0% found this document useful (0 votes)
54 views12 pages

EBSD Calssification

Uploaded by

Meghraj Prajapat
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
54 views12 pages

EBSD Calssification

Uploaded by

Meghraj Prajapat
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 12

www.nature.

com/scientificreports

OPEN Efficient few‑shot machine learning


for classification of EBSD patterns
Kevin Kaufmann1, Hobson Lane2,3, Xiao Liu4 & Kenneth S. Vecchio1,4*

Deep learning is quickly becoming a standard approach to solving a range of materials science
objectives, particularly in the field of computer vision. However, labeled datasets large enough to
train neural networks from scratch can be challenging to collect. One approach to accelerating the
training of deep learning models such as convolutional neural networks is the transfer of weights
from models trained on unrelated image classification problems, commonly referred to as transfer
learning. The powerful feature extractors learned previously can potentially be fine-tuned for a new
classification problem without hindering performance. Transfer learning can also improve the results
of training a model using a small amount of data, known as few-shot learning. Herein, we test the
effectiveness of a few-shot transfer learning approach for the classification of electron backscatter
diffraction (EBSD) pattern images to six space groups within the 4/m32/m point group. Training


history and performance metrics are compared with a model of the same architecture trained from
scratch. In an effort to make this approach more explainable, visualization of filters, activation maps,
and Shapley values are utilized to provide insight into the model’s operations. The applicability to
real-world phase identification and differentiation is demonstrated using dual phase materials that are
challenging to analyze with traditional methods.

Data science-based methods to materials development and analysis have gained great popularity in recent
­years1–13. Deep learning algorithms are of significant interest owing to their excellent performance without sig-
nificant feature engineering, and the ubiquity of these methods will likely continue owing to the outperformance
of systems directly designed by humans. While often difficult to assess how and why these ‘black box algorithms’
are capable of performing these tasks, these methods can provide significant value or spark new i­nsights14,15.
Application of these tools to image-based tasks in materials science has proved to be useful for c­ lassification16–19,
­segmentation20–22, and other ­objectives23–25. While deep learning provides significant opportunities for the
advancement of materials science, robust application of these tools often requires much larger datasets than are
typically available within the materials science community. Utilizing the knowledge deep neural networks have
learned from other domains offers an opportunity to develop models in these domains, where data is sparse and
further collection and labeling could be slow or t­ edious26–28.
Convolutional neural networks (CNNs) are a class of deep learning models that have proven effective for
analyzing image d ­ ata29. Before a CNN can be applied to a given task, it must learn to assign importance (learnable
weights and biases) to various aspects of the image that maximize the network’s differentiation capabilities. Two
general strategies exist for training convolutional neural networks: (1) the weights can be randomly initialized,
or (2) the weights can be transferred from a model pre-trained on a separate but related task, often in a nearby
domain with significantly more data, and then refined for the current objective. The first approach, commonly
referred to as "training from scratch", requires a large dataset to avoid overfitting and perform robustly on new,
real-world examples. The second approach, referred to as "transfer learning", can significantly reduce the number
of training examples required, accelerate the training process, and retain or exceed the performance garnered
by training from ­scratch27,30–32. The transfer learning method is motivated by the human ability to intelligently
apply previously learned knowledge to solve new problems faster or with better ­solutions33. Despite the potential,
knowledge transfer from a given source domain is not guaranteed to improve performance in the target domain
and can in fact hinder p ­ erformance26,32. Furthermore, one of the requirements to use this approach is that the
images in the new domain must conform to the processed shape and structure determined at the outset of the
previous training. For models pre-trained on ­ImageNet34, a library of over one million images labeled with one
thousand classes, the expected input is usually 299 × 299 pixels with 3-channels (one each for RGB). When dealing

1
Department of NanoEngineering, UC San Diego, La Jolla, CA 92093, USA. 2Tangible AI LLC, San Diego,
CA 92037, USA. 3Department of Healthcare Research and Policy, UC San Diego-Extension, San Diego, CA 92037,
USA. 4Materials Science and Engineering Program, UC San Diego, La Jolla, CA 92093, USA. *email: kvecchio@
eng.ucsd.edu

Scientific Reports | (2021) 11:8172 | https://doi.org/10.1038/s41598-021-87557-5 1

Vol.:(0123456789)
www.nature.com/scientificreports/

with one-channel grayscale images, such as those typically collected from electron diffraction studies, the decision
to use transfer learning necessitates the stacking of a single image into a pseudo-color i­ mage31.
The number of labeled images that can reasonably be collected must also be considered for the appropriate
training and application of a CNN. Computer vision research has recently been motivated by children’s ability
to learn novel visual concepts almost effortlessly after accumulating sufficient past k­ nowledge35. In deep learning
and computer vision, learning visual models of object categories has notoriously required tens of thousands of
training ­examples36; however, recent research has demonstrated that it is possible to classify images accurately
using relatively few labeled examples with the appropriate combination of pretraining of the CNN layers on
unrelated image classification training s­ ets37,38, adversarial or unsupervised l­ earning39,40, network p
­ runing41, and
micro architecture t­ uning42. Once several categories have been learned the hard way, learning new categories
should become more efficient. The increased efficiency allows for a lesser number of images to be used in train-
ing, referred to as a “few shots”.
Electron backscatter diffraction (EBSD) patterns (EBSPs) are an excellent case study for the use of few-shot
transfer learning toward accelerating analysis of electron diffraction data. The scanning electron microscope
(SEM)-based method involves the capture of 2D diffraction patterns produced from an incident electron beam
scattering, diffracting, and escaping from a well-polished ‘bulk’ ­sample43. The collected diffraction patterns con-
tain significant structural and chemical information and are similar to those collected in other techniques such
as convergent beam electron diffraction (CBED)44,45. Despite the vast amount of information in the patterns,
conventional EBSD has primarily focused on determining the three-dimensional orientation of individual grains
in crystalline ­materials43,46–48. Furthermore, the commercial technique typically relies on Hough-based indexing
with a look-up table of interplanar angles constructed from the set of selected reflectors for phases specified by
the ­user49. This generally allows for phase differentiation of sufficiently distinct crystal s­ tructures50–52, but the
process remains susceptible to structural ­misclassification53–55. Improvements to phase differentiation have been
proposed and developed including dictionary i­ ndexing56–59, spherical i­ ndexing60–62, and more recently machine
­learning63. While each offers significant advantages over the Hough-based method, these tools continue to require
assumptions about the number of phases and/or their structure. For example, the dictionary-based approach
requires simulation of a “master” pattern for each potential phase and every experimental pattern is matched
against a dictionary of simulated patterns for all potential p ­ hases56. The highest similarity match is selected as the
most likely phase and orientation. Another available solution to phase differentiation is combining an electron
backscattered pattern (EBSP) with information from energy dispersive X-ray spectroscopy (EDS) or wavelength
dispersive X-ray spectroscopy (WDS)64–66. While these have been adopted commercially, the singular EBSP is still
analyzed with the Hough-based method and an expert user is required to evaluate the plausibility of returned
matches. Applications using hand-drawn lines overlaid on individual Kikuchi diffraction patterns have been
developed for determining the Bravais lattice or point group; however, they remain limited by at least one of the
following: analysis time per pattern, the need for an expert crystallographer, or necessitating multiples of the
same diffraction pattern with different SEM ­settings67. Clearly, there remains a need for a rapid EBSP classifica-
tion tool capable of functioning with one EBSD pattern while remaining suitable for even the most novice user.
Recently, the EBSD community has begun to explore the use of convolutional neural networks as a foun-
dation for addressing modern EBSD ­challenges16,68–70. Despite the marked advances these works have made,
the requirements for ­simulating71 or collecting experimental diffraction patterns from a significant number of
materials and crystallographic orientations remains a limiting factor. In this work, we test the validity of few-shot
transfer learning, starting from ImageNet weights, applied to classify EBSPs to one of six space groups. Herein,
space group refers to the symmetry group of a configuration in three-dimensional space. We compare the time
to converge, the individual kernels (weights of the neurons), activation maps, and the performance of models
trained from scratch or by transfer learning. Though there has been considerable progress on interpretability
of machine learning systems, mostly in the field of eXplainable AI (XAI)72, fully understanding the internal
mechanisms of deep neural networks is still an area of active r­ esearch15,73. Visualization of the most similar
weights each model independently learned, their respective activation maps, and Shapley values can increase
understanding of how the artificial intelligence models accomplishes its task. Building these XAI foundations
can increase trust in the model’s later predictions and help identify when the prediction is incorrect. In addition
to evaluating each model’s performance on holdout data, each model is also tested with 6900 EBSPs collected
from a ­Ni90Al10 sample outside the training, testing, and validation data, and a space group map is generated
from the individual classifications.

Results and discussion


Training metrics. Each model is first trained until the validation loss converged using a small number of the
available diffraction patterns from each material in the six available space groups. The training and validation
loss were recorded at the end of each pass through the training set (i.e. an epoch) to monitor the model’s per-
formance as the weights are updated. The validation loss is also monitored to determine when training should
cease to prevent model underfitting or overfitting owing to a fixed number of training epochs. The loss function
(categorical cross-entropy) is plotted in Supplementary Fig. S1 for reference. The goal of the neural network is
to minimize the loss function by maximizing its prediction of the correct class. Figure 1a shows statistics for the
training and validation loss for the model trained from scratch in grayscale. While the average training loss is
low (0.84 ± 0.04) by the end of the first epoch, the average validation loss is observed to be 11.5 ± 2.5. In com-
parison, by initializing the model with weights learned from ImageNet, the average training and validation loss
are both observed to be small (0.56 ± 0.02 and 0.52 ± 0.3, respectively) from the start (Fig. 1b). Figure 1c shows
a magnified view of Fig. 1b to show the detailed training history. The improved starting loss, presumably owing
to the strong filters learned in pre-training, also aids in rapid model convergence. Including the 15 epochs used

Scientific Reports | (2021) 11:8172 | https://doi.org/10.1038/s41598-021-87557-5 2

Vol:.(1234567890)
www.nature.com/scientificreports/

Figure 1.  Training and validation history statistics for the two models. (a) The model trained from scratch
using grayscale images. (b) The model trained starting from ImageNet weights using 3-channel stacked images.
Data is plotted on the same y-axis as in (b). (c) The data in (b) for the transfer learning model plotted on a
different scale. Error bars are one standard deviation from five trials per approach.

Precision Recall
Trained from scratch 0.93 0.92
Transfer learning 0.97 0.96

Table 1.  Classification metrics. The class-weighted average Precision and Recall on the test data for each
model. Both metrics are improved using the transfer learning approach to training the model. The same test
data was used to benchmark each model.

to establish convergence, the model trained from scratch required 50 ± 12 epochs, while the transfer learning
model required only 26 ± 3 epochs; this represents a twofold reduction in the average number of passes through
the training set. Since each epoch with 2400 diffraction patterns requires two minutes on this hardware, a time
savings of nearly an hour per training event is gained. With the amount of training data available to the CNN
expected to grow, and therefore the time per epoch expected to increase, the time savings will become more
pronounced.

Performance on holdout data. The time saving advantages garnered by fine-tuning the weights of an
existing model are only valuable if the model performs similarly well or exceeds the model trained from scratch.
Table 1 shows the class-weighted Precision and Recall for the best performing model from each approach, for
which further analysis into the internal operations will be studied. See Supplementary Tables S1 and S2 for a
breakdown of Precision and Recall by space group for the transfer learned and trained from scratch models,
respectively. Both metrics are improved for the model using transfer learning even though the training, valida-
tion, and test sets were held constant. This implies that the feature extractors learned from ImageNet are not only
relevant to this new domain, but also at least as valuable as those learned from scratch. This likely results from

Scientific Reports | (2021) 11:8172 | https://doi.org/10.1038/s41598-021-87557-5 3

Vol.:(0123456789)
www.nature.com/scientificreports/

Figure 2.  Kernel pairs with lowest Euclidean distance. Kernel pairs are grouped by column between the two
models. The leftmost pair has the lowest Euclidean distance, with the kernels becoming less similar moving to
the right.

the more general nature of the feature extractors necessary for optimal performance on I­ mageNet74. It is possible
that increasing the size of the dataset used to train the model from scratch could increase its performance to near
that of the few-shot transfer learning model; however, the cost to training time would be notable. For this study,
it was also important to keep the dataset and its partitions fixed for more deterministic comparisons.

Visualization. Since deep neural networks are used in this study, meaningful information about the internal
mathematical operations performed necessitates studying several aspects of the model’s inner workings. The
first study involves visualization of the filters (or kernels) and corresponding feature maps. Filters and feature
maps from the earliest layers in the model are most useful since deeper layers become more abstract. To identify
corresponding filters in the two models, the Euclidean distance between all possible pairs of filters in a selected
layer from the grayscale and pseudo RGB models were calculated and the similarity was ranked. The four most
similar kernels between the two models in the first convolution layer are shown in Fig. 2. The pairs are grouped
in columns between the model trained from scratch and using transfer learning. The independent co-learning
of these filters suggest they are very valuable for feature tuning and selection on EBSPs. At this low-level in the
model, the filters will predominantly be designed to identify edges at various orientations. The next layer will
likely combine those edges into corners and small points, and the subsequent layers will figure out larger and
larger shapes/features, such as the number of and relative angles between Kikuchi lines. Two of these four filters
are recognizable as classical edge detection operators. The second filter has converged close to the x-direction
Sobel edge detection ­operator75 and the third filter resembles the Gabor filters with theta θ = 135◦. Visualization
of the feature maps resulting from these four filters can provide insight into their function.
Figure 3 shows the result of each filter from Fig. 2 being individually convolved over an input image from the
six space groups studied. All feature maps shown are from the grayscale model. The first filter, and therefore the
most similar filter between the two models, is primarily observed to perform an inversion on the input image.
As a result, the band edges in each image are activated (white regions) and become more distinct. It is reason-
able to speculate that deeper parts of the network extract the angle and relative locations of these intersections,
a function that was also suggested in Ding et al.’s recent w ­ ork70. The fourth filter is observed to have normalized
the contrast of the input image, perhaps to reduce the effects of atomic scattering factors (Z-contrast) observed
in prior w­ ork68. Further analysis of the class-specific feature importance can further improve understanding of
the neural network’s methodology.

Feature importance. Measurement and visualization of feature importance is another method to increase
understanding of the deep neural network. This type of analysis is performed to help determine whether one
should trust a prediction and why. There are a number of techniques available for CNNs including Gradient-
weighted Class Activation Mapping (Grad-CAM)76, activation ­maximization77, ­LIME78, and Shapley v­ alues79.
Several of those listed have been effectively demonstrated in other works involving EBSD p ­ atterns16,17,68,70; how-
ever, this is the first to use Shapley values. In game theory, Shapley values are a solution to fairly distributing the
gains and costs of several actors working in coalition. By definition, each Shapley value is the average expected
marginal contribution of one actor after all possible combinations have been considered. In this case, the actors
are the features in the images. Essentially, the Shapley value is the average expected marginal contribution of one
actor after all possible combinations have been considered. While not perfect, this has proven a fair approach
to allocating value in a variety of fi­ elds80–82. The results of this analysis, further described in the Methods sec-

Scientific Reports | (2021) 11:8172 | https://doi.org/10.1038/s41598-021-87557-5 4

Vol:.(1234567890)
www.nature.com/scientificreports/

Figure 3.  Selected feature maps from the first convolution layer. Feature maps extracted from the grayscale
model corresponding to the filters shown previously in Fig. 2. One input image per space group is shown. From
top to bottom, the six space groups are Pm3m, Pm3n,Fm3m,Fd3m,Im3m, and Ia3d.

tion, are shown in Fig. 4. Refer to Supplementary Fig. S2 for a demonstration of SHAP analysis on handwritten
numbers.
The previously unseen input images are shown in the top row, and as semi-transparent grayscale backings
behind each of the explanations. After random selection of the input image, it was verified that the model cor-
rectly identified the space group (thus pseudo-random selection). The middle row in Fig. 4 corresponds to the
explanations for the correct prediction, while the bottom row displays the explanations for the next most likely
(i.e. incorrect) space group. All 6 Shapley explanations ordered by most to least likely space group for each input
are displayed in Supplementary Fig. S3. The positive (red) and negative (blue) contributions to each prediction
are primarily at diffraction maxima (e.g. zone axes), band intersections, and outlining band edges. This lends
credence that the model is indeed utilizing information grounded in the physics of EBSD. The clustering of
Shapley values near zone axes is further reaffirming given the abundance of information and their role in clas-
sical diffraction pattern i­ ndexing83.

Performance in practice. It is also of importance to demonstrate and compare the efficacy of both models
in a real-world context, not only the patterns set aside for testing. In this case, the EBSD mapping of a dual phase
sample serves to demonstrate that both model training strategies are capable in situations where the diffraction
patterns are not collected from an ideal, single-phase material. Figure 5 top-left shows a backscattered electron
image of a N­ i90Al10 (wt%) sample containing a Ni-rich matrix (space group 225) along with N ­ i3Al precipitates
(space group 221) appearing raised from the surface. It is of importance to note that the model has not yet
encountered a solid-solution phase such as the Ni with Al matrix present in this sample. The Al content in the Ni
matrix was determined to be 19.7% ± 0.53% (at%). The Al chemistry of the N ­ i3Al precipitates was determined to

Scientific Reports | (2021) 11:8172 | https://doi.org/10.1038/s41598-021-87557-5 5

Vol.:(0123456789)
www.nature.com/scientificreports/

Figure 4.  Visual explanation of feature contributions. Shapley values are computed for each input image to
gauge the importance of features in the EBSPs. The first row is the raw input image. The second row corresponds
to the Shapley values for the correct prediction. Row three corresponds to the first incorrect classification as
ranked by softmax probability. From left to right, the six space groups are Pm3m Pm3n,Fm3m,Fd3m,Im3m, and
Ia3d.

Figure 5.  Phase mapping a dual-phase sample. Top left shows a backscattered electron image of the area to be
mapped. The ­Ni3Al precipitates appear raised in the Ni-rich matrix. Top middle shows the Hough transform-
based phase map. Red pixels are identified as Ni or ­Ni3Al with equal certainty, while black pixels were not
solved. Top right is the inverse pole figure map in the Y-direction. Bottom left is the phase map produced using
the predictions from the grayscale CNN. Bottom middle shows the aluminum EDS map. Bottom right is the
phase map produced using the predictions from the few-shot transfer learning approach. There are 6900 total
EBSPs (pixels). Scale bar = 25 µm.

Scientific Reports | (2021) 11:8172 | https://doi.org/10.1038/s41598-021-87557-5 6

Vol:.(1234567890)
www.nature.com/scientificreports/

221 223 225 227 229 230


Scratch 984 0 5822 10 72 12
Transfer 1501 0 5342 0 56 1
Difference 517 0 (480) (10) (16) (11)

Table 2.  Tabulated predictions for the dual-phase sample. The number of patterns classified to each space
group by the respective model. The difference is calculated by subtracting the number predicted by the model
trained from scratch from the number predicted by the few-shot transfer learning model. Parentheses denotes
the transfer learning model predicted fewer EBSPs belonging to the respective class. From left to right, the six
space groups are Pm3m, Pm3n,Fm3m,Fd3m,Im3m, and Ia3d.

be 25.7% ± 0.74% (at%). While partially selected for this reason, the material was primarily chosen since space
groups 221 and 225 have two of the lowest F1-scores in each model and can readily be produced in this singular
sample. Furthermore, it was important that the phases could be differentiated visually and easily to the reader,
such as with EDS maps. While this means the phases could potentially be differentiated if the EBSD operator
assigned reference chemistries to the phases in advance of collecting the EBSPs, the purpose of this demonstra-
tion is really to compare the transfer learning and trained from scratch models’ abilities to identify the space
group without further information. For the most complete phase identification in EBSD (i.e. lattice parameters),
multiple analysis methods (e.g. XRD and EDS) may need to be employed.
The Hough-based phase map in Fig. 5 is a representative image of the expected results from commercial
systems. The phase map consists of entirely one phase (shown in red) and Oxford Aztec software is observed to
predict Ni and N­ i3Al with an equal number of bands and mean angular deviation (MAD), effectively a combined
measure of certainty, for each diffraction pattern. It is the order of phase selection by the user in Oxford Aztec
software that ultimately determines whether Ni or ­Ni3Al is selected in this case. Example diffraction patterns
from each phase are shown in Supplementary Fig. S4. Black pixels remain unindexed by the Hough method,
typically a result of poor diffraction pattern quality. The inverse pole figure (IPF) in the Y-direction is provided
along with the EDS map for aluminum to elucidate where the N ­ i3Al is expected to be found within the Ni matrix.
The last two images in Fig. 5 show the phase maps produced by the scratch (­ MLS) and few-shot transfer learned
­(MLT) models. Over a statistical number of diffraction patterns, the two models are expected to produce similar
answers with some variance, although the test set would suggest that the transfer learning model will generally
outperform. Comparing the results with the Al EDS map and IPF Y, both models perform well at identifying
the space group of each EBSP with a low false positive rate for the four space groups known not to be present.
There are almost certainly some errors with regard to the EBSPs classified as space group 221 ( Pm3m) or 255
( Fm3m), the Precision and Recall of each model alert the user of this in advance; however, the results over these
6900 EBSPs demonstrate the improvement over the Hough-based approach and, more importantly for this study,
the robustness of a few-shot transfer learning approach.
The number of EBSPs each model classifies to the available space groups is tallied in Table 2. A total of
6900 diffraction patterns were individually identified by each neural network without any other information
provided. The transfer learning model is observed to have a reduced misclassification rate to the space groups
227 ( Fd3m), 229 ( Im3m), and 230 ( Ia3d ) in this phase map. The largest difference between the two models is
for the diffraction patterns classified to space group 221; the class with the lowest Precision for each of the two
models. The total difference of 517 diffraction patterns only equates to 7% of the total diffraction patterns; well
within reason and the expected margin of error between the two models, particularly between these two space
groups for the current models. Of those 517, only 480 of these predictions differ between space groups 221 and
225, the other 37 differences are due to false positives (i.e. 227, 229, and 230) in the model trained from scratch.
The phase fraction of N ­ i3Al likely lies somewhere between what is predicted by these two models. The results
of this comparison also suggests that there is future opportunity to construct an approach leveraging Bayesian
deep learning or an ensemble of (i.e. at least two) individually trained models making individual classifications
combined with model averaging (e.g. voting) to reduce variance, provide insight into overall uncertainty, and
identify when “no solution” is an appropriate a­ nswer84–86.
Thus, a few-shot transfer learning approach to classifying electron backscatter diffraction patterns is an attrac-
tive method for leveraging the knowledge a deep neural network has attained in a previous context. The convolu-
tional neural network-based approach to diffraction pattern classification is advantageous in that it requires little
or no a-priori knowledge of the phases in a new sample and can readily be improved or expanded to new classes
with the inclusion of new data. The similarity of EBSD patterns to those from techniques such as CBED suggests
the few-shot transfer learning approach could also apply and potentially be more beneficial given the slower rate
of data collection with other electron diffraction methods. Limitations of the current models exist in the number
of space groups currently differentiable and the “black box” nature of neural networks. The number of space
groups the model can learn to differentiate can be continuously expanded as more data becomes available for
training. One of the goals of this work is to discern whether the few-shot transfer learning approach can be used
to reduce the amount of data necessary for robust expansion to all 230 space groups or other diffraction pattern
classification tasks. Indeed, we find that this approach does not hinder the model’s performance on holdout or
entirely new data and offers accelerated training time. While it can be difficult to precisely determine how the
CNN performs this task, recent advances in eXplainable AI (e.g. SHAP) provides tools for developing insight
and trust in the model’s predictions. The combination of ease of scaling, flexibility of the framework, and ability

Scientific Reports | (2021) 11:8172 | https://doi.org/10.1038/s41598-021-87557-5 7

Vol.:(0123456789)
www.nature.com/scientificreports/

to assess aspects of a model’s decision process support the utilization of CNNs and few-shot transfer learning as
another tool for phase differentiation and symmetry identification in electron diffraction.

Materials and methods


Materials.  Eighteen different single-phase materials, comprising 6 of the 10 space groups within the
4/m32/m point group, were selected for training the space group classification CNN. Suitable samples for the
remaining 4 space groups could not be obtained. The six space groups are Pm3m, Pm3n,Fm3m,Fd3m,Im3m, and
Ia3d . Numerically, these are space groups 221, 223, 225, 227, 229, and 230. Space groups 221 and 223 are primi-
tive cubic, 225 and 227 are face centered cubic, and 229 and 230 are bodycentered cubic. Each of the six space
groups share the threefold rotary inversion necessary for inclusion in the 4/m32/m point group. Supplemen-

tary Table S3 details the similarities and differences between the symmetry operations of the six space groups.
The materials were FeAl, NiAl, ­Ni3Al, ­Fe3Ni, ­Cr3Si, ­Mo3Si, Ni, Al, NbC, TaC, TiC, Si, Ge, W, Ta, Fe, ­Al4CoNi2, and
­Al4Ni3. These materials were of low texture, typically less than 2 times random in any direction. Refer to Kauf-
mann et al. for the distributions of orientation, band contrast, and mean angular deviation for these s­ amples69.
A dual-phase material known to challenge Hough-based EBSD was fabricated to demonstrate and compare
the capabilities of each CNN training approach. An additional constraint for the material selected was that the
two space groups be identifiable within an EDS map, even though this meant an operator could force the Hough-
based method to differentiate the phases by chemistry if they knew the phases in advance. An ingot of N ­ i90Al10
(wt%) was arc melted and processed via hot rolling at 600 °C to 45% reduction in thickness followed by aging
at 600 °C for 4 h and air cooled. X-ray diffraction (XRD) using a Rigaku Miniflex X-ray Diffractometer with a
1D detector, a step size of 0.02°, 5° per minute scan rate, and Cu Kα radiation (wavelength λ = 1.54059 Å) was
performed to confirm the existence of phases belonging to space groups 221 ( Pm3m) and 225 ( Fm3m) (Sup-
plementary Fig. S5).

Electron backscatter diffraction pattern collection. EBSD patterns (EBSPs) were collected as previ-
ously described in Kaufmann et al.68. Diffraction patterns were collected using a Thermo Scientific (formerly
FEI) Apreo scanning electron microscope (SEM) equipped with an Oxford Symmetry EBSD detector utilized
in high resolution (1244 × 1024) mode. The geometry of the setup was held constant as follows. The working
distance was 18.1 mm ± 0.1 mm. Oxford Aztec software was used to set the detector insertion distance to 160.2
and the detector tilt to -3.1. The imaging parameters were 20 kV accelerating voltage, 51 nA beam current,
0.8 ms ± 0.1 ms dwell time, and 30-pattern averaging. The Hough indexing parameters were 12 Kikuchi bands, a
Hough resolution of 250, and band center indexing.
After collecting high resolution EBSPs from each material, all patterns collected were exported as tiff images.
The images were resized for the CNN using the resize function in scikit-image. All collected data for each mate-
rial was individually assessed by the neural network, and the collection of images for each sample may contain
partial or low-quality diffraction patterns, which could decrease the accuracy of their identification. The test
data was not filtered to better assess the model as it would be applied in practice.

Neural network architecture. The well-studied convolutional neural network architecture X


­ ception87
was selected as the basis architecture for fitting a model that determines which space group a diffraction pattern
originated from. The Xception architecture was used without modifications for training the model from scratch
and the transfer learning process to facilitate comparison of the training metrics, performance, and internal
workings. Selection of this network was partially based on Xception or derivatives of Xception being used previ-
ously in the EBSD ­community16,68–70. Xception is also a standard model with ImageNet weights readily available
in deep learning APIs such as ­Keras88. A schematic of the convolutional neural network operating on an EBSP is
provided in Fig. S6. Due to space constraints, only the resultant feature maps from selected convolutional layers
are shown after image input and before the 2048-dimensional vector. For a complete description of the Xception
architecture, please refer to Fig. 5 in Xception: Deep Learning with Depthwise Separable C ­ onvolutions87.

Neural network training. For both the transfer learning and from scratch approaches, training was per-
formed using 400 diffraction patterns per space group. The diffraction patterns supplied at training were evenly
divided between the number of materials per space group that the model had access to during training. For
example, if the model was given two materials of the same space group during training, 200 diffraction pat-
terns per material were made available. The validation set contained 100 diffraction patterns per space group,
equivalent to the standard 80:20 train/validation split. The validation set was only used to monitor the training
progress and model convergence. The test set contains the rest of the patterns (a total of 145,453 images; refer to
Supplementary Table S1 for class distribution) that were not used for training or validation. The images selected
for training, testing, and validation were the exact same for transfer learning and for from scratch learning.
Model hyperparameters were selected or tuned as follows. Adam (adaptive moment estimation) optimization
with a learning rate of 0.00189, and a minimum delta of 0.001 as the validation loss were employed for convergence
criteria. Adam is chosen for its ability to work well with little hyperparameter tuning, relatively low memory
requirements, and its ability to smooth the steps of gradient descent using momentum. Monitoring of valida-
tion loss, i.e. early stopping criteria, was employed instead of a fixed number of epochs to allow both models the
necessary epochs to converge while keeping the risk of overfitting to the training data low. The patience criteria
for validation loss convergence was set to 15 epochs to allow for sufficient certainty that the model had converged
and was unlikely to meaningfully improve. The weight decay was set to 1­ e−5 following previous optimization
­work87. The CNNs were implemented with ­TensorFlow90 and the Keras ­API88 and model training was performed
using an NVIDIA Titan V.

Scientific Reports | (2021) 11:8172 | https://doi.org/10.1038/s41598-021-87557-5 8

Vol:.(1234567890)
www.nature.com/scientificreports/

Diffraction pattern classification. Each diffraction pattern collected, but not used in training (> 140,000
images), was evaluated in a random order by the corresponding trained CNN model without further informa-
tion. The output classification of each diffraction pattern was recorded, saved in a (.csv) file, and tabulated. Preci-
sion and Recall were calculated for each material and each space group using Scikit-learn91. Precision (Eq. 1) for
each class (e.g. 230) is defined as the number of correctly predicted images out of all photos predicted to belong
to that class (e.g. 230). Recall (Eq. 2) is the number of correctly predicted images for each class divided by the
actual number of images for the class. F1-score is the weighted harmonic of the Precision and Recall and is par-
ticularly valuable in situations where the number of test images per class is variable. A high F1-score means the
model has low false positives and low false negatives.
true positives
Precision = , (1)
true positives + false positives

true positives
Recall = . (2)
true positives + false negatives

Neural network insight. Comparisons between the resultant models from pseudo RGB transfer learning
and training the model from scratch offer an opportunity to understand how the CNNs go about their given task.
Visualization techniques are implemented using the keras-vis package v0.4.177. Filters from the first layer of each
model are extracted and plotted as 3 × 3 matrices with m ­ atplotlib92. The first layer was targeted since the earliest
filters represent lower level features such as colors and edges. The Euclidean distance between the individual fil-
ter arrays in each model was computed using the NumPy linear algebra toolbox to compute L2 n ­ orms93 and the
four most similar learned filters between the two models were identified. The outputs (a.k.a. feature maps) of the
first layer corresponding to these four filters are also extracted to examine the activations of the two approaches.
The feature maps from earlier convolution layers are more useful since deeper layers operate in feature space and
are therefore more difficult to ­understand29,70. Lastly, Shapley values, deeply rooted in game ­theory94,95, are esti-
mated using the DeepSHAP tools in S­ HAP96. SHAP uses a distribution of background samples, approximates the
model with a linear function between each background data sample and the current input to be explained, and
assumes the input features are independent to compute approximate SHAP values. The sum of the SHAP values
equals the difference between the expected model output (averaged over the background dataset) and the cur-
rent model output. One hundred images per space group were used as background samples in conjunction with
one new input image per space group. The total number of diffraction patterns used as a background follows the
SHAP software protocols. When plotted as an overlay, red pixels represent positive SHAP values that increase the
probability of the class, while blue pixels represent negative SHAP values that reduce the probability of the class.
An example of SHAP analysis on handwritten digits from the MNIST ­database97 is shown in Supplementary
Fig. S2. For a given image, the presence and absence of features that positively correlate with a class are shown in
red, while negative correlations are shown in blue. As an example, in the image of a ‘four’ the lack of a connec-
tion on top makes it a four instead of a nine. Combined, these insights into the operations of the neural network
can further substantiate the validity of the transfer learning approach, increase trust by better understanding the
model’s methods, and provide indications in cases where the model is incorrect about future predictions.

Data availability
The datasets generated during and/or analyzed during the current study are available from the corresponding
author on reasonable request. The code will be available in the GitHub repository https://​github.​com/​krkau​
fma/​EBSD_​trans​fer_​learn​ing and in the online Zenodo repository (https://​doi.​org/​10.​5281/​zenodo.​35649​37)98.

Received: 28 November 2020; Accepted: 31 March 2021

References
1. O’Mara, J., Meredig, B. & Michel, K. Materials data infrastructure: A case study of the citrination platform to examine data import,
storage, and access. JOM 68, 2031–2034 (2016).
2. Ward, L. et al. Matminer: An open source toolkit for materials data mining. Comput. Mater. Sci. 152, 60–69 (2018).
3. Mrdjenovich, D. et al. propnet: A knowledge graph for materials science. Matter 2, 464–480 (2020).
4. Tabor, D. P. et al. Accelerating the discovery of materials for clean energy in the era of smart automation. Nat. Rev. Mater. 3, 5–20
(2018).
5. Ong, S. P. et al. Python materials genomics (pymatgen): A robust, open-source python library for materials analysis. Comput.
Mater. Sci. 68, 314–319 (2013).
6. Saal, J. E., Kirklin, S., Aykol, M., Meredig, B. & Wolverton, C. Materials design and discovery with high-throughput density func-
tional theory: The open quantum materials database (OQMD). JOM 65, 1501–1509 (2013).
7. Jha, D. et al. ElemNet: Deep learning the chemistry of materials from only elemental composition. Sci. Rep. 8, 17593 (2018).
8. Oviedo, F. et al. Fast and interpretable classification of small X-ray diffraction datasets using data augmentation and deep neural
networks. NPJ Comput. Mater. 5, 1–9 (2019).
9. Kaufmann, K. et al. Discovery of high-entropy ceramics via machine learning. NPJ Comput. Mater. 6, 42 (2020).
10. DeCost, B. L. & Holm, E. A. A computer vision approach for automated analysis and classification of microstructural image data.
Comput. Mater. Sci. 110, 126–133 (2015).
11. McAuliffe, T. P. et al. Advancing characterisation with statistics from correlative electron diffraction and X-ray spectroscopy, in
the scanning electron microscope. Ultramicroscopy 211, 112944 (2020).
12. Kaufmann, K. & Vecchio, K. S. Searching for high entropy alloys: A machine learning approach. Acta Mater. 198, 178–222 (2020).

Scientific Reports | (2021) 11:8172 | https://doi.org/10.1038/s41598-021-87557-5 9

Vol.:(0123456789)
www.nature.com/scientificreports/

13. Qiao, Z., Welborn, M., Anandkumar, A., Manby, F. R. & Miller, T. F. OrbNet: Deep learning for quantum chemistry using symmetry-
adapted atomic-orbital features. J. Chem. Phys. 153, 124111 (2020).
14. Holm, E. A. In defense of the black box. Science (80-). 364, 26–27 (2019).
15. Adadi, A. & Berrada, M. Peeking inside the black-box: A survey on explainable artificial intelligence (XAI). IEEE Access 6,
52138–52160 (2018).
16. Kaufmann, K. et al. Crystal symmetry determination in electron diffraction using machine learning. Science (80-). 367, 564–568
(2020).
17. Foden, A., Previero, A. & Britton, T. B. Advances in electron backscatter diffraction. Preprint at http://​arxiv.​org/​abs/​1908.​04860
(2019).
18. Ziletti, A., Kumar, D., Scheffler, M. & Ghiringhelli, L. M. Insightful classification of crystal structures using deep learning. Nat.
Commun. 9, 2775 (2018).
19. Modarres, M. H. et al. Neural network for nanoscience scanning electron microscope image recognition. Sci. Rep. 7, 1–12 (2017).
20. Stan, T., Thompson, Z. T. & Voorhees, P. W. Optimizing convolutional neural networks to perform semantic segmentation on large
materials imaging datasets: X-ray tomography and serial sectioning. Mater. Charact. 160, 110119 (2020).
21. DeCost, B. L., Lei, B., Francis, T. & Holm, E. A. High throughput quantitative metallography for complex microstructures using
deep learning: A case study in ultrahigh carbon steel. Microsc. Microanal. 25, 21–29 (2019).
22. Roberts, G. et al. Deep learning for semantic segmentation of defects in advanced STEM images of steels. Sci. Rep. 9, 1–12 (2019).
23. de Haan, K., Ballard, Z. S., Rivenson, Y., Wu, Y. & Ozcan, A. Resolution enhancement in scanning electron microscopy using deep
learning. Sci. Rep. 9, 1–7 (2019).
24. Xie, T. & Grossman, J. C. Crystal graph convolutional neural networks for an accurate and interpretable prediction of material
properties. Phys. Rev. Lett. 120, 145301 (2018).
25. Spurgeon, S. R. et al. Towards data-driven next-generation transmission electron microscopy. Nat. Mater. https://​doi.​org/​10.​1038/​
s41563-​020-​00833-z (2020).
26. Rosenstein, M. T., Marx, Z., Kaelbling, L. P. & Dietterich, T. G. To transfer or not to transfer. In Neural Information Processing
Systems (NIPS ’05) Workshop Inductive Transfer: 10 Years Later (2005).
27. Jha, D. et al. Enhancing materials property prediction by leveraging computational and experimental data using deep transfer
learning. Nat. Commun. 10, 1–12 (2019).
28. Thompson, J. A. F., Schonwiesner, M., Bengio, Y. & Willett, D. How transferable are features in convolutional neural network
acoustic models across languages? In ICASSP, IEEE International Conference on Acoustics, Speech and Signal Processing—Proceed-
ings 2019-May, 2827–2831 (Institute of Electrical and Electronics Engineers Inc., 2019).
29. LeCun, Y., Bengio, Y. & Hinton, G. Deep learning. Nature 521, 436–444 (2015).
30. Pan, X. et al. Multi-task Deep learning for fine-grained classification/grading in breast cancer histopathological images. In Studies
in Computational Intelligence 810, 85–95 (Springer, 2020).
31. Xie, Y. & Richmond, D. Pre-training on grayscale imagenet improves medical image classification. In The European Conference on
Computer Vision (ECCV) Workshops 11134 (Springer, 2018).
32. Gonzalez, J., Bhowmick, D., Beltran, C., Sankaran, K. & Bengio, Y. Applying knowledge transfer for water body segmentation in
Peru. Preprint at http://​arxiv.​org/​abs/​1912.​00957 (2019).
33. Pan, S. J. & Yang, Q. A survey on transfer learning. IEEE Trans. Knowl. Data Eng. 22, 1345–1359 (2010).
34. Deng, J. et al. ImageNet: A large-scale hierarchical image database. In IEEE Conference on Computer Vision and Pattern Recognition
248–255 (Institute of Electrical and Electronics Engineers (IEEE), 2010). https://​doi.​org/​10.​1109/​cvpr.​2009.​52068​48
35. Bloom, P. How Children Learn the Meanings of Words (MIT Press, 2000).
36. Felzenszwalb, P. F. & Huttenlocher, D. P. Pictorial structures for object recognition. Int. J. Comput. Vis. 61, 55–79 (2005).
37. Fei-Fei, L., Fergus, R. & Perona, P. One-shot learning of object categories. IEEE Trans. Pattern Anal. Mach. Intell. 28, 594–611
(2006).
38. Lake, B. M., Salakhutdinov, R. & Tenenbaum, J. B. Human-level concept learning through probabilistic program induction. Science
(80-). 350, 1332–1338 (2015).
39. Zhang, R., Che, T., Ghahramani, Z., Bengio, Y. & Song, Y. MetaGAN: An Adversarial Approach to Few-Shot Learning. In NeurIPS
2018 2365–2374 (2018).
40. Liu M.-Y. et al. Few-shot unsupervised image-to-image translation. In Proceedings of the IEEE/CVF International Conference on
Computer Vision (2019).
41. Li, H., Kadav, A., Durdanovic, I., Samet, H. & Graf, H. P. Pruning filters for efficient ConvNets. In 5th Int. Conf. Learn. Represent.
1–13 (2016).
42. Guo, Z. et al. Single path one-shot neural architecture search with uniform sampling. Preprint at http://​arxiv.​org/​abs/​1904.​00420
(2019).
43. Schwartz, A. J., Kumar, M., Adams, B. L. & Field, D. P. Electron Backscatter Diffraction in Materials Science (Springer
Science+Business Media, LLC, 2009). https://​doi.​org/​10.​1007/​978-0-​387-​88136-2.
44. Vecchio, K. S. & Williams, D. B. Convergent beam electron diffraction study of ­Al3Zr in Al–Zr AND Al–Li–Zr alloys. Acta Metall.
35, 2959–2970 (1987).
45. Vecchio, K. S. & Williams, D. B. Convergent beam electron diffraction analysis of theT 1 (­ Al2CuLi) phase in Al–Li–Cu alloys.
Metall. Trans. A 19, 2885–2891 (1988).
46. Tong, V. S., Knowles, A. J., Dye, D. & Britton, T. B. Rapid electron backscatter diffraction mapping: Painting by numbers. Mater.
Charact. 147, 271–279 (2019).
47. Thomsen, K., Schmidt, N. H., Bewick, A., Larsen, K. & Goulden, J. Improving the accuracy of orientation measurements using
EBSD. Microsc. Microanal. 19, 724–725 (2013).
48. Zhu, C., Kaufmann, K. & Vecchio, K. S. Novel remapping approach for HR-EBSD based on demons registration. Ultramicroscopy
208, 112851 (2020).
49. Lassen, N. C. K. Automated Determination of Crystal Orientations from Electron Backscattering Patterns (The Technical University
of Denmark, 1994).
50. Britton, T. B. et al. Factors affecting the accuracy of high resolution electron backscatter diffraction when using simulated patterns.
Ultramicroscopy 110, 1443–1453 (2010).
51. Hielscher, R., Bartel, F. & Britton, T. B. Gazing at crystal balls: Electron backscatter diffraction pattern analysis and cross correlation
on the sphere. Ultramicroscopy 207, 112836 (2019).
52. Foden, A., Collins, D. M., Wilkinson, A. J. & Britton, T. B. Indexing electron backscatter diffraction patterns with a refined template
matching approach. Ultramicroscopy 207, 112845 (2019).
53. Karthikeyan, T., Dash, M. K., Saroja, S. & Vijayalakshmi, M. Evaluation of misindexing of EBSD patterns in a ferritic steel. J.
Microsc. 249, 26–35 (2013).
54. Chen, C. L. & Thomson, R. C. The combined use of EBSD and EDX analyses for the identification of complex intermetallic phases
in multicomponent Al–Si piston alloys. J. Alloys Compd. 490, 293–300 (2010).
55. McLaren, S. & Reddy, S. M. Automated mapping of K-feldspar by electron backscatter diffraction and application to 40Ar/39Ar
dating. J. Struct. Geol. 30, 1229–1241 (2008).

Scientific Reports | (2021) 11:8172 | https://doi.org/10.1038/s41598-021-87557-5 10

Vol:.(1234567890)
www.nature.com/scientificreports/

56. Ram, F. & De Graef, M. Phase differentiation by electron backscatter diffraction using the dictionary indexing approach. Acta
Mater. 144, 352–364 (2018).
57. Chen, Y. H. et al. A dictionary approach to electron backscatter diffraction indexing. Microsc. Microanal. https://​doi.​org/​10.​1017/​
S1431​92761​50007​56 (2015).
58. Singh, S. et al. High resolution low kV EBSD of heavily deformed and nanocrystalline Aluminium by dictionary-based indexing.
Sci. Rep. https://​doi.​org/​10.​1038/​s41598-​018-​29315-8 (2018).
59. Ram, F., Wright, S., Singh, S. & De Graef, M. Error analysis of the crystal orientations obtained by the dictionary approach to EBSD
indexing. Ultramicroscopy 181, 17–26 (2017).
60. Day, A. P. Spherical EBSD. J. Microsc. https://​doi.​org/​10.​1111/j.​1365-​2818.​2008.​02011.x (2008).
61. Lenthe, W. C., Singh, S. & Graef, M. D. A spherical harmonic transform approach to the indexing of electron back-scattered dif-
fraction patterns. Ultramicroscopy 207, 112841 (2019).
62. Zhu, C., Kaufmann, K. & Vecchio, K. Automated reconstruction of spherical Kikuchi maps. Microsc. Microanal. https://​doi.​org/​
10.​1017/​S1431​92761​90007​10 (2019).
63. McAuliffe, T. P., Dye, D. & Britton, T. B. Spherical-angular dark field imaging and sensitive microstructural phase clustering with
unsupervised machine learning. Ultramicroscopy 219, 113132 (2020).
64. Nowell, M. M. & Wright, S. I. Phase differentiation via combined EBSD and XEDS. J. Microsc. 213, 296–305 (2004).
65. Goehner, R. P. & Michael, J. R. Phase identification in a scanning electron microscope using backscattered electron Kikuchi pat-
terns. J. Res. Natl. Inst. Stand. Technol. 101, 301–308 (1996).
66. Dingley, D. J. & Wright, S. I. Phase identification through symmetry determination in EBSD patterns. In Electron Backscatter
Diffraction in Materials Science (eds. Schwartz, A., Kumar, M., Adams, B. & Field, D.) 97–107 (Springer, 2009). https://​doi.​org/​10.​
1007/​978-0-​387-​88136-2
67. Li, L. & Han, M. Determining the Bravais lattice using a single electron backscatter diffraction pattern. J. Appl. Crystallogr. https://​
doi.​org/​10.​1107/​S1600​57671​40259​89 (2015).
68. Kaufmann, K., Zhu, C., Rosengarten, A. S. & Vecchio, K. S. Deep neural network enabled space group identification in EBSD.
Microsc. Microanal. 26, 447–457 (2020).
69. Kaufmann, K. et al. Phase Mapping in EBSD using convolutional neural networks. Microsc. Microanal. 26, 458–468 (2020).
70. Ding, Z., Pascal, E. & De Graef, M. Indexing of electron back-scatter diffraction patterns using a convolutional neural network.
Acta Mater. 199, 370–382 (2020).
71. Callahan, P. G. & De Graef, M. Dynamical electron backscatter diffraction patterns. Part I: Pattern simulations. Microsc. Microanal.
19, 1255–1265 (2013).
72. Explainable AI: Interpreting, Explaining and Visualizing Deep Learning (eds. Samek, W., Montavon, G., & Vedaldi, A.) 11700,
(Springer International Publishing, 2019)
73. Carter, S., Armstrong, Z., Schubert, L., Johnson, I. & Olah, C. Activation atlas. Distill 4, e15 (2019).
74. Graff, C. A. & Ellen, J. Correlating Filter Diversity with Convolutional Neural Network Accuracy 75–80 (Institute of Electrical and
Electronics Engineers (IEEE), 2017). https://​doi.​org/​10.​1109/​icmla.​2016.​0021.
75. Kanopoulos, N., Vasanthavada, N. & Baker, R. L. Design of an image edge detection filter using the sobel operator. IEEE J. Solid-
State Circuits 23, 358–367 (1988).
76. Selvaraju, R. R. et al. Grad-CAM: Visual explanations from deep networks via gradient-based localization. In Proceedings of the
IEEE International Conference on Computer Vision 2017, 618–626 (2017).
77. Kotikalapudi, R. keras-vis. https://​github.​com/​ragha​kot/​keras-​vis (2017).
78. Ribeiro, M. T., Singh, S. & Guestrin, C. ‘Why Should I Trust You?’ Explaining the Predictions of Any Classifier. (2016). https://​doi.​
org/​10.​1145/​29396​72.​29397​78
79. Lundberg, S. M., Allen, P. G. & Lee, S.-I. A Unified approach to interpreting model predictions. In Advances in Neural Information
Processing Systems 30 (eds. Guyon, I. et al.) 4765–4774 (Curran Associates, Inc., 2017).
80. Chen, H., Janizek, J. D., Lundberg, S. & Lee, S.-I. True to the Model or True to the Data? Preprint at http://​arxiv.​org/​abs/​2006.​
16234 (2020).
81. Ghorbani, A. & Zou, J. Data shapley: Equitable valuation of data for machine learning. In 36th International Conference on Machine
Learning, ICML 2019 2019, 4053–4065 (International Machine Learning Society (IMLS), 2019).
82. Wu, M., Wicker, M., Ruan, W., Huang, X. & Kwiatkowska, M. A game-based approximate verification of deep neural networks
with provable guarantees. Theor. Comput. Sci. 807, 298–329 (2020).
83. Winkelmann, A., Britton, T. B. & Nolze, G. Constraints on the effective electron energy spectrum in backscatter Kikuchi diffrac-
tion. Phys. Rev. B 99, 064115 (2019).
84. Yuanyuan, C. & Zhibin, W. Quantitative analysis modeling of infrared spectroscopy based on ensemble convolutional neural
networks. Chemom. Intell. Lab. Syst. 181, 1–10 (2018).
85. Blundell, C., Cornebise, J., Kavukcuoglu, K. & Wierstra, D. Weight uncertainty in neural networks. In 32nd Int. Conf. Mach. Learn.
ICML 2015 2, 1613–1622 (2015).
86. Hinton, G., Vinyals, O. & Dean, J. Distilling the Knowledge in a Neural Network. Preprint at http://​arxiv.​org/​abs/​1503.​02531
(2015).
87. Chollet, F. Xception: Deep learning with depthwise separable convolutions. In Proceedings of the IEEE International Conference
on Computer Vision 1251–1258 (The Computer Vision Foundation, 2017).
88. Chollet, F. K. (2015). https://​github.​com/​keras-​team/​keras.
89. Kingma, D. P. & Ba, J. L. Adam: A method for stochastic optimization. In 3rd International Conference on Learning Representations,
ICLR 2015—Conference Track Proceedings 1–15 (2015).
90. Abadi, M. et al. TensorFlow: A system for large-scale machine learning. In Proceedings of the 12th USENIX Symposium on Operating
Systems Design and Implementation 265–283 (USENIX, 2016).
91. Pedregosa, F. et al. Scikit-learn: Machine learning in python. J. Mach. Learn. Res. 12, 2825–2830 (2011).
92. Hunter, J. D. Matplotlib: A 2D graphics environment. Comput. Sci. Eng. 9, 99–104 (2007).
93. Van Der Walt, S., Colbert, S. C. & Varoquaux, G. The NumPy array: A structure for efficient numerical computation. Comput. Sci.
Eng. 13, 22–30 (2011).
94. Shapley, L. S. Stochastic games. Proc. Natl. Acad. Sci. U. S. A. 39, 1095–1100 (1953).
95. Shapley, L. S. A value for n-person games. In Contributions to the Theory of Games (AM-28), Volume II (eds. Kuhn, H. W. & Tucker,
A. W.) 307–317 (Princeton University Press, 1953).
96. Lundberg, S. M. & Lee, S.-I. A Unified Approach to Interpreting Model Predictions. In Advances in Neural Information Processing
Systems Volume 30, 4765–4774 (2017).
97. LeCun, Y., Bottou, L., Bengio, Y. & Haffner, P. Gradient-based learning applied to document recognition. Proc. IEEE 86, 2278–2323
(1998).
98. krkaufma. krkaufma/Electron-Diffraction-CNN v1.0.1. (2019). Code available at https://​doi.​org/​10.​5281/​ZENODO.​35649​37.

Acknowledgements
The authors would like to thank William M. Mellor for composing Supplementary Table S3.

Scientific Reports | (2021) 11:8172 | https://doi.org/10.1038/s41598-021-87557-5 11

Vol.:(0123456789)
www.nature.com/scientificreports/

Author contributions
K.K. and H.L. co-designed the initial project scope and aims. K.K. performed the bulk of the programming
work with guidance from H.L. K.K. drafted the early versions of the manuscript and figures. X.L. performed
the experimental work for the ­Ni90Al10 sample. K.S.V. guided the development of the project and reviewed and
revised the early drafts of the manuscript. All authors participated in analyzing and interpreting the final data
and contributed to the discussions and revisions of the manuscript.

Funding
Supported by the U.S. Department of Defense (DoD) [through the National Defense Science and Engineering
Graduate Fellowship (NDSEG) Program] and the ARCS Foundation, San Diego Chapter (K.K.); and the Oer-
likon Group (K.S.V.).

Competing interests
The authors declare no competing interests.

Additional information
Supplementary Information The online version contains supplementary material available at https://​doi.​org/​
10.​1038/​s41598-​021-​87557-5.
Correspondence and requests for materials should be addressed to K.S.V.
Reprints and permissions information is available at www.nature.com/reprints.
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional affiliations.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International
License, which permits use, sharing, adaptation, distribution and reproduction in any medium or
format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the
Creative Commons licence, and indicate if changes were made. The images or other third party material in this
article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the
material. If material is not included in the article’s Creative Commons licence and your intended use is not
permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder. To view a copy of this licence, visit http://​creat​iveco​mmons.​org/​licen​ses/​by/4.​0/.

© The Author(s) 2021

Scientific Reports | (2021) 11:8172 | https://doi.org/10.1038/s41598-021-87557-5 12

Vol:.(1234567890)

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy