process. Modelling can serve as a versatile and efficient tool
in quantifying the effect of essential parameter variation on system performance. This
is necessary to determine if a system meets its inspection specification requirements.
6.6. CONCLUSIONS 86

Acknowledgements

The authors wish to thank: Ontario Power Generation management for supporting
this work; as well as Stuart Craig from Atomic Energy of Canada Limited and John
Sedo for valuable discussions.
87

Chapter 7

Principal Components Analysis of Multi-frequency

Eddy Current Data Used to Measure Pressure

Tube to Calandria Tube Gap

Shaddy Shokralla, Jordan Morelli, and Thomas Krause

Abstract: Principal components analysis (PCA) involves transforming a set of cor-


related observations into a set of linearly uncorrelated variables, which can reveal
simplified trends in data. Multifrequency eddy current testing contains correlations
across different test frequencies. In this work, PCA is used to extract unique informa-
tion from multifrequency eddy current data sets, used to measure the pressure tube
(PT) to calandria tube (CT) gap, in CANDU (CANada Deurterium Uranium) fuel
channels. Advantages include compressed data acquisition, allowing for increased in-
spection speed, and monitoring for variation in physical parameters using a reduced
number of variables. PCA employing analytical input model data is validated against
PCA employing data from physical experiments.
Keywords: principal components analysis, multi-frequency, eddy current, pressure
tube, gap
7.1. INTRODUCTION 88

7.1 Introduction

Multi-frequency eddy current non-destructive evaluation (NDE) is employed to pro-


vide enhanced sensitivity to test surface condition, over single frequency eddy current
NDE [52, 76]. Advantages are: different depths of penetration for each frequency,
varying sensitivity to lift-off or fill factor, variable noise sensitivity, ability to vary rel-
ative drive amplitudes and capability to perform multi-frequency mixing for removal
of signal artefacts.
Measurement of pressure tube (PT) to calandria tube (CT) gap, necessary for
monitoring the separation between PT and CT for CANDU fuel channels is one
application that has exploited multi-frequency eddy current NDE to provide for col-
lectively greater sensitivity to changes in PT to CT gap [3]. Different excitation
frequencies (4 kHz, 8 kHz, and 16 kHz) respond with varying sensitivity to changes
in PT to CT gap, in addition to other confounding factors, and collectively provide
enhanced response over application of only an individual frequency, in addition to
providing additional measurement redundancy. Factors affecting pressure tube to
calandria tube gap measurement have been analytically modelled using Dodd and
Deeds equations [16, 77]. This has been shown to provide practical benefit to the
inspection qualification process [20, 77], including the ability to model eddy current
response to any combination of relevant factors (PT wall thickness, PT resistivity,
and PT to CT gap) affecting gap measurement.
Principal components analysis (PCA) involves employing orthogonal transforma-
tions to convert a set of observations, which are correlated, into a set of linearly
uncorrelated variables, called principal components [11, 12]. The main advantage of
PCA is that complex data sets can be reduced to data sets of lower dimension to
7.1. INTRODUCTION 89

reveal unapparent and/or simplified trends in the data [11, 12]. Particular to conven-
tional eddy current testing, principal components analysis (PCA) has been employed
to increase reliable interpretation of steam generator tube signals [37]. PCA has also
been applied to pulsed eddy current, and was first shown to provide enhanced classi-
fication of defects [38]. Further applications of PCA to pulsed eddy current include
detection of defects in multilayer aluminum lap joints [39, 40, 49], steel [42], and
aircraft structures [43, 44, 45]. Independent component analysis (ICA), a technique
used to separate multivariate signals into independent non-gaussian signals, has also
been used on eddy current data towards flaw characterization [78]. Integration of ICA
and PCA has been used for defect detection and separation for eddy current pulsed
thermography video [79].
Inspection speed is a critical factor in non destructive examination of nuclear com-
ponents, such as steam generator tubes [80], since shortened inspections will reduce
facility downtime. Techniques which have been employed to increase eddy current
inspection speed of steam generator tubes, have involved multiplexing, simultaneous
injection, or employment of eddy current arrays [81, 82, 83]. This is the first work
to apply PCA to a multi-frequency eddy current application (using analytically gen-
erated model data), where the ability to monitor variation in physical parameters of
interest, using a reduced number of variables is shown. Acquisition of the reduced
number of variables is an alternative to previously employed strategies for increasing
speed of eddy current inspection, as relevant inspection data acquired in real time
is significantly reduced. PCA using model data is validated against PCA data from
physically acquired data.
7.2. PRESSURE TUBE TO CALANDRIA TUBE GAP
MEASUREMENT 90

7.2 Pressure Tube to Calandria Tube Gap Measurement

Acquisition of multi-frequency (4 kHz, 8 kHz, and 16 kHz) eddy current data is per-
formed by a transmit-receive eddy current probe. The eddy current probe is mounted
on an inspection head, which is driven axially, while rotating circumferentially within
a pressure tube [70].
The eddy current probe is spring loaded and therefore makes contact with the
pressure tube surface during acquisition of gap measurement data. Also, eddy current
probe transmit and receive coils are aligned in the axial direction of the pressure tube.
Due to a relatively small center-to-center spacing between the transmit and receive
coil (11 mm), compared to the 330 mm pressure tube inner circumference, a planar
approximation of the geometries in proximity to the gap probe has been employed
[77]. Figure 7.1 identifies coil positions, relative to other component geometries.
Eddy current probe amplitude response increases with increased gap variation,
as this interaction is fundamentally a calandria tube lift-off response [70, 69]. Data
from multiple non-concentric pressure tube to calandria tube segments is acquired to
generate calibration data. In calibration sets the pressure tube wall thickness is varied
between segments, representing beginning of life, mid-life, and end-of-life conditions,
since pressure tube wall thickness decreases with in-reactor life [3]. For each eddy
current frequency, calibration data combined with eddy current data acquired in-
channel along with ultrasonic measurements, are used to generate in-channel gap
measurements [70].
The skin depth equation for a plane magnetic field incident on a planar surface
[5] is given by
7.2. PRESSURE TUBE TO CALANDRIA TUBE GAP
MEASUREMENT 91

Figure 7.1: Material components in the vicinity of the gap probe, where pressure tube
and calandria tube are approximated as flat plates.


J
= exp − x ,
n o
J0 (7.1)
δ

where J0 is the surface eddy current density, J is the eddy current density at x
mm into the conductor, and δ is the electromagnetic skin depth (mm) of penetration
into the conductor. δ is given by [5]

r
ρ
δ = 50 , (7.2)
f µr

where ρ is resistivity (µΩ·cm), f is frequency (Hz), and µr is relative permeability.


7.3. PRINCIPAL COMPONENTS ANALYSIS 92

Eddy current density at δ has been reduced to 1/e of J0 [5].


Eddy current amplitude response to changes in gap are attenuated for larger
pressure tube wall thicknesses, as the density of eddy currents decrease with increasing
x, as described by Equation 7.1. Furthermore, Equation 7.2 identifies that decreased
pressure tube resistivity, and conversely, increased probe excitation frequency, result
in decreased skin depth, also incurring attenuation of eddy current amplitude response
to the presence of the CT.
The combined effects of sensitivity to lift-off and skin depth form the basis for
explaining PT to CT gap measurement.

7.2.1 Analytical Model

A 2D analytical model has been shown to be sufficient for approximating responses to


changes in PT to CT gap [77]. This 2D model is derived from closed form solutions
to Maxwell’s electromagnetic equations, as developed by Dodd and Deeds [16], and
also applied in other works [72, 73].

7.3 Principal Components Analysis

A data set is represented as X = [x1 , x2 , . . . , xp ], an N × p matrix, where the means


of x1 , x2 , . . . , xp are normalized to zero. Where Σ = X T X/(N − 1) is the covariance
matrix, eigenvectors and eigenvalues of Σ are given by

U T ΣU = Λ. (7.3)

Columns of U , eigenvectors of Σ, are principal components, and diagonal matrix


entries of Λ, λ1 , . . . , λp are eigenvalues of Σ [11, 12], where λ1 ≥ λ2 ≥, . . . , ≥ λp .
7.3. PRINCIPAL COMPONENTS ANALYSIS 93

When raw data vectors in X are in different units, the correlation matrix can be
used instead of the covariance matrix to scale data appropriately [11]. To employ the
correlation matrix for PCA, vectors xi , i ∈ {1, . . . , p} are scaled by 1/σi , where σi2 is
the variance of xi .
Principal component scores Z = [z1 , z2 , . . . , zp ] are computed as Z = XU .
Eigenvalues λ1 , . . . , λp are the variances of the principal components, where the
percent variance explained by values of principal component i (scores) is given by
P
λi / j λj × 100%.
The original data set X can be recovered from principal component scores via the
computation X = ZU T .
Pm P
Principal component scores z1 , z2 , . . . , zm , i=1 λi / j λj ≈ 1, contain the major-
ity of variance accounted for in principal component scores z1 , z2 , . . . , zp . Therefore,
principal components u1 , u2 , . . . , um can be used in computing a subset of the full set
of principal component scores, and also approximating the original data set X from
this same subset. Considering Z = XU , computation of z1 , z2 , . . . , zm , and recon-
structing X (represented by the approximation X̃) from these scores is performed via
the following transformations,

[z1 , z2 , . . . , zm ] = X[u1 , u2 , . . . , um ], (7.4)

and

X̃ = [z1 , z2 , . . . , zm ][u1 , u2 , . . . , um ]T . (7.5)


7.4. PCA APPLIED TO ANALYTICAL MODEL DATA 94

7.4 PCA Applied to Analytical Model Data

The analytical model based on Dodd and Deeds equations [16] was employed to model
eddy current response to a number of varying PT and CT conditions. A modelling
program emulating Dodd and Deeds equations [77, 84] was applied to determine the
voltage developed in the receive coil of the probe at frequencies 4 kHz, 8 kHz, and
16 kHz. Modelling conditions that were employed in the simulation and subsequent
validation are given in Table 7.1, where all permutations of given parameters were
modeled. The resistivities (varying between 45 µΩ · cm and 60 µΩ · cm) bound
resistivities for non-irradiated PTs [5, 74]. Similarly, the range of wall thickness
values (3.48 mm and 4.36 mm) employed in modelling was chosen to emulate the
range of wall thicknesses of in-service pressure tubes.

Table 7.1: Modelled Pressure Tube and Calandria Tube Parameters.

Parameter Values
0, 0.5, 1.0, 1.5, 2.0
PT to CT Gap (mm) 2.5, 3.0, 4.0, 5.0, 6.0
7.0, 8.0, 9.0, 11.0, 16.0
PT Wall Thickness (mm) 3.48, 3.76, 4.36
PT Resistivity (µΩ · cm) 45, 49, 52.4, 56.9, 60

7.4.1 PCA Analysis

The Matlab function pca [85] was applied to the analytical model data, where resistive
and reactive components of individual frequency responses were represented as row
vectors in X (Sect. 7.3). A set of principal component scores was output from
the PCA analysis. The percentage of the total variance explained by each principal
component (ordered in terms of significance) is as follows: 96.92%, 2.76%, 0.29%,
7.4. PCA APPLIED TO ANALYTICAL MODEL DATA 95

0.02%, 0.00%, and 0.00%.


As individual frequency (4 kHz, 8 kHz, 16 kHz) responses to PT to CT gap,
PT resistivity, and PT wall thickness are correlated (all frequencies are sensitive to
changes in these quantities), the large weighting on the first principal component,
and small weightings on subsequent components is expected.
The large weighting on the first principal component demonstrates the principle
of data compression. Where the original eddy current data was represented in 6
dimensions, ∼ 97% of the data variance is explained by 1 dimension, post PCA
transformation.
Figure 7.2 is useful in identifying the relative effect of varying any of the physical
parameters against principal component 1, 2, and 3, scores. Increasing resistivity has
the effect of shifting modelled data in the positive principal component 1 direction;
clusters of modelled data (with varying gap and wall thickness values) are plotted for
each resistivity. Each of these clusters is shown as three strings of points, where each
string has different gap values, for each of the wall thicknesses modelled (see Table
7.1 for values of modelled physical parameters).
For a particular gap and resistivity combination, increasing wall thickness has the
effect of shifting modelled data in the negative principal component 1 direction, but in
either component 2, and 3 directions, dependent on the particular gap and resistivity.
For the case of a particular resistivity and wall thickness combination, increasing
gap has the effect of shifting modelled data in the positive principal component 1
direction, and also in either component 2 and 3 directions, dependent on the particular
resistivity and wall thickness.
Figure 7.2: PCA Scores 1, 2, and 3, plotted together. Analytically modelled data is clustered in five groups,
corresponding to the different modelled resistivity values as labeled. As indicated in Table 7.1, each
cluster has a 0 to 16 mm gap change for three different wall thicknesses.
7.4. PCA APPLIED TO ANALYTICAL MODEL DATA
96
7.5. COMPRESSED DATA ACQUISITION 97

7.4.2 Relationship to Skin Depth Equation

As mentioned in Section 7.2, measurement of pressure tube to calandria tube gap


is primarily a calandria tube lift-off measurement, obfuscated by the presence of the
pressure tube [70, 69]. Variations in eddy current probe response are monotonic with
variation in the distance between the probe and the calandria tube.
Equation 7.2 reveals that the effect of pressure tube resistivity variation on eddy
current voltage response for each frequency employed in pressure tube to calandria
tube gap measurement is direct, and nonlinear. The skin depth equation describes
the constant in the exponent for the rate of exponential decay. So it is evident that
increased wall thickness change has an inverse exponential relationship with the eddy
current density induced on the calandria tube.
The relationship between resistivity, wall thickness, and minimum-to-maximum
gap (0 mm to 16 mm) voltage response is represented concisely in Figure 7.3. The
peak-to-peak variation of principal component score 1, having the largest variance,
is calculated for every wall thickness and resistivity pairing. Clearly, peak-to-peak
voltage varies directly with resistivity, while it varies inversely with wall thickness,
consistent with Equations 7.1 and 7.2.

7.5 Compressed Data Acquisition

Figure 7.4 provides a flow chart illustrating the process of compressed data acquisition
for multi-frequency eddy current data, where PCA has been employed, as presented
in Section 7.3. In the described process, data is acquired along the first m principal
component vectors, where m is less than p, the total dimension of multi-frequency
eddy current data. p = 2 × q, where q is the number of frequencies employed, and 2
7.5. COMPRESSED DATA ACQUISITION 98

Figure 7.3: Score 1 peak-to-peak variation (due 0 to 16 mm change in gap) for model
data.

represents horizontal and vertical voltages plotted on the impedance plane. Process
hardware (performing arithmetic operations) [86], can be used to linearly transform
p dimensional multi-frequency eddy current data into m dimensional PCA scores (via
Equation 7.4), thereby realizing data compression during real time acquisition.
Original uncompressed multi-frequency eddy current data sensed at the probe, can
be reconstructed from compressed data, by a simple linear transformation (Equation
7.5), again employing m principal components. Process hardware or software can
be employed during the reconstruction step, as this is performed post-acquisition.
The closer m is to the dimension of the original uncompressed multi-frequency eddy
current data (p), the more the reconstructed signals will represent the originally sensed
7.5. COMPRESSED DATA ACQUISITION

Figure 7.4: Data Acquisition Flowchart.


99
7.5. COMPRESSED DATA ACQUISITION 100

signals, prior to compression. For multi-dimensional correlated data sets, however, as


will be demonstrated, m does not need to approach p for an accurate reconstruction
to be performed.
Six-dimensional multi-frequency eddy current data used for PT to CT gap mea-
surement is highly correlated with respect to changes in PT wall thickness, PT resis-
tivity, and PT to CT gap, and can therefore be compressed and reconstructed using a
lower number of principal components. Figure 7.5 illustrates reconstruction capabil-
ity for 16 kHz analytically modelled eddy current data, which has been compressed
using a lower number of principal components, than the dimension (six) of the orig-
inal multi-frequency data set. All voltage responses for 4 kHz, 8 kHz, and 16 kHz
frequencies, for all permutations of values listed in Table 7.1 were used to generate
principal components. Plotted in Figure 7.5, gap is varied from 0 mm to 16 mm
(gap values provided in Table 7.1). PT wall thickness and resistivity are fixed at 3.76
mm and 52.4 µΩ·cm, respectively. Analytically modelled, uncompressed eddy current
data, is shown to be reconstructed from a lower number of principal components than
the original six-dimensional eddy current data. The greater the number of principal
components used in compression and reconstruction of originally uncompressed ET
data, the greater the reproducibility of the original data set.
Table 7.2 provides a comparison of reproduction capability, dependent on the
number of principal component scores utilized in compression and reconstruction.
All voltage responses, derived from combinations of physical parameters identified in
Table 7.1 for 4 kHz, 8 kHz, and 16 kHz frequencies were used in this analysis (675
data points). Mean and standard deviation of % error are provided, charted against
number of principal components. Amplitude error (for a given ET response, for a
7.5. COMPRESSED DATA ACQUISITION 101

Figure 7.5: Reconstructed 16 kHz ET data on Lissajous plot.

given PT resistivity, PT wall thickness and PT to CT gap, represented on a Lissajous


plot) is given by

p
Amplitude error = (x0 − x00 )2 + (y0 − y00 )2 , (7.6)

where (x0 , y0 ) is the original uncompressed voltage response (resistive and reac-
tive components) and (x00 , y00 ) is the reconstructed voltage response. % error for an
approximated voltage response is given as

amplitude error
% error = , (7.7)
max gap response

where max gap response is defined as the maximum amplitude response due to a
7.6. EXPERIMENTAL VALIDATION 102

0 mm to 16 mm gap change (for 16 kHz frequency, 3.48 mm PT wall thickness, and


60 µΩ · cm PT resistivity).
Confirming the previous assertion, the greater the number of principal components
used for data compression and reconstruction, the greater the reconstruction fidelity.
Using m = 3 (half the dimension of the original data set) principal components, mean
and standard deviation of % error are less than 1 %.
Table 7.2: Mean and standard deviation of percentage voltage error.

Principal Mean Standard


Component(s) % Error Deviation
of % Error
1 10.20 7.14
2 3.11 2.49
3 0.85 0.6
4 0.07 0.07
5 0.02 0.02
6 0.00 0.00

7.6 Experimental Validation

Two sets of samples with changing pressure tube wall thickness, pressure tube resis-
tivity, and pressure tube to calandria tube gap were used for experiments discussed in
this paper. The first sample varied PT to CT gap continuously between 0.8 mm and
16 mm. Pressure tube to calandria tube wall thickness also varied, as two PT sections
had their inner diameters machined to different depths. To summarize, this sample
was made up of three sections with wall thicknesses 3.48 mm, 3.84 mm, and 4.36 mm
(resistivity of the PT was ∼ 53.8 µΩ · cm at 20 ◦ C). The experimental configuration
is described in detail elsewhere [77].
The other sample set that was used included changes to PT resistivity and gap.
7.6. EXPERIMENTAL VALIDATION 103

Two different pressure tube samples, each with differing resistivity (51.38 µΩ · cm
and 55.95 µΩ · cm) were cut and welded, and their respective wall thicknesses were
machined such that they were equal within a few microns. A composite pressure tube
was housed within a calandria tube such that the gap varied continuously, from just
above 0 mm to just below 16 mm [77].

7.6.1 Wall Thickness Variation

Validation of model responses to changes in wall thickness was performed by compar-


ison of modeled responses to change in both wall thickness and gap, against physical
measurements, where both wall thickness and gap were varied. Physical measure-
ments of eddy current responses to changes in wall thickness and gap employed in
the comparison made use of gap values (listed in Table 7.1) of 1.0 mm and greater,
as well as wall thickness values of 3.48 mm and 4.36 mm.

PCA and Regression

Weighted PCA (where the inverse variable variances are used as weights) was per-
formed on physical measurements of eddy current responses to changes in PT wall
thickness and PT to CT gap (detailed in Section 7.6). The percentage of the total
variance explained by each principal component (ordered in terms of significance) is
as follows: 87.94%, 11.66%, 0.36%, 0.03%, 0.00%, and 0.00%.
A polynomial surface of the form

f (x, y) = p00 + p10 x + p01 y + p20 x2 + p11 xy + p02 y 2 , (7.8)

where x represents PT to CT gap (mm), and y represents PT resistivity (µΩ·cm),


7.6. EXPERIMENTAL VALIDATION 104

is fit to the first principal component score values shown in Figure 7.6 (blue circles),
such that f (x, y) = 48.82 + 0.91x − 22.87y − 0.01694x2 − 0.1077xy + 2.439y 2 .
Towards validating the model data against experimental data, PCA was performed
on a subset of the model data, where parameters were restricted to those with values
near the physical parameter values present in the physical experiment, where PT wall
thickness and PT to CT gap were varied. More specifically, PT resistivity was fixed at
52.4 µΩ · cm, while CT resistivity was fixed at 74 µΩ · cm. The percentage of the total
variance explained by each principal component (ordered in terms of significance) is
as follows: 87.98%, 11.90%, 0.09%, 0.03%, 0.00%, and 0.00%.
A polynomial surface of the same form used in surface regression for PCA of
experimental data (Equation 7.8), is fit to the first principal component score values
for the model data, shown in Figure 7.6 (yellow data), such that f (x, y) = 52.05 +
1.862x − 24.91y − 0.02486x2 − 0.3097xy + 2.784y 2 .
Examination of Figure 7.6 reveals that polynomial fits with respect to the exper-
imental and model data (restricted to approximate physical parameter values) are
similar, serving to validate the PCA approach for PT wall thickness and PT to CT
gap variations.

7.6.2 Resistivity Variation

Eddy current scans of a composite pressure tube, having two sections with different
resistivities (51.38 µΩ · cm and 55.95 µΩ · cm) were taken. A calandria tube enclosed
the pressure tubes during the scan, to simulate in-channel conditions. For each axial
position along the pressure tube, gap varied uniformly between minimum and max-
imum (i.e. cross sectional gap profiles are identical along the pressure tube in the
7.6. EXPERIMENTAL VALIDATION 105

Figure 7.6: Component 1 score with regression surfaces, for variations in PT wall
thickness and PT to CT gap. Experimental data is shown in blue while
analytical model data is shown in yellow.

axial direction).

PCA and Regression

Weighted PCA (where the inverse variable variances are used as weights) was per-
formed on physical measurements of eddy current responses to changes in PT resis-
tivity and PT to CT gap (detailed in Section 7.6.2). The percentage of the total
variance explained by each principal component (ordered in terms of significance) is
as follows: 81.46%, 18.05%, 0.43%, 0.05%, 0.01%, and 0.00%.
A polynomial of the same form as Equation 7.8, where x represents PT to CT
7.6. EXPERIMENTAL VALIDATION 106

gap (µΩ · cm), and y represents PT wall thickness (mm), is fit to the first princi-
pal component score values shown in Figure 7.7 (blue circles), such that f (x, y) =
−20.57 + 0.337x + 0.309y − 0.033x2 + 0.010xy + 0y 2 .

Figure 7.7: Component 1 score with regression surfaces, for variations in PT resis-
tivity and PT to CT gap. Experimental data is shown in blue while
analytical model data is shown in yellow.

Towards validating the model data, PCA was performed on a subset of the model
data. Parameters were restricted to those with values near physical parameter values
present in the physical experiment, where PT resistivity and PT to CT gap were
varied. More specifically, PT wall thickness was fixed at 4.36 mm, while CT resistivity
was fixed at 74 µΩ · cm. The percentage of the total variance explained by each
principal component (ordered in terms of significance) is as follows: 88.32%, 11.11%,
0.54%, 0.03%, 0.00%, and 0.00%.
7.7. DISCUSSION 107

A polynomial surface of the form of Equation 7.8 is fit to the first principal
component score values shown in Figure 7.7 (yellow circles), such that f (x, y) =
−46.98 − 0.3423x + 0.8312y − 0.018x2 + 0.01466xy + 0y 2 .
Examination of Figure 7.7 reveals that polynomial fits with respect to the experi-
mental and model data exhibit similar trends, serving to validate the PCA approach
for PT resistivity and PT to CT gap variations.

7.6.3 Differences Between Experimental and Analytical PCA

The weighted PCA of both the experimental and analytical model data, generated
via Dodd and Deeds equations [16] generally agree and reveal similar trends (Figures
7.6 and 7.7). Disagreement in PCA output for the two differently generated data
sets can be attributed to difference in experimental and analytical phase response to
variations in physical quantities: PT wall thickness, PT resistivity, and PT to CT
gap. Limitations of Dodd and Deeds equations may be the result of the multi-coaxial
coil approximation used to model the experimental transmit-receive configuration
[77]. This may be the cause for minor difference between experimental and analytical
response, exhibited by score 1. Agreement between analytical and experimental am-
plitude responses of the entire signal to variations in these physical quantities have,
however, been validated [77].

7.7 Discussion

7.7.1 Normalization of Phase and Scale

Employing weighted PCA in representing ET data requires a normalization of phase


and scale. Conventional ET applications involve rotating the phase of the ET response
7.7. DISCUSSION 108

to allow for the signals of interest to align in the vertical direction of the impedance
plane. Employing PCA allows for this rotation to be emulated by computation of
principal components; direction of subsequent principal component vectors, identify
subsequent directions of maximum data variance.
Normalization of scale is achieved via the weighted PCA process, as resultant
variance of columns of Σ are 1. This has the advantage of normalizing data sets be-
longing to different multi-frequency ET signals, having different gain settings, and/or
acquired by different probes and instruments.

7.7.2 Practical Implementation

Prior to field use of PCA, for interpretation of multi-frequency eddy current gap
measurement data, consideration should be given to practical implications of field
implementation. Probe lift-off due to oxide and sharp pressure tube inner diameter
surface changes will induce impedance plane responses, which would also be required
as inputs into a fully inclusive PCA model for analysis of gap measurement data.
Other physical effects to consider include variable cable impedance for different cable
lengths and changing probe position, thermal signal drift, electrical noise, and prox-
imity to other reactor components outside the calandria tube (e.g. Liquid Injection
Shutdown System nozzles).
In implementing compressed data acquisition, information content contained in the
last few principal component scores will be lost upon recovery of the full dimensional
multi-frequency data set. Although it has been shown that responses to gap, wall
thickness, and resistivity are captured within the first three principal component
scores (Figure 7.2), generalization of compressed data acquisition to other applications
7.8. CONCLUSION 109

should verify that responses of particular physical phenomena of interest, which could
be present in the last few principal component scores, are not lost. To mitigate loss
of inspection data of interest, data used to generate principal components, should
include responses to relevant physical phenomena. Also, bench testing should be
employed to ensure a sufficient number of principal components are used in data
acquisition, for identification of physical phenomena.

7.8 Conclusion

Principal component analysis (PCA) has been performed on multi-frequency eddy


current data modeled via Dodd and Deeds equations [16], where PT resistivity, PT
wall thickness, and PT to CT gap have varied within the range expected in PT to
CT gap measurement.
∼ 97% of the original 6 dimensional data variance is represented by a single dimen-
sion (PCA score 1) post PCA analysis, demonstrating the high degree of correlation
across the multi-frequency data sets. Peak-to-peak variation of PCA score 1, is in-
versely proportional to increasing wall thickness and decreasing resistivity, consistent
with the skin depth equation.
Compressed data acquisition, demonstrating attainability of increased inspection
rate, can be realized by collecting the first m principal component scores representing
the significant majority of data variance. This compressed data acquistion has been
emulated for analytical model data, and it has been demonstrated how the originally
sensed multi-frequency eddy current data can be approximated by a linear transfor-
mation of the scores.
A comparison of PCA performed on experimental multi-frequency eddy current
7.8. CONCLUSION 110

data, and on eddy current data analytically derived from Dodd and Deed equations
[16] via computer program [84], was performed. Agreement was observed between
the PCA performed on analytical data and experimental data, with small differences
attributed to incomplete representation of the transmit-receive experimental data by
Dodd and Deeds equations.

Acknowledgment

The authors would like to thank Dr. Ross Underhill and Stuart Craig for valuable
discussions, as well as Ryan Howard from Ontario Power Generation for supporting
this work.
111

Chapter 8

Discussion

Traditional methods employed to examine eddy current measurement data have in-
volved examination of impedance plane diagrams, and more recently, C-Scan displays.
These methods have provided an intuitive and basic introduction for the interrogation
of eddy current data with minimal analysis; voltage responses to physical phenomena
such as lift off, resistivity changes, or the presence of an edge or defect have clear and
pre-defined responses for an operator to detect and interpret. Acquisition trajectory is
implicit in how voltage responses are displayed, as curves on impedance planes corre-
spond to voltage response to probe movement relative to a test piece, and C-Scan axes
correspond to probe location. The techniques developed in this thesis demonstrate
the effectiveness of particular analysis techniques applied to data that are independent
of acquisition location. These methods rely on representation of inspection data, in
the context of other data acquired either locally or globally. For example, the surface
profiling technique employs directional derivatives in multiple directions, relying on
neighbouring data to highlight areas of probe lift off. Furthermore, weighted prin-
cipal components analysis of multi-frequency eddy current data normalizes principal
component scores by their standard deviations [87]. This allows for normalization
112

of phase and scale for data sets acquired by different instruments and gain settings,
providing a common framework for representation of multi-frequency eddy current
data, and analyst review, as shown in the third manuscript provided in this thesis
(Chapter 7).
The surface profiling technique developed in this thesis employs a transmit-receive
probe, initially developed for pressure tube to calandria tube gap measurement. This
surface profiling technique has been employed at Ontario Power Generation for the
detection and characterization of near-surface pressure tube flaws, providing inde-
pendence (with respect to ultrasonic measurements) for detection and disposition of
pressure tube artefacts. Worthy to note is the dual functionality of the gap probe,
now employed for both relative eddy current measurement, towards gap estimation,
and pressure tube artefacts (including flaw) detection. Specialized engineered tools
having drastically different, yet critical functionalities, are rare, and development pro-
cesses for analysis tools employing the gap probe should be noted in evaluation of
other engineered tools in extending their functionality beyond their initially intended
purposes.
Recent advances in signal processing have introduced compressed sensing as a
methodology allowing for reducing traditional acquisition data density requirements
(per the Shannon-Nyquist theorem [88]), needed for full reconstruction. In simi-
lar fashion, compressed data acquisition, facilitated through calculation of principal
components of multi-frequency eddy current data, allows reduction of acquisition data
density requirements, while providing 99.9 % reconstruction capability, using only half
the normally acquired data, and still 97 % reconstruction capability using only one
principal component. Although different in their approach and requirements, these
113

two methods for reducing requirements for data acquisition can be employed where
the cost (whether due to cost of materials, or induced delays) of acquiring data is
prohibitive.
114

Chapter 9

Conclusions

The objective of this thesis was to develop a comprehensive characterization of eddy


current gap data in order to extract further information on fuel channel condition,
and to identify generalized applications for multi-frequency eddy current data. This
has been performed through development and commissioning of a surface profiling
tool, PCA analysis of gap data using validated Dodd and Deeds equations, and iden-
tification of a methodology for compressed data acquisition of multi-frequency eddy
current data.
Surface profiling of CANDU fuel channel pressure tube artefacts has been real-
ized through probe lift-off location extraction, reliant on a dot product computation
of complex valued variables (partial derivative of eddy current data, and the lift off
vector). Local pressure tube deformation due to spacer loading, constrictions, and
mechanical wear marks, having protrusions on the order of 200 µm, 50 µm - 100 µm,
and 50 µm - 100 µm, respectively, have been successfully imaged. Identification of
these artefacts reveals fitness for service condition of fuel channels, as deviant spacer
location and fretting are factors considered in assessing fuel channel degradation and
planning for replacement. The surface profiling tool has been independently validated,
115

using ultrasonic gauging measurements collected in reactor within fuel channels. Fi-
nally, this tool is generalizable to other conductive surfaces and eddy current probes.
The conditions for extending functionality beyond pressure tube applications, include
high eddy current data density and relatively large surface area of the probe, com-
pared to artefacts required to be imaged. As a result of work done towards this
thesis, the surface profiling tool is now a commissioned fuel channel inspection tool,
employed at Ontario Power Generation. It is used in tandem with ultrasonic mea-
surements to assess fitness for service of fuel channels. Analyst technicians have been
trained in its use, and the condition of numerous fuel channels has been evaluated
using the tool.
Contributing to a comprehensive characterization of gap eddy current data, inde-
pendent and redundant information across multi-frequency data sets has been iden-
tified through principal components analysis. Eddy current responses to PT to CT
gap, PT wall thickness, and PT resistivity have been modeled via Dodd and Deed
solutions to Maxwell’s equations. 97% of the data variance is represented by the first
principal component scores, while 99.7% (nearly 100%) is represented by the first
two principal component scores, reflecting a high degree of redundancy across multi-
frequency data sets. Minimum-to-maximum response of the first principal component
to gap variation, with varying PT wall thickness and resistivity is consistent with ex-
pected responses, considering the skin depth equation. The model data has been
validated against physical experiments, varying PT to CT gap, PT wall thickness,
and PT resistivity. Demonstration of compressed data acquisition for multi-frequency
gap data has identified a methodology, which is generalizable to potential applications
116

that acquire redundant information, and are limited due to acquisition hardware con-
straints. In the case of multi-frequency eddy current gap data, using half the number
of principal components (compared to the dimension of the original data set) allows
for acquisition and reconstruction fidelity to within 1%, for both mean and standard
deviation errors.
Finally, Dodd and Deeds solutions have been experimentally validated using sam-
ples with varying PT to CT gap, PT resistivity, and PT wall thickness variations.
This has facilitated the application of principal components analysis of eddy cur-
rent gap data using model data generated via Dodd and Deeds solutions. PCA of
model data was used to extract the significant components that would also arise
in experimental data due to variations in PT resistivity, wall thickness and PT to
CT gap. Employment of Dodd and Deeds solutions is generalizable to conductive
surfaces, whose intersection with eddy current probe magnetic field spread, can be
approximated as flat plates.
117

Chapter 10

Future Work

In this section the potential extension of the work of this thesis is introduced and
discussed. For purposes of discussion, future work can be grouped into two different
areas: implementation of generalized applications for multi-frequency eddy current
data, and development of mathematical tools for further characterization of the fuel
channel condition.
Generalized applications can be realized for the surface profiling methods, mod-
elling employing Dodd and Deeds equations, and principal components analysis of
multi-frequency eddy current data. Surface profiling of conductive surfaces can be
realized for numerous geometries, given eddy current data is collected with sufficient
data density, the direction of the lift-off vector is known, and the surface area of the
probe is larger than area of surface protrusions requiring imaging. Increased degrees
of freedom of probe movement (requiring probe redesign) will allow for ability to im-
age surface variations of increased geometrical variation. Furthermore, modelling of
factors affecting measured eddy current impedance employing Dodd and Deeds solu-
tions can be extended beyond the gap probe and fuel channel geometries, provided the
area in which the interaction between the magnetic field induced by the eddy current
118

probe, and conductive material can be approximated by a flat plate. Finally, in addi-
tion to employing principal components analysis (PCA) to highlight the relationship
between physical variables and high dimensional multi-frequency eddy current data,
PCA can be employed in the realization of compressed data acquisition. It has been
demonstrated how compressed data acquisition can increase the acquisition speed
of eddy current inspection given fixed limitations of data acquisition hardware. To
realize compressed data acquisition for eddy current systems, process hardware can
be employed to reduce the dimensionality of the data sensed by the probe for data
acquisition, while not overly restricting ability to reconstruct the originally sensed
data.
A number of mathematical tools could be exploited to further characterize the
fuel channel condition. PCA could be combined with techniques such as cluster anal-
ysis, support vector machines, or artificial neural networks to simultaneously extract
multi-parameter data, collected from in-reactor fuel channels. Reduction of data set
dimensionality via PCA can provide for tools exploiting geometric and topological
arguments, which are either not available or are only apparent at higher dimensions.
The potential for extraction of resistivity from pressure tube data, which could give
information on pressure tube condition in terms of irradiation induced effects of creep
on hardness and ductility and potentially hydrogen/deuterium pick-up, is a clear ex-
ample of an important candidate parameter for extraction. Specific to the problem
of PT resistivity extraction, as made evident from examination of the skin depth
equation, different excitation frequencies have variable sensitivity to resistivity, and
variation in this sensitivity may provide sufficient independence across measurements
to solve for resistivity. Towards extraction of new information identifying physical
119

characteristics of pressure tube data, consideration must be given for pre-requisite


field calibrations that must be performed.
Finally, an improvement in PT to CT gap estimation accuracy may be possible,
given compensation for variables such as base PT and CT resistivity and their respec-
tive temperatures at time of inspection. If these variables could be derived using the
current eddy current acquisition system, they would be candidate input variables into
a new and improved algorithm for gap estimation. Another example is compensation
for in-channel probe lift-off variation, made viable through field calibrations and gap
calculation algorithm advancement. An increase in gap estimation accuracy could
provide increased confidence in fuel channel fitness for service, providing sufficient
impetus for prolonged CANDU reactor operation.
BIBLIOGRAPHY 120

Bibliography

[1] “CANTEACH image library.” https://canteach.candu.org/Pages/


ImageLibrary.aspx. Accessed: 2016-04-15.

[2] M. Scholz, Approaches to analyse and interpret biological profile data. PhD thesis,
University of Potsdam, 2006.

[3] S. Shokralla and T. Krause, “Methods for evaluation of accuracy with multiple
essential parameters for eddy current measurement of pressure tube to calandria
tube gap in CANDU reactors,” CINDE Journal, vol. 35, no. 1, pp. 5–8, 2014.

[4] Canadian Standards Association, Mississauga, Ontario, Canada, N285.4-09 -


Periodic Inspection of CANDU Nuclear Power Plant Components, 2009.

[5] V. S. Cecco, G. V. Drunen, and F. L. Sharp, Eddy Current Testing Manual on


Eddy Current Method. Atomic Energy of Canada Limited, Chalk River, Ontario,
Canada, 1981.

[6] D. J. Griffiths, Introduction to Electrodynamics. Upper Saddle River, New Jersey:


Prentice Hall, third ed., 1999.
BIBLIOGRAPHY 121

[7] M. Trelinski, “Selected CANDU pressure tube degradation mechanisms and in-
spection related issues,” in 17th World Conference on Nondestructive Testing,
October 2008.

[8] M. Trelinski, “Inspection of CANDU reactor pressure tubes using ultrasonics,”


in 17th World Conference on Nondestructive Testing, October 2008.

[9] T. W. Krause, J. Schankula, and S. P. Sullivan, “Eddy current detection of spac-


ers in the fuel channels of CANDU nuclear reactors,” in Twenty-Third Annual
Conference of the Canadian Nuclear Society, (Toronto, Canada), June 2002.

[10] J. B. Marion and M. A. Heald, Classical Electromagnetic Radiation. Orlando,


Florida: Academic Press, second ed., 1980.

[11] I. T. Jolliffe, Principal Component Analysis. NY: Springer Series in Statistics,


second ed., 2002.

[12] J. M. Lattin, J. D. Carroll, and P. E. Green, Analyzing Multivariate Data.


Brooks/Cole-Thomson Learning, first ed., 2003.

[13] P. Horan, P. R. Underhill, and T. W. Krause, “Pulsed eddy current detection


of cracks in F/A-18 inner wing spar without wing skin removal using modified
principal component analysis,” NDT& E International, vol. 55, no. 0, pp. 21 –
27, 2013.

[14] “MultiScan MS5800 for tube inspection.” www.olympus-ims.com/en/


ms-5800-tube-inspection. Accessed: 2016-08-3.

[15] “Zetec TC7700 remote data acquisition unit.” zetec.ru/pdf/tc-7700.pdf. Ac-


cessed: 2016-08-3.
BIBLIOGRAPHY 122

[16] C. V. Dodd and W. E. Deeds, “Analytical solutions to eddy-current probe-coil


problems,” Journal of Applied Physics, vol. 39, no. 6, pp. 2829–2839, 1968.

[17] C. V. Dodd and W. E. Deeds, “Some eddy-current problems and their integral
solutions,” Tech. Rep. Contract No. W-7505-eng-26, Oak Ridge National Labo-
ratory, April 1969.

[18] C. C. Cheng, C. V. Dodd, and W. E. Deeds, “General analysis of probe coils near
stratified conductors,” International Journal of Nondestructive Testing, vol. 3,
pp. 109–130, 1971.

[19] J. A. Baron, “Qualification of inspection systems in the CANDU nuclear indus-


try,” CINDE Journal, vol. 35, no. 1, pp. 10–14, 2014.

[20] European Network for Inspection and Qualification, Luxembourg, ENIQ Recom-
mended Practice 6: The Use of Modelling in Inspection Qualification, 2nd ed.,
2011.

[21] P. R. Underhill and T. W. Krause, “Enhancing probability of detection and


analysis of bolt hole eddy current,” Journal of Nondestructive Evaluation, vol. 30,
pp. 237–245, 2011.

[22] R. Bachnak and S. King, “Non-destructive evaluation and flaw visualization


using an eddy current probe,” in Third International Conference on Systems,
(Cancun, Mexico), pp. 124–139, 2008.
BIBLIOGRAPHY 123

[23] R. P. R. Hasanzadeh, A. R. Moghaddamjoo, S. H. H. Sadaghi, A. H. Rezaie, and


M. Ahmadi, “Optimal signal-adaptive maximum likelihood filter for enhance-
ment of defects in eddy current c-scan images,” NDT&E International, vol. 41,
pp. 371–381, 2008.

[24] P. Xiang, Automatic multi-frequency rotating-probe eddy-current data analysis.


PhD thesis, Iowa State University, 2005.

[25] M. Bodruzzaman and S. Zein-Sabatto, “Estimation of micro-crack lengths using


eddy-current c-scan images and neural-wavelet transform,” in IEEE Southeast-
con, (Huntsville, AL, USA), pp. 551 – 556, April 2008.

[26] B. R. Groshong, G. L. Bilbro, and W. E. Snyder, “Eddy current image restoration


by constrained gradient descent,” Journal of Nondestructive Evaluation, vol. 10,
pp. 127–137, 1991.

[27] A. A. Diaz and R. A. Mathews, Assessment of Eddy Current Testing for the
Detection of Cracks in Cast Stainless Steel Reactor Piping Components. Pacific
Northwest National Laboratory, 2006.

[28] J. R. Buck, M. M. Daniel, and A. C. Singer, Computer Explorations in Signals


and Systems Using MATLAB. Upper Saddle River, NJ: Prentice Hall, second ed.,
2002.

[29] S. Damelin and W. Miller, The Mathematics of Signal Processing. Cambridge


University Press, 2011.

[30] M. R. Spiegel, S. Lipschutz, J. J. Schiller, and D. Spellman, Complex Variables.


New York, USA: McGraw-Hill, second ed., 2009.
BIBLIOGRAPHY 124

[31] J. Canny, “A computational approach to edge detection,” Pattern Analysis and


Machine Intelligence, IEEE Transactions on, vol. PAMI-8, no. 6, pp. 679–698,
1986.

[32] R. Gonzalez and R. Woods, Digital Image Processing. Upper Saddle River, New
Jersey, USA: Pearson Education, Inc., third ed., 2007.

[33] K. Pearson, “On lines and planes of closest fit to systems of points in space,”
Philisophical Magazine, vol. 2, no. 11, pp. 559–572, 1901.

[34] H. Hotelling, “Analysis of a complex of statistical variables into principal com-


ponents,” Journal of Educational Psychology, vol. 24, pp. 417–441, 1933.

[35] A. K. Helmy and G. S. El-Taweel, “Authentication scheme based on principal


component analysis for satellite images,” International Journal of Signal Pro-
cessing, Image Processing and Pattern Recognition, vol. 2, no. 3, 2009.

[36] K. I. Kim, K. Jung, and H. J. Kim, “Face recognition using kernel principal
component analysis,” IEEE Signal Processing Letters, vol. 9, February 2002.

[37] Y. H. Kim, S. J. Song, J. S. Hur, E. L. Kim, C. J. Yim, Y. H. Choi, S. C. Kang,


and M. H. Song, “Inversion of eddy current test signals obtained from steam
generator tubes,” Key Engineering Materials, vol. 297-300, pp. 2219–2224, 2005.

[38] A. Sophian, G. Y. Tian, D. Taylor, and J. Rudlin, “A feature extraction technique


based on principal component analysis for pulsed eddy current NDT,” NDT& E
International, vol. 36, no. 1, pp. 37 – 41, 2003.
BIBLIOGRAPHY 125

[39] V. Babbar, P. Underhill, C. Stott, and T. Krause, “Finite element modeling of


second layer crack detection in aircraft bolt holes with ferrous fasteners present,”
NDT & E International, vol. 65, no. 0, pp. 64 – 71, 2014.

[40] C. Stott, P. Underhill, V. Babbar, and T. Krause, “Pulsed eddy current detection
of cracks in multilayer aluminum lap joints,” Sensors Journal, IEEE, vol. 15,
pp. 956–962, Feb 2015.

[41] C. A. Stott, “Pulsed eddy current inspection of second layer wing structure,”
Master’s thesis, Royal Military College of Canada, 2014.

[42] Y. He, G. Tian, H. Zhang, M. Alamin, A. Simm, and P. Jackson, “Steel corrosion
characterization using pulsed eddy current systems,” Sensors Journal, IEEE,
vol. 12, pp. 2113–2120, June 2012.

[43] M. Pan, Y. He, G. Tian, D. Chen, and F. Luo, “PEC frequency band selection
for locating defects in two-layer aircraft structures with air gap variations,” In-
strumentation and Measurement, IEEE Transactions on, vol. 62, pp. 2849–2856,
Oct 2013.

[44] Y. He, F. Luo, M. Pan, F. Weng, X. Hu, J. Gao, and B. Liu, “Pulsed eddy
current technique for defect detection in aircraft riveted structures,” NDT & E
International, vol. 43, no. 2, pp. 176 – 181, 2010.

[45] Y. He, M. Pan, D. Chen, and F. Luo, “PEC defect automated classification
in aircraft multi-ply structures with interlayer gaps and lift-offs,” NDT & E
International, vol. 53, no. 0, pp. 39 – 46, 2013.
BIBLIOGRAPHY 126

[46] G. Yang, J. Kim, L. Udpa, and S. Udpa, “Analysis of pulsed eddy current - gmr
data using principal component analysis,” Studies in Applied Electromagnetics
and Mechanics, vol. 33, pp. 207–214, 2010.

[47] J. Kim, G. Yang, L. Udpa, and S. Udpa, “Classification of pulsed eddy current
gmr data on aircraft structures,” NDT&E International, vol. 43, no. 2, pp. 141–
144, 2010.

[48] P. Horan, R. Underhill, and T. W. Krause, “Pulsed eddy current detection of


cracks in F/A-18 inner wing spar at large lift-off using modified principal compo-
nent analysis,” International Journal of Applied Electromagnetics and Mechan-
ics, vol. 45, no. 1, pp. 287–292, 2014.

[49] P. Horan, P. Underhill, and T. Krause, “Real time pulsed eddy current detection
of cracks in F/A-18 inner wing spar using discriminant separation of modified
principal components analysis scores,” Sensors Journal, IEEE, vol. 14, pp. 171–
177, Jan 2014.

[50] G. Tian, A. Sophian, D. Taylor, and J. Rudlin, “Wavelet-based pca defect clas-
sification and quantification for pulsed eddy current ndt,” IEEE Proceeding -
Science, Measurement and Technology, vol. 152, no. 4, pp. 141–148, 2005.

[51] B. Gao, H. Zhang, W. L. Woo, G. Y. Tian, L. Bao, and A. Yin, “Smooth nonneg-
ative matrix factorization for defect detection using microwave nondestructuve
testing and evaluation,” IEEE Transactions on Instrumentation and Measure-
ment, vol. 63, no. 4, pp. 923–934, 2014.
BIBLIOGRAPHY 127

[52] B. Sasi, B. P. C. Rao, and T. Jayakumar, “Dual-frequency eddy current non-


destructive detection of fatigue cracks in compressor discs of aero engines,” De-
fence Science Journal, vol. 54, no. 4, pp. 563–570, 2004.

[53] R. La, B. Benoist, B. de Barmon, M. Talvard, R. Lengell, and P. Gaillard, “Mes-


sine, a parametric three-dimensional eddy current model,” Research in Nonde-
structive Evaluation, vol. 12, no. 2, pp. 65–86, 2000.

[54] C. Reboud, D. Premel, G. Pichenot, D. Lesselier, and B. Bisiaux, “Development


and validation of a 3d model dedicated to eddy current non-destructive testing of
tubes by encircling probes,” International Journal of Applied Electromagnetics
and Mechanics, vol. 25, no. 1-4, pp. 313–317, 2007.

[55] European Network for Inspection and Qualification, Luxembourg, ENIQ Recom-
mended Practice 1: Influential/Essential Parameters, 2nd ed., 2005.

[56] K. Cahill, Physical Mathematics. Cambridge: Cambridge University Press,


first ed., 2013.

[57] G. Tian, Z. Zhao, and R. Baines, “The research of inhomogeneity in eddy current
sensors,” Sensors and Actuators A: Physical, vol. 69, no. 2, pp. 148 – 151, 1998.

[58] B. G. Koronev, Bessel Functions and Their Applications. CRC Press, first ed.,
2002.

[59] J. Grman and L. Syrova, “Classifiability analysis of indications by eddy-current


testing of nuclear power plant heat-exchanger tube,” in Radioelektronika, 2009.
RADIOELEKTRONIKA ’09. 19th International Conference, pp. 287–290, 2009.
BIBLIOGRAPHY 128

[60] S. Tian, Z. Chen, M. Ueda, and T. Yamashita, “Signal processing schemes for
eddy current testing of steam generator tubes of nuclear power plants,” Nuclear
Engineering and Design, vol. 245, no. 0, pp. 78 – 88, 2012.

[61] S. T. Craig, T. W. Krause, B. V. Luloff, and J. J. Schankula, “Eddy current


measurement of remote tube positions in CANDU reactors,” in 16th World Con-
ference on Nondestructive Testing, (Montreal, Canada), August 2004.

[62] J. C. Aldrin and J. S. Knopp, “Crack characterization method with invariance


to noise features for eddy current inspection of fastener sites,” Journal of Non-
destructive Evaluation, vol. 25, December 2006.

[63] ASM International, Metals Park, OH, USA, Materials Handbook. Volume 17:
Nondestructive Evaluation and Quality Control. p. 168, ninth ed., 1989.

[64] M. Wrzuszczak and J. Wrzuszczak, “Eddy current flaw detection with neural
network applications,” Measurement, vol. 38, no. 2, pp. 132 – 136, 2005.

[65] G. Chen, A. Yamaguchi, and K. Miya, “A novel signal processing technique for
eddy-current testing of steam generator tubes,” Magnetics, IEEE Transactions
on, vol. 34, no. 3, pp. 642–648, 1998.

[66] J. Grman, R. Ravas, and L. Syrova, “Application of wavelet transformation in


eddy current testing of steam generator tubes,” in Instrumentation and Measure-
ment Technology Conference, 2001. IMTC 2001. Proceedings of the 18th IEEE,
vol. 1, pp. 392–396 vol.1, 2001.
BIBLIOGRAPHY 129

[67] E. G. Price, B. A. Cheadle, G. R. Evans, and M. W. Hardie, COG-91-65. Up-


date of Operating Experience with Cold-Worked Zr-2.5% Nb Pressure Tubes in
CANDU Reactors. Atomic Energy of Canada Limited, 1991.

[68] S. Shokralla, “Reducing the negative effects of grating through a-scan post pro-
cessing when compensating for poor element directivity,” Canadian Institute for
NDE, vol. 31, no. 4, pp. 8–11, 2010.

[69] R. Taneja, S. K. Parulkar, S. S. Taliyan, M. Singh, and G. Govindarajan, “Eddy


current measurement of annular gap between pressure tube and calandria tube
in indian pressurized heavy water reactors (PHWRs),” in 14th World Conference
on Non Destructive Testing, vol. 2, pp. 377–380, 1996.

[70] S. Shokralla, T. W. Krause, and J. Morelli, “Surface profiling with high density
eddy current non-destructive examination data,” NDT&E International, vol. 62,
pp. 153 – 159, March 2014.

[71] J. D. Jackson, Classical Electrodynamics, ch. 5, p. 220. New York, USA: Wiley
& Sons, third ed., 1999.

[72] V. S. Cecco, J. R. Carter, and S. P. Sullivan, “An eddy current technique for
detecting and sizing surface cracks in carbon steel,” Materials Evaluation, vol. 51,
pp. 572–577, May 1993.

[73] L. S. Obrutsky, V. S. Cecco, S. P. Sullivan, and D. Humphrey, “Transmit-receive


eddy current probes for circumferential cracks in heat exchanger tubes,” Mate-
rials Evaluation, vol. 54, pp. 93–98, January 1996.
BIBLIOGRAPHY 130

[74] E. G. Price, TDVI-368. Thermal Conductivity, Electrical Resistivity and Specific


Heat of CANDU Contructional Zirconium Alloys and AISI Type 403 End Fitting.
Atomic Energy of Canada Limited, 1980.

[75] C. de Boor, A Practical Guide to Splines. New York, USA: Springer-Verlag,


first ed., 1978.

[76] D. Horn and R. Roiha, “Multifrequency analysis of eddy current data,” in 16th
World Conference on NDT, August 2004.

[77] S. Shokralla, S. Sullivan, J. Morelli, and T. W. Krause, “Modelling and validation


of eddy current response to changes in factors affecting pressure tube to calandria
tube gap measurement,” NDT&E International, vol. 73, no. 0, pp. 15 – 21, 2015.

[78] M. Cacciola, G. Ripepi, G. Yang, G. Y. Tian, and F. C. Morabito, “ICA based


algorithms for flaw classification in pulsed eddy current data: A study,” in Neural
Nets WIRN10 - Proceedings of the 20th Italian Workshop on Neural Nets, Vietri
sul Mare, Salerno, Italy, May 27-29 2010, pp. 162–171, 2010.

[79] L. Cheng, B. Gao, G. Y. Tian, W. L. Woo, and G. Berthiau, “Impact damage


detection and identification using eddy current pulsed thermography through
integration of PCA and ICA,” Sensors Journal, IEEE, vol. 14, pp. 1655–1663,
May 2014.

[80] L. Obrutsky, J. Renaud, and R. Lakhan, “Overview of steam generator tube-


inspection technology,” in NDT in Canada, 2009.

[81] M. O’Connor, “Multifrequency eddy current instrumentation: Understanding


the specifications,” Materials Evaluation, vol. 64, no. 4, pp. 391–393, 2006.
BIBLIOGRAPHY 131

[82] L. Obrutsky, J. Renaud, and R. Lakhan, “Steam generator inspections: Faster,


cheaper and better, are we there yet?,” in IV Conferencia Panamericana de
END, 2007.

[83] M. O’Connor, “Design of an eddy-current array probe for crack sizing in steam
generator tubes,” NDT& E International, vol. 36, no. 7, pp. 515 – 522, 2003.

[84] S. T. Craig, “Surfacecalc evaluation of pt/ct gap eddy-current signals under a


range of resistivity and wall-thickness conditions,” Report 153-128840-REPT-003
Rev. 0, IMD AECL, May 2013.

[85] “Mathworks pca.” http://www.mathworks.com/help/stats/pca.html. Ac-


cessed: 2015-07-26.

[86] F. Jovic, Process Control Systems Principles of design, operation and interfacing.
Springer Netherlands, first ed., 1992.

[87] Z. Fan, E. Liu, and B. Xu, Weighted Principal Component Analysis, pp. 569–574.
Berlin, Heidelberg: Springer Berlin Heidelberg, 2011.

[88] A. J. Jerri, “The shannon sampling theorem;its various extensions and applica-
tions: A tutorial review,” Proceedings of the IEEE, vol. 65, pp. 1565–1596, Nov
1977.

Вам также может понравиться