This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.
Paper The following article is Open access

Microtexture region segmentation of eddy current testing data using a structural prior

, , and

Published 2 April 2024 © 2024 The Author(s). Published by IOP Publishing Ltd
, , Special Issue on Inverse Problems in Civil, Mechanical and Aerospace Engineering Citation Laura Homa et al 2024 Inverse Problems 40 055013 DOI 10.1088/1361-6420/ad366e

0266-5611/40/5/055013

Abstract

Microtexture regions (MTRs) are collections of grains with similar crystallographic orientation. Because their presence in titanium alloys can significantly impact aerospace component life, a nondestructive method to detect and characterize MTR is needed. In this work, we propose to use data from two nondestructive evaluation methods, eddy current testing (ECT) and scanning acoustic microscopy (SAM), in order to recover the boundary and dominant crystallographic orientation of each MTR in a specimen. ECT is an electromagnetic method that is sensitive to changes in crystallographic orientation associated with MTR; however, its low resolution prevents it from resolving MTR boundaries well. In contrast, SAM is a high frequency ultrasound method that is able to resolve MTR boundaries but is not sensitive to orientation. This paper proposes an algorithm to characterize MTR that makes use of a method known as covariance generalized matching component analysis. This method is used to build a surrogate linear forward model that relates MTR boundaries and orientation to ECT data. The model is inverted using the SAM data as a structural prior. We demonstrate this technique using simulated ECT and experimental SAM data from a large grain titanium specimen.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 license. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Titanium (Ti) alloys see wide use in the aerospace industry. Certain Ti alloys have two ductile microstructural phases–hexagonal close packed (HCP) α and body centered cubic (BCC) β–with the α-phase dominating the microstructure by volume fraction. In these alloys, α-phase grains can aggregate into regions of similar crystallographic orientation to produce microtexture regions (MTRs) [1]. MTRs can localize strain, leading to crack nucleation, extension, and significant reduction in the lifespan of aerospace parts [2, 3]. As such, a method to characterize MTR is necessary for quality control.

Several inspection methods are sensitive to the size and/or orientation of MTR, each with trade-offs in resolution, acquisition time, and cost. These techniques include scanning electron microscopy electron backscatter diffraction (SEM EBSD), polarized light microscopy (PLM), scanning acoustic microscopy (SAM), and eddy current testing (ECT).

SEM EBSD and PLM are microscopy techniques for microstructure characterization. EBSD determines crystallographic orientation by measuring the diffraction pattern produced by electrons interacting with the target's crystal lattice [4]. Conversely, PLM exploits the optical anisotropy of metals like α-Ti with polarized light to measure crystallographic orientation. Of these, SEM EBSD is the gold standard for MTR characterization, providing information on the size and orientation of surface grains and their aggregation into MTR. However, both SEM EBSD and PLM have practical limitations for characterization of MTR in real structures: most notably, a requirement for highly-polished flat surfaces.

SAM is high frequency ultrasonic testing that is sensitive to small changes in the complex modulus of the target material. Certain designs of a SAM transducer can produce Rayleigh surface waves within the target that are ideal for characterizing surface-breaking MTR. Typical SAM transducers can image features smaller than 100 µm in these materials, and the velocity of the propagating surface wave directly correlates to MTR stiffness as a function of orientation. However, the spherically-focused lens found on typical RSW SAM transducers results in non-unique orientation information.

ECT is an electromagnetic technique that uses alternating current passing through a coil of wire to generate a primary magnetic field. When coupled to a conductive body, this magnetic field creates a localized electric field and secondary magnetic field. The secondary magnetic field opposes the primary magnetic field and produces a change in the impedance of the coil as a function of the conductivity of the underlying structure. The local conductivity of the Ti alloy depends on the orientation of the microstructure and its aggregation into MTR; thus, ECT is sensitive to MTR orientation. In particular, it has been shown ECT is sensitive to changes in crystallographic orientation that alter MTR conductivity by as little as 6% [58]. While previous work has demonstrated the use of ECT for material property characterization (see, for instance, [911]), the main limitation of ECT is spatial resolution: features smaller than 350 µm can only be detected in isotropic, homogeneous materials [12]. Therefore, ECT is sensitive to MTR orientation but struggles to resolve MTR boundaries and size.

SAM and ECT are appealing choices because they can both be performed in a practical environment on complex geometries with as-machined surfaces. Ultimately, we want to use these modalities to produce a spatial map of MTR in a specimen, similar to what is shown in figure 1. Figure 1 depicts a set of EBSD data of a real titanium alloy specimen and the corresponding MTR segmentation computed using the open-source software DREAM.3D [13]. Each MTR is colored based on its average crystallographic orientation. The goal of this paper is to propose a method in which the boundary information from SAM can be combined with orientation information from ECT to produce a similar result to figure 1(c).

Figure 1.

Figure 1. Sample EBSD data and its corresponding MTR segmentation. (a) The legend for the images. (b) The EBSD data. Each point is colored based on the crystallographic orientation of the specimen at that point. (c) The MTR segmentation of the specimen. Each MTR is colored based on the average crystallographic orientation.

Standard image High-resolution image

The idea of combining imaging modalities that convey different information but have varying levels of spatial resolution has been used widely in the medical field. For instance, a variety of papers have suggested the use of anatomical information from magnetic resonance imaging (MRI) to improve the reconstruction of positron emission tomography (PET) data (see, for instance, [1417]). Other work has focused on using computed tomography (CT) or MRI images to inform electrical impedance tomography (EIT) inversion (see [1821]) as well as diffuse optical tomography (DOT) inversion [22]. The goal of these works is to use a high-resolution method to build a structural prior that can be used to inform the inversion of a lower-resolution method that contains different information. However, in spite of their wide use in the medical field, structural priors have not been applied to nondestructive evaluation methods. While there is a history of data fusion in nondestructive evaluation (see, for instance, [23, 24]), the proposed methods focus on damage detection and characterization, rather than material characterization. Work that does focus on material characterization, such as that in [25, 26], does not consider any kind of data inversion. Thus, we feel there is an opportunity to translate inversion techniques typically used in medical imaging to nondestructive evaluation.

In this work, this is achieved using a method known as covariance generalized matching component analysis (CGMCA). The purpose of CGMCA is to find common features between two types of data by mapping each to a common domain. We use CGMCA to build a surrogate linear forward model relating MTR boundaries and orientation to the corresponding ECT signal. We train the CGMCA maps on the MTR segmentation of EBSD data matched to the corresponding ECT data of that microstructure. The MTR segmentation of a new set of ECT data is found by mapping the ECT data to the common domain, and then mapping the common domain data back to the segmentation domain. The general outline of this method is shown in figure 4. The SAM data is incorporated when the common domain data is mapped back to the segmentation domain, which requires the inversion of one of the CGMCA maps. In particular, the SAM data is used to construct a structural prior that is applied in the inversion back to the segmentation domain. We demonstrate the method on experimental SAM and simulated ECT data of a large grain titanium specimen.

The paper is laid out as follows. Section 2 provides a description of the test specimen and data types, section 3 describes the analysis methods, section 4 describes the implementation of the method, section 5 shows the results, and section 6 provides conclusions.

1.1. Note on crystallographic orientation

Crystallographic orientation refers to the set of rotations that generates the orientation of a grain with respect to a global frame of reference. In this work, it is described in terms of three Euler angles $(\psi_1,\theta, \psi_2)$: ψ1 is rotation about the z-axis, θ the subsequent rotation about the new x-axis, and ψ2 the final rotation about the new z-axis [27]. Generally, $\psi_1 \in [0, 2\pi]$, $\theta \in [0, \pi]$, and $\psi_2 \in [0, 2\pi]$. However, symmetry of the HCP crystal permits rotations to be expressed in the ranges $\psi_1 \in [0, 2\pi]$, $\theta \in [0, \pi/2]$, and $\psi_2 \in [0, \pi/3]$. Note that the axis of the HCP crystal parallel to the z-axis is called the c-axis, and the plane perpendicular to the z-axis is called the basal plane. A diagram of the unit cell for grains with hexagonal symmetry depicted in the three rotations that define crystallographic orientation is shown in figure 2.

Figure 2.

Figure 2. The unit cell for grains with hexagonal symmetry, shown in the three rotations that define crystallographic orientation.

Standard image High-resolution image

2. Experimental and simulated data

The methods proposed here will be demonstrated on experimental SAM and simulated ECT data of a large grain titanium alloy specimen. The PLM data of this specimen is shown in figure 3. Each point in the image is colored based on the crystallographic orientation of the specimen at that location. The sample specimen shown in figure 1 consists of grains with sizes on the order of 10 µm that aggregate into MTR. In contrast, the grains in the specimen shown in figure 3 have sizes on the order of several millimeters. Each solid colored region in figure 3(a) represents a single grain with a single orientation, instead of a collection of grains with similar orientation. This is a simpler case than a specimen with MTR, but it serves as a proof of concept for the methods we demonstrate here.

Figure 3.

Figure 3. The large grain Ti sample as viewed by three different modalities. (a) Polarized light microscopy, (b) scanning acoustic microscopy, and (c), (d) eddy current testing. The eddy current testing data is shown for two rotations of the ECT coil, specifically (c) 0 and (d) 90.

Standard image High-resolution image

2.1. Scanning acoustic microscopy

Ultrasound measurements are performed at spatial locations, in a grid pattern with 50 µm spacing, across the sample with a single element ultrasonic transducer. Both the sample and transducer are immersed in water to ensure proper coupling of acoustic energy. The transducer is an Olympus V3330 with a nominal frequency of 50 MHz, 6.35 mm diameter element, and a 5.08 mm focal length. The transducer also includes an integrated fused silica delay line that allows the ultrasonic wave to form a uniform wave front prior to focusing. The transducer is de-focused to slightly shift the time-of-flight of the reflected longitudinal wave from that of the Rayleigh surface wave. This is required to facilitate data analysis. Time-of-flight refers to the time required for the generated sound wave to propagate to and interact with the sample and then reflect back to the transducer to be measured.

A digitized voltage, directly proportional to the acoustic pressure across the entire transducer element, is recorded at each spatial location as a function of time. This is commonly referred to as an A-scan. A time gate is defined—in this case, a time window within which we expect the Rayleigh surface wave to propagate back to the transducer element. This time gate is allowed to shift in time as we scan across the sample by tracking the time-of-flight of the higher amplitude reflected longitudinal wave in a separate tracking gate. This accounts for any subtle local variations in sample geometry.

The maximum value of the voltage within the time gate for the Rayleigh surface wave is taken from the A-scan captured at each spatial location and used to produce the image shown in figure 3(b). The color scale in this image shows the measured maximum voltage within the gate, which is directly proportional to the amplitude of the Rayleigh surface wave, mapped across the surface of the sample. This form of presentation is commonly referred to as a C-scan.

The amplitude of this back-scattered Rayleigh surface wave has been found previously to be sensitive to grain structure, though the exact mechanisms are not fully understood [28]. Furthermore, as previously mentioned, since this is a single element transducer with a spherically focused beam, we are simultaneously measuring the back-scattered response as the wave propagates along the surface radially in all directions from the point of focus and therefore we do not obtain useful orientation information.

2.2. Eddy current forward model

All ECT data used in this paper are simulated. The forward model used to simulate ECT data is the approximate impedance integral (AII) model (see [7]), which is based on Maxwell's equations. We briefly review its formulation here. In differential form, Ampere's Law and Faraday's Law are given by

Equation (1)

Equation (2)

where ω is the angular frequency, µ is the permeability, ε is the scalar permittivity, and $\overline{\sigma}$ is the anisotropic conductivity tensor. We assume there are no external fields and that the only current sources are the induced currents in the conductive material and the displacement currents. From [29], (1) and (2) can be combined to yield the expression

Equation (3)

where S is the surface enclosing a flaw in the conductive medium and V is the volume of the flaw. For our application, S is the entire surface of the sample. From [30], this expression can be related to the impedance of an ECT coil to yield

Equation (4)

where all fields with the subscript a are the result of exciting the ECT coil above an isotropic homogeneous material with conductivity σa , while all fields with the b subscript are the result of exciting the coil above an anisotropic, polycrystalline material. Due to the relatively low conductivity changes in a rotated grain, the Born approximation can be applied to (4). Under this approximation, we have that $\mathbf{H}_a \approx \mathbf{H}_b$ and $\mathbf{E}_a \approx \mathbf{E}_b$, in which case (4) becomes

Equation (5)

The rotated, anisotropic conductivity tensor $\overline{\sigma}_b$ is a function of the local crystallographic orientation and is given by

Equation (6)

where

Equation (7)

and ci and si are the sine and cosine of the ith Euler angle, respectively. Within α-titanium, conductivity differs along the c-axis but is isotropic in the basal plane. ECT is sensitive to changes in the first two Euler angles because they correspond to changes in the orientation of the c-axis. However, it is not sensitive to changes in the third Euler angle due to the fact that it is simply a final rotation in the basal plane. The fact that $\overline{\sigma}_b$ in the final model is dependent on only ψ1 and θ is consistent with the fact that ECT is not sensitive to changes in the third Euler angle within α-titanium. For further details on this model, we refer to ([7]).

Figures 3(c) and (d) show the imaginary portion of the simulated ECT data of the large grain specimen for two rotations of an elliptical ECT coil with axes equal to 200 µm and 635 µm. The simulation was a two-step process. First, the reference electric field Ea from (5) was computed using COMSOL Multiphysics. Specifically, a linear FEM simulation was used to compute the reference electric field values at the surface of an isotropic, homogeneous half-space. Second-order triangular elements with size equal to one-third of the skin depth were used with Lagrange interpolation. Given the field Ea , the change in impedance for a fixed location of the coil was computed by solving (5) using numerical integration in MATLAB. The conductivity tensor $\overline{\sigma_b}$ was computed at each point in the specimen using the orientations provided by the PLM data shown in figure 3(a). The spatial resolution of the PLM data was 15 µm, which is a sufficiently small step size to ensure the numerical integration converges. Note that the integration is performed over the span of the coil field, which is a 2 mm square centered at the coil location. The full scan of the specimen is performed by moving the center of the simulated coil 100 µm, interpolating Ea at the new location, and recomputing (5) over the new span of the coil.

Because this model only requires a single solve of an FEM model to obtain Ea , it is computationally efficient. A full raster scan of the specimen can be performed without needing to solve an FEM model at each measurement point. On an Intel Core i9 desktop with a 2.4 GHz CPU the simulation took around 230 s. Note that the full PLM dataset is spatially larger than the portion shown in figure 3(a). As such, we did not need to be concerned about edge effects in our simulation.

We chose to use simulated ECT data due to the fact that the method requires the ECT data and SAM data to be registered to each other, which is non-trivial. In this paper, we registered the SAM data to the PLM data using a landmark registration algorithm built into MATLAB. By simulating ECT data from the PLM data, we could ensure that it would be registered to the SAM data. The development of an algorithm to register SAM and ECT data is a topic of ongoing work, but is not the topic of this paper. While we did choose to use simulated ECT data, we note that the AII model was validated against experimental ECT data in [7]. Thus, we feel it is appropriate to use here.

Lastly, we use an elliptical absolute ECT coil which exhibits directional sensitivity to the underlying crystallographic orientation. The elliptical coil is necessary to recover information about both the first and second Euler angle. This is opposed to using a circular coil, which is only sensitive to the second Euler angle. We simulated ECT data from two coil rotations, both of which will be used in the inversion. As will be demonstrated in section 5, we require data from multiple rotations of the ECT coil to obtain unique orientation information.

3. Numerical methods

3.1. Segmentation using matching component analysis

Matching component analysis (MCA) is a data analysis method which is used to determine common features between two data domains by mapping each to a lower dimensional common domain [31]. It was originally developed for transfer learning; one of its initial applications was to find a common domain between experimental and simulated synthetic aperture radar (SAR) data to improve classification. We use an extension of MCA known as covariance generalized matching component analysis (CGMCA) [32]. For this application, the two data domains of interest are (1) the underlying segmentation of a known microstructure into MTR and (2) the corresponding ECT data of the microstructure.

Let $X_1 \in \mathbb{R}^{d_1}$ and $X_2 \in \mathbb{R}^{d_2}$ be random variables modeling the two data domains. Then CGMCA seeks to find maps g1 and g2 that are solutions to the constrained optimization problem

Equation (8)

Equation (9)

Equation (10)

Equation (11)

where $k \unicode{x2A7D} \min(d_1, d_2)$ is the dimension of the common domain, and the symmetric positive definite matrix $\Gamma_i \in \mathbb{R}^{k \times k}$ is the covariance matrix prescribed to the map gi . It was shown in [32] that if we require that

Equation (12)

for $A_i \in \mathbb{R}^{k \times d_i}$ and $b \in \mathbb{R}^k$, then a closed form solution exists to the CGMCA optimization problem. Note that if the choice of $\Gamma_i$ in (11) is restricted to the identity matrix, this reduces to the original MCA problem. The algorithm to compute the CGMCA maps is provided in appendix.

The basic procedure for segmentation using CGMCA was developed in [33]. The CGMCA maps g1 and g2 are trained on a set of known microstructures with known MTR segmentation matched to the corresponding ECT data of each microstructure, as shown in figure 4. Given a new set of ECT data y with unknown segmentation, we first map y to the common domain using g2. We then recover the unknown segmentation by inverting g1 to map back to the segmentation domain. Formally, we wish to find the unknown segmentation $s^*$ such that

Equation (13)

Recall that g1 and g2 are affine linear transformations; thus we can find $s^*$ by solving the linear system

Equation (14)

Note that $A_1 \in \mathbb{R}^{k \times d_1}$, with $k \lt d_1$; as such, this system is underdetermined. A possible solution is given by

Equation (15)

However, as was shown in [33], this solution will not behave as a segmentation should, i.e. it will not be spatially piecewise constant. Thus, additional constraints on the solution are needed.

Figure 4.

Figure 4. Basic outline of CGMCA segmentation process.

Standard image High-resolution image

In [33], the piecewise constant requirement was enforced by utilizing level set inversion to find $s^*$, while in [34], a penalty term was added to the inversion. Specifically, in [34], a regularization method developed for image deblurring in [35] was applied, which we review briefly in the next subsection. This paper proposes a new method to impose the piecewise constant requirement using the SAM data as structural information in the inversion, as will be discussed later in the section.

3.2. Image deblurring prior

Rather than estimate s directly, the method proposed in [35] estimates the increments (i.e. the difference between neighboring pixels) of s. Denote the vector containing the horizontal and vertical increments of s by α. This method assumes that the entries of α are mutually independent and follow a conditionally Gaussian prior. The variance of each increment is also an unknown and is estimated as part of the inverse problem. More specifically, the method estimates the reciprocal of the variance of each increment, which we denote by λ. In [35], an exponential hyperprior is applied to each entry of λ to encode the idea that the majority of the increments will be close to zero, while still allowing for jumps when needed.

We briefly review the implementation of the image deblurring prior here; for further detail, we direct the reader to [35]. The biggest challenge in implementing this prior is writing the forward model in terms of α. Let Ne denote the number of increments in s. Note that we assume in the following calculations that $s \in \mathbb{R}^{N \times N}$ has been stacked into a single column vector of length N2. We can compute α from s,

Equation (16)

where $B \in \mathbb{R}^{N_e \times N^2}$ is a difference operator that computes both the horizontal and vertical increments of s. Because $N_e \gt N^2$, we cannot invert B to obtain s from α. It can be shown that $\text{rank}(B) = N^2 - 1$. Denote by $B_2 \in \mathbb{R}^{N^2-1 \times N^2}$ the matrix that contains a set of linearly independent rows of B, and let $\alpha_2 \in \mathbb{R}^{N^2-1}$ denote the vector that contains the corresponding edges.

The fact that the number of linearly independent edges is equal to $N^2-1$ agrees with the fact that an image cannot be reconstructed from increments alone; therefore, we must apply another constraint. Let $v \in \mathbb{R}^{N^2}$ be a vector with non-zero entries and let ν be the projection of $s$ onto $v$, that is, $\nu = v^Ts$. Define the matrix $C \in \mathbb{R}^{N^2 \times N^2}$ as $C = [B_2^T \; \; v]^T$, and then we can write

Equation (17)

Because C is full rank, it can be inverted to find $s$ given the increments α 2 and ν. This allows us to write the forward model in terms of the increments.

Next, define the augmented vector $\xi \in \mathbb{R}^{N_e + 1}$ and the permutation matrix $P \in \mathbb{R}^{N^2 \times N_e + 1}$ as

Equation (18)

Now we can write an equation to recover $s$ from ξ , specifically

Equation (19)

In the inverse problem, there is an additional constraint placed on α ; specifically, that the four increments surrounding each interior point in $s$ sum to 0. This allows us to solve for the remaining entries of α in terms of α 2 . (See section 4 in [35] for more details.) This condition is enforced by the matrix $M \in \mathbb{R}^{(N-1)^2 \times N_e+1}$, so that we require $M\xi = 0$. Lastly, define the matrix D as

Equation (20)

Our forward model in terms of ξ is given by

Equation (21)

We have a Gaussian error term in our forward model to signify that the segmentation will not map exactly to the ECT data in the common domain. Additionally, this term expresses the fact that we require only that $M\xi \approx 0$, instead of $M\xi = 0$. We could have two different error terms for our CGMCA forward model and the additional constraint; however, for the sake of simplicity, we combine it into one.

We assume that the entries of the unknown ξ are mutually independent and follow a conditionally Gaussian prior. Furthermore, each entry of the reciprocal of the variance λ follows an exponential hyperprior. The optimal values of ξ and λ are given by the maximum a posteriori (MAP) estimates, which are solutions to the minimization problem

Equation (22)

The entries of the diagonal matrix $\Lambda \in \mathbb{R}^{N_e + 1 \times N_e+1}$ are given by $\sqrt{\lambda}$ and β is the reciprocal of the mean of the exponential hyperprior.

The procedure to estimate ξ and λ is iterative, and requires alternating between optimizing ξ and λ . The procedure is as follows:

  • (i)  
    Let h = 0. Initialize λh .
  • (ii)  
    Update ξ by minimizing (22) with respect to ξ ; specifically, find ξ h by solving
    Equation (23)
    in the least squares sense. Note that we first multiplied (22) through by γ2 in order to obtain (23).
  • (iii)  
    Update λ by minimizing (22) with respect to λ . The update for λ is explicit, and is given by
    Equation (24)
  • (iv)  
    Increase h to h + 1 and repeat from step (ii) until convergence.

In this work, we run this algorithm for a fixed number of iterations. Then $s^* = C^{-1}P\xi^*$, where $\xi^*$ is the optimal value of ξ as found by the iterative procedure.

3.3. Structural prior

In this work, we propose applying a penalty term that incorporates structural information from the SAM data. We discuss the implementation of our method here. Denote the $\ell$th entry of the unknown discretized segmentation $s \in \mathbb{R}^{N \times N}$ as $s_\ell$ for $1 \unicode{x2A7D} \ell \unicode{x2A7D} N^2$ and denote the set of indices of the neighboring pixels to $s_\ell$ as $n_\ell$. We assume that $s_\ell$ is roughly equal to the weighted average of its neighbors, that is,

Equation (25)

where $\omega_\ell^k$ is the kth entry of the weight vector $\omega_\ell \in \mathbb{R}^{1 \times N^2}$. Note that the weight vector differs for each entry $s_\ell$.

In our application, we extract the MTR boundaries from the SAM data. This edge information is contained in the matrix e, which is the same size as the unknown segmentation s. The $\ell$th entry of e is defined as follows,

Equation (26)

Given the edges, the weight vector is formed in the following way. If $k \notin n_\ell$, then $\omega_\ell^k = 0$. Otherwise, $\omega_\ell^k$ is given by

Equation (27)

where $0 \lt \epsilon \lt 1$ is the coupling constant. The smaller the value of ε, the more neighboring pixels that lie on an edge are allowed to differ from their non-edge neighbors. This construction results in a spatially piecewise constant solution; however, unlike our previous methods which attempted to directly estimate MTR boundaries from ECT data, this approach assumes that we have prior knowledge about the boundaries from the SAM data.

Define the $\ell th$ row of the matrix $R_\epsilon \in \mathbb{R}^{N^2 \times N^2}$ by

Equation (28)

Then the full regularization operator $L_\epsilon \in \mathbb{R}^{N^2 \times N^2}$ for the structural prior is given by

Equation (29)

where I is the $N^2 \times N^2$ identity matrix. Using this regularization method, the value of $s^*$ is given by

Equation (30)

where ρ is the regularization parameter. This implies that $s^*$ can be found by solving the linear system

Equation (31)

in the least squares sense. Note that unlike the method described in the previous subsection, this technique does not require an iterative process to find $s^*$.

4. Numerical implementation

4.1. Edge detection on SAM data

We used a standard canny edge detection algorithm built into MATLAB to detect grain boundaries from the SAM data shown in figure 3(b). Figure 5 shows the detected edges overlaid on top of maps of the first and second Euler angles. Note that although SAM cannot provide unique orientation information, it is sensitive to changes in both the first and second Euler angles. The grain boundaries as determined by the SAM data match up very well with the true grain boundaries from the specimen. These edges are used to build our structural prior when we perform inversion on the ECT data.

Figure 5.

Figure 5. Edges as detected from SAM data, overlaid on (a) map of the second Euler angle and (b) map of the first Euler angle. The black lines represent edges. The colorbars are both in radians, ranging from 0 to $\pi/2$.

Standard image High-resolution image

4.2. Training of CGMCA maps

We generated a set of $n = 18\,000$ simulated microstructures and simulated ECT data sets to train the CGMCA maps. Each microstructure is 1 mm × 1 mm in size and consists of two regions, with each region defined by a single crystallographic orientation. A sample microstructure is shown in figure 6. Each pixel in the simulated microstructure represents an area of $25 \; \mu \text{m} \times 25 \; \mu\text{m}$. For each microstructure, we simulated ECT data using an elliptical coil with axes equal to 200 µm and 635 µm. The assumed liftoff of the probe was 90 µm. The simulated data was recorded with a step size of 100 µm. The ECT data collection was simulated for two rotations of the elliptical coil (0 and 90); the imaginary portion of simulated ECT data for a sample simulated microstructure for both rotations is also shown in figure 6.

Figure 6.

Figure 6. A sample set of training data for the CGMCA maps. (a) The training segmentation, labeled based on the value of the second Euler angle in each region. (b), (c) Imaginary portion of corresponding simulated ECT data for two rotations of the ECT coil, specifically (b) 0 and (c) 90.

Standard image High-resolution image

Our first test was to segment the data along the second Euler angle. Thus, the regions in each training segmentation were labeled based on the value of the second Euler angle for that region. Note that each training segmentation was downsampled from 25 µm step size to 50 µm in order to match the resolution of the SAM data.

We trained the CGMCA maps on the segmentations matched to the simulated ECT data sets. Let Z1 and Z2 denote the matrices that contain the training data for the segmentation and ECT data, respectively. The dimensions of a single training segmentation are N×N, where N = 21. Each column of $Z_1 \in \mathbb{R}^{N^2 \times n}$ contains one of the training segmentations stacked into a single vector. A single set of ECT measurements is of the size M×M, where M = 11. Each column of $Z_2 \in \mathbb{R}^{4M^2 \times n}$ contains the real and imaginary portion of the ECT signal from each rotation for the corresponding segmentation in Z1 stacked into a single column vector.

We chose the matrices Γ1 and Γ2 for each CGMCA map in the same manner as in [33]. Specifically, we assume that $\Gamma_1 = \Gamma_2$ and assign a single covariance matrix $\Gamma \in \mathbb{R}^{k \times k}$ to the common domain. Let $\mu_1 \in \mathbb{R}^{N^2}$ be the mean vector of the columns of Z1. Denote the matrix containing the centered segmentation data sets as S1, where S1 is given by

Equation (32)

Then denote the SVD of S1 by

Equation (33)

and denote the projection of S1 onto the first k columns of S1 as $P_1 \in \mathbb{R}^{k \times n}$. The covariance matrix Γ is given by

Equation (34)

where $p_\ell$ is the $\ell$th column of P1 and $\overline{p}$ is the mean vector of the columns of P1. This choice of Γ was shown to work well for in the segmentation algorithm in [33].

The dimension of the common domain k is chosen based on the singular values of S1, in a similar manner as in [33]. Specifically, k is chosen as the index where the singular values of S1 begin to level off. Here, we let k = 50.

4.3. Processing a large ECT data set

Note that our training ECT data sets have dimensions 1 mm × 1 mm. However, the ECT data of the large grain Ti specimen represents an area of 33.4 mm × 18 mm. Thus, we need a way to process a larger data set through the segmentation algorithm. We achieve this by processing the data in 1 mm × 1 mm squares. The process is outlined below.

  • (i)  
    Extract a 1 mm × 1 mm square of data from the large ECT data set to form y.
  • (ii)  
    Map y to the common domain, i.e. compute $A_2y + b_2$.
  • (iii)  
    Find $s^*$ that satisfies
    Equation (35)
    while also satisfying the constraints imposed by either the image deblurring prior or the structural prior.
  • (iv)  
    Repeat for another extracted data set that partially overlaps the previous one.
  • (v)  
    Repeat steps (i)–(iv) until the entire data set is processed.

Once each of the ECT data subsets have been processed, we need to combine them into a single final segmentation, which we denote by $S$. Let $x_\ell$ denote the $\ell$th spatial location in S. After completing the process outlined above, suppose there are a total of $m_\ell$ smaller segmentations that cover $x_\ell$. Denote the kth segmentation that contains $x_\ell$ by $s^{\ell k}$. Then the value of the large segmentation S at the point $x_\ell$ is given by

Equation (36)

Processing the large data set in smaller portions is beneficial because each portion can be run independently in parallel. Additionally, smaller data sets are computationally faster to run through the segmentation algorithm. The next section will discuss the application of the segmentation algorithm to the large grain ECT data.

5. Results and discussion

5.1. Segmentation along second Euler angle

As previously mentioned, We simulated ECT data sets of the large grain specimen for two rotations of the ECT coil (specifically 0 and 90). The imaginary portion of the noiseless simulated data sets is shown in figures 3(c) and (d). To both the real and imaginary portions of this data, we added zero-mean Gaussian noise with standard deviation equal to 5% of the maximum value of the signal. Note that although we only show the imaginary portions, both the real and imaginary portions of the ECT data were used to estimate the segmentation.

We applied the process outlined in section 4.3 to the noisy ECT data to find the segmentation of the specimen along changes in the second Euler angle. We estimated the segmentation assuming both the SAM structural prior and the image deblurring prior. For the structural prior, we used a value of ε = 0.005 for the coupling constant and ρ = 5 for the regularization parameter. For the image deblurring prior, we chose $\gamma = 2e-3$. The results are shown in figure 7. The image deblurring prior does well at estimating the orientation of the individual regions. However, it struggles to resolve some of the boundaries and smaller features. In contrast, the segmentation using the structural prior has sharp boundaries, and is able to resolve the small features that are missed with the image deblurring prior.

Figure 7.

Figure 7. Segmentation results for the second Euler angle. (a) Segmentation of ECT data using the image deblurring prior. (b) Segmentation of ECT data using SAM as a structural prior. (c) True value of second Euler angle at each point. (d)–(f) Zoomed-in view of the outlined areas in (a)–(c), respectively.

Standard image High-resolution image

To further illustrate this point, we also show a zoomed-in image of a subset of each result in figures 7(d)–(f). These images highlight the inability of ECT to resolve certain features on its own, such as the small teardrop region in the bottom middle of the image. In addition, it is obvious that the estimated boundary of the larger region is much clearer with the addition of the structural prior.

Another consideration in comparing these two methods is the computational time required to compute each segmentation. Both estimates were computed on a Intel Core-i9 with a 2.4 GHz CPU with 8 cores. The result using the image deblurring prior took around 30 min to compute, while the structural prior took approximately 12 min. This is due to the fact that the image deblurring prior requires an iterative process to find the unknown segmentation, while the structural prior only needs to solve a single linear system. Thus, the addition of structural information greatly decreases the computational time.

5.2. Effect of missing edges

We also consider the effect of missing information in the structural prior. We removed some of the edges shown in figure 5 and reran the ECT inversion using the structural prior. The results are shown in figure 8. Note that we only show the result for the zoomed-in area where edges were removed, as the rest of the solution remains the same. For comparison, we also show the result with all edge information, as shown in figure 7.

Figure 8.

Figure 8. Results of structural prior inversion with missing information. (a) Original edges from SAM data. (b) Results from inversion using all edges. (c) Partial edge information. (d) Inversion with partial edges.

Standard image High-resolution image

Without edge information, the algorithm reverts to a smoothness prior. While it will still detect an orientation change, boundary information is completely lost. Furthermore, the teardrop shaped feature is no longer visible. Ideally, we would like for the algorithm to revert to the image deblurring prior if no structural information is available. Although the boundary of the small feature is not well defined using the image deblurring prior (as seen in figure 7(d), the feature is still visible. Future work will modify the structural prior so that it will revert to the image deblurring prior if necessary, similar to what was done in [19], which utilized total variation regularization if structural information was not available.

5.3. Effect of multiple ECT coil rotations

As previously mentioned, we use ECT data from multiple rotations of the ECT coil in order to produce the images in figure 7. However, for illustrative purposes, we also estimated the segmentation using ECT data from a single rotation of the coil. In particular, we computed the segmentation using data from only the 0 rotation and only the 90 rotation using both the image deblurring prior and the structural prior. These results are shown in figure 9. When using the image deblurring prior, neither rotation on its own is sufficient to recover information about the grain edges. However, even with the edge information provided by the structural prior, multiple rotations of the coil are still needed to provide correct orientation information. This is due to the fact that two different values of θ can produce the same ECT response for a single rotation of the coil depending on the value of ψ1. Adding a second rotation of the coil resolves this uncertainty, as is evident in figure 9.

Figure 9.

Figure 9. Results of segmentation algorithm using either one or two rotations of the ECT coil. The top row shows segmentation results with the image deblurring prior using ECT data from (a) only the 0 rotation, (b) only the 90 rotation and (c) both 0 and 90 rotations. The bottom row (d)–(f) shows the same results as (a)–(c) but for the structural prior.

Standard image High-resolution image

5.4. Segmentation along first Euler angle

As shown in [33], we can also use this method to segment the ECT data along changes in the first Euler angle, with a caveat. Recall that for the HCP crystal, $\psi_1 \in [0, 2\pi]$. However, due to the symmetry of the crystal, it is more difficult to develop a labeling scheme segmentation along the first Euler angle. In particular, certain crystallographic orientations can be represented by multiple different values of ψ1. For instance, if $\theta = \pi/2$, then ψ1 = 0 or $\psi_1 = \pi$ describe the same orientation. This ambiguity will cause the algorithm to fail. To avoid this non-uniqueness issue, we enforce $\psi_1 \in [0, \pi/2]$ when running the segmentation algorithm. Developing a labeling system for ψ1 is the topic of future work.

We relabeled the training segmentations based on the value of the first Euler angle and reran the process described in section 4.3 assuming both the image deblurring prior and the structural prior. The results are shown in figure 10. Similar to the results with the second Euler angle, the estimates of the orientation using either prior are very similar. However, as before, the image deblurring prior is unable to resolve many of the boundaries between grains, while the structural prior provides much sharper boundaries. We note that ECT is generally less sensitive to the value of ψ1, particularly if the value of θ is close to zero. As such, the estimates of ψ1 using either prior are much less accurate for values of θ close to zero.

Figure 10.

Figure 10. Segmentation results for first Euler angle. (a) Segmentation assuming the image deblurring prior. (b) Segmentation assuming the structural prior. (c) True value of first Euler angle at each point in the specimen.

Standard image High-resolution image

We can combine the segmentation results for each Euler angle into a single image. Figure 11 shows the combined results for the image deblurring method and the structural prior. Each point in the result is colored based on the estimated value of the first and second Euler angle at that point. Also shown in figure 11 is the PLM data of the specimen. The result using structural prior matches the PLM data very well. The advantage of the structural prior is twofold: it improves the resolution of the final image, and it speeds up computation time. The structural prior shows promise to take advantage of the benefits of each imaging modality.

Figure 11.

Figure 11. Combined segmentation of the first and second Euler angle using (a) image deblurring prior and (b) the structural prior. (c) True value of the first and second Euler angle at each point in the specimen.

Standard image High-resolution image

6. Conclusions

This work demonstrated a method to combine SAM and ECT data in order to generate a segmentation of a large grain titanium specimen. The SAM data was used to provide boundary information, while the ECT data provided information on the crystallographic orientation of each region. Using SAM data for structural information provided better detail and faster computational times than using ECT data on its own.

Furthermore, this work shows that nondestructive evaluation methods could benefit from adopting algorithms developed for medical imaging. Although the application spaces are quite different, nondestructive evaluation and medical imaging share similar issues. Here, we drew inspiration from medical imaging research to mitigate the issue of low spatial resolution with ECT. We will continue to consider work from the medical field to address challenges in nondestructive evaluation.

While in this work we only considered ECT data that had been corrupted by Gaussian noise, we recognize that other sources of noise may be present with experimental ECT data. In particular, there may be uncertainty in the parameters associated with the geometry of the ECT coil, such as liftoff, tilt, or size. We were able to assume with simulated data that these parameters were known exactly. We anticipate that with these additional sources of uncertainty, the boundaries detected with the image deblurring prior may suffer and the structural prior will become even more necessary.

As previously mentioned, further improvements to the structural prior can still be made. In its current formulation, it will revert to a smoothness prior if no boundary information is provided. However, it would be ideal if instead it reverted to the image deblurring prior, as this yields a more desirable result than the smoothness prior. This will be the topic of future work.

The ultimate goal is to apply this method to a realistic microstructure. The large grain specimen has much harder boundaries than a specimen with true MTR, meaning that the boundaries are generally easier to locate in the SAM data. As such, a more sophisticated method will likely be required to determine MTR boundaries from SAM data. This is the topic of future work.

Acknowledgments

The authors gratefully acknowledge funding from the Air Force Office of Scientific Research through Grant 21RXCOR037 under the Dynamic Data and Information Processing program. Dr Homa and Mr Lesthaeghe also acknowledge support from the Air Force Research Laboratory through Contract FA8650-19-D-5230.

Data availability statement

The data cannot be made publicly available upon publication because they are not available in a format that is sufficiently accessible or reusable by other researchers. The data that support the findings of this study are available upon reasonable request from the authors.

Appendix: Computation of CGMCA maps

This section provides the details on how to compute the CGMCA maps; specifically how to compute Ai and bi for each data domain. Let $Z_i \in \mathbb{R}^{d_i \times n}$ denote the data matrix for the ith data domain, where the $\ell$th column of Zi contains the $\ell$th sample and n is the number of samples from each data domain. The algorithm to compute Ai and bi is as follows.

Algorithm 1. CGMCA algorithm.
1: for $i \in \{1 , 2 \}$ do
2:    Compute sample mean µi from each data matrix Zi
3:    Compute centered data matrix $S_i = Z_i - \mu_i$
4:    Compute thin SVD $S_i = U_{S_i } \Sigma_{S_i } V_{S_i }^T$
5:    Compute thin SVD $\Gamma_i = U_{\Gamma_i } \Sigma_{\Gamma_i }^2 U_{\Gamma_i }^T$
6:    if $\text{rank}( S_i ) \lt \text{rank}( \Gamma_i )$ then
7:      return error
8:    else
9:      Compute matrix $\Sigma_{\Gamma_i } = ( \Sigma_{\Gamma_i }^2 )^{1 / 2 }$
10:    end if
11: end for
12: Compute matrices $A = \Sigma_{\Gamma_1 } U_{\Gamma_1 }^T U_{\Gamma_2 } \Sigma_{\Gamma_2 }$ and $B = V_{S_1 }^T V_{S_2 }$
13: Compute full SVDs $A = U_A \Sigma_A V_A^T$ and $B = U_B \Sigma_B V_B^T$
14: Compute $D_1 = U_A U_B^T$, $D_2 = V_A V_B^T$
15: for $i \in \{1, 2\}$ do
16:    Compute linear map $A_i = U_{\Gamma_i } \Sigma_{\Gamma_i } D_i \Sigma_{S_i }^{-1} U_{S_i }^T$
17:    Compute translation $b_i = -A_i \mu_i$
18: end for
Please wait… references are loading.
10.1088/1361-6420/ad366e