Next Article in Journal
Marine GIS as a Tool to Support Backscatter Data Analysis for Zooplankton Investigations
Next Article in Special Issue
VFR: The Underwater Acoustic Target Recognition Using Cross-Domain Pre-Training with FBank Fusion Features
Previous Article in Journal
Development and Testing of a High-Resolution Three-Dimensional Seismic Detection System for Gas Hydrate
Previous Article in Special Issue
Experimental Study on the Target–Receiver Formation Problem with the Exploitation of Coherent and Non-Coherent Bearing Information
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Convolutional Autoencoding of Small Targets in the Littoral Sonar Acoustic Backscattering Domain

1
Department of Electrical and Computer Engineering, University of Iowa, Iowa City, IA 52242, USA
2
Panama City Division, Naval Surface Warfare Center, Panama City, FL 32407, USA
*
Author to whom correspondence should be addressed.
J. Mar. Sci. Eng. 2023, 11(1), 21; https://doi.org/10.3390/jmse11010021
Submission received: 1 November 2022 / Revised: 2 December 2022 / Accepted: 10 December 2022 / Published: 23 December 2022
(This article belongs to the Special Issue Application of Sensing and Machine Learning to Underwater Acoustic)

Abstract

:
Automated target recognition is an important task in the littoral warfare domain, as distinguishing mundane objects from mines can be a matter of life and death. This is initial work towards the application of convolutional autoencoding to the littoral sonar space, with goals of disentangling the reflection noise prevalent in underwater acoustics and allowing recognition of the shape and material of targets. The autoencoders were trained on magnitude Fourier transforms of the TREX13 dataset. Clusters in the encoding space representing the known variable of measurement distance between the target and the sensor were found. An encoding vector space of around 16 dimensions appeared sufficient, and the space was shown to generalize well to unseen data.

1. Introduction

The oceanic acoustic environment poses many challenges to autonomous target characterization not present in other signaling domains. Reflections off the surface of the ocean and the seafloor, as well as clutter like fish, rocks and bubbles add noise and echoes that greatly impede clarity [1]. Robustly defining and selecting features for underwater signals is necessary for automatic target recognition tasks [2,3,4,5,6].
We want to apply techniques from the deep learning community [7] and provide new methods of target representation and analysis, specifically employing the various types of autoencoders. Deep learning has had recent successes in the underwater sonar domain such as in [8] where it was used to automate sonar processing.
Our end goal is to develop a model that can robustly provide insight into acoustic backscattering returns for object detection with application to mine detection [9]. This would at minimum constitute object detection and ordinance classification regardless of noise, to be utilized in conjunction with an autonomous underwater vehicle [10,11].
The goal of informing object classification, also called automated target recognition (ATR), via geometric characteristics detected in the acoustic signal is of particular interest for maritime search [12,13,14,15]. Unlike other applications such as natural language processing, the nature of the maritime environment causes the cost of data collection for ATR training purposes to be extraordinarily high, involving days of on-water tests to collect a limited set of sonar returns on relatively similar targets within a target field [16]. The high cost of data collection necessitates other methods for data collection, including the development of simulated data [17], aggregation of heterogeneous datasets [18], and physics-based parameter characterisation during the ATR process [19,20]. Among these data-centric approaches, physics-based methods show considerable promise due to the direct involvement of the model in the learning process.
In maritime applications, it is often the case that while acoustic imagery is unavailable, engineering parameters related to a target’s shape, size, and other characteristics are known. The interplay between these physical traits and the sonar returns allows for the potential to train algorithms to detect these physical features with increased accuracy for a given amount of training data compared to other methods. This is due to the direct involvement of the physical characteristics without secondary approximations that may occur when training on simulated data.
The key concept to be applied to the acoustic backscattering data is the autoencoder. One of the earliest uses of an autoencoder was [21] in 1987 as documented in [22]. Many variants and extensions of this concept have developed since then, but the core idea remains. A network is trained in an unsupervised manner such that inputs of high dimension are reduced to lower dimensional encodings akin to the classic principal component analysis (PCA) technique.
This key dimensional advantage provides the background motivation for adopting the autoencoder for compact yet effective feature representation of sonar backscattering from small targets such as considered in this work. Lower-dimensional feature engineering are inherently less vulnerable to unstructured noise from the sonar environment, and also provide a computational advantage towards eventual target classification. The backscattering data is a two dimensional signal domain, so our autoencoders are convolutional by construction, with an aim to provide smoothly connected lower-dimensional feature encoding of the targets of interest. The main disadvantage of an autoencoder is that it is a physics-agnostic and purely computational tool. Therefore, it is unable to provide direct physical interpretation of the individual values in the encodings that the network produces.
The scope of this work is constrained to explore the representational power of the autoencoder in the context of real sonar targets over field data. In this context, given the noise suppression and computational advantages inherent in lower-dimensional feature encoding, this lack of immediate physical interpretation does not pose a fundamental limitation towards successful feature encoding and subsequent reconstruction. Therefore, the compact representational advantage makes the autoencoder a useful tool towards lower-dimensional feature engineering, and in future work we aim to explore the physical meaning of autoencoded features using extensive simulations of a variety of sonar targets under diverse environmental conditions.
The key contributions of this work are as follows.
(i)
We take the preexisting convolutional autoencoder network concept [23] and apply it to a different data domain: underwater sonar backscattering from small elastic targets.
(ii)
We approximate the dimensionality of a feature space for this new data through empirical analysis, and then
(iii)
visualize the multimodal network reconstruction error distributions collected through 5-fold cross-validation.
This enables feature interpretation based on the autoencoder framework in terms of potentially lower dimensional representations.
In future work, we will aim for encoding vector interpretability, similar to how the InfoGAN [24] works. InfoGAN is a type of generative adversarial network (GAN), similar in concept to an inverted autoencoder [25,26], that generates data from an interpretable vector of variables with defined units, and a noise vector. Out goal is disentangling a noise vector from an interpretable coding vector that would contain object class as well as useful information that could include distance from the target, measurement angle, and ground sediment type. Additionally, we may branch off to use such results to generate underwater acoustic signal in a different way from existing techniques such as in [27].

2. Materials and Methods

2.1. Principle Component Analysis

Autoencoders take inspiration from the process of principal component analysis (PCA) [28], a method by which data dimensionality is reduced. Given a set of data vectors, one finds a a linear projection to a lower dimensional space, or principle subspace, while maximizing the variance of the projected data. This is useful for data compression and feature extraction [29,30].
Given a distribution of points in a high dimensional vector space, the eigenvectors of the covariance matrix form a new set of perpendicular components, with the highest magnitude eigenvalue associated with the most principle component, and in sequence of magnitudes beyond that. To perform dimensionality reduction, the component space is projected onto a lower dimension principle subspace that is created by removing the least principle components.

2.2. Autoencoders

Autoencoding networks learn a low dimensional subspace to the space of the input data. Unlike with PCA, the components of the encoding are not orthogonal and linear vectors, but rather nonlinear functions on an input and the network’s trained weights [30]. The goal is not perfect reconstruction; instead important characteristics are ideally emphasized by reducing the inputs to a tight bottleneck [23].
An autoencoder is a neural network comprised of two subnetworks in sequence. The first subnetwork is called the encoder, and performs dimension reduction, resulting in the encoding. This encoding is then fed into the second subnetwork, referred to as a decoder, and it attempts to reverse the effects of the encoder.
The encoder learns a mapping from the input data space to a low dimension vector space, Equation (1), which is the encoding. In our use case, the input space is a set of two dimensional signals that can be represented as M × N matrices of real values. The matrices are fed through a series of decimating convolutional layers separated with regularization and nonlinear activation layers. The result of the final convolution has a regularization and nonlinear activation applied to it before it is reshaped into a vector. This vector has a dense neuron layer applied to it to produce the encoding.
E : R M × N R k
The decoder attempts to reverse the mapping performed by the encoder, Equation (2), with a sequence of dilating transposed convolutional layers to reconstruct the input data. The structure is effectively an inverse of the encoder’s: a dense layer followed by regularization and activation and then the transposed convolutional layers with the same separation layers. The final transposed convolutional layer produces the reconstruction attempt.
D : R k R M × N
It is in practice impossible to have D = E 1 because of the lossy mapping, so this process finds the most important features of the input space.
PCA performed on input images yields component images that can be linearly combined while decoding, but the components of autoencoders are not orthogonal, so this cannot be mimicked.
The training process of autoencoders is the same as with other neural networks. Back-propagation and stochastic gradient descent are utilized to minimize a loss function. Batch normalization layers are utilized to reduce generalization error [31] and constrain the values between the network layers [32]—to reduce the likelihood of exploding gradients.
The loss function for an autoencoder is related to how the error between the input and output is evaluated. The sum of the element-wise mean squared error (MSE) between the input and reconstruction is a common choice for the error metric in autoencoding.
MSE x , D E x = i = 1 M j = 1 N x i j D E x i j
Signal-like data is typically passed through convolutional neuron layers in neural networks instead of the more general-case fully connected layers that are tantamount to matrix multiplication [33].
The autoencoder can be called “sparse” if a sparsity constraint is applied to the values of the encoding vector [34]. Sparse autoencoders are used to find interpretable features for in other tasks [23]. It penalizes the loss function if the expected value an individual component of the encoding diverges from a small value, with the expected value estimated for each training batch. KL-Divergence is only defined for probability values, so the encoding space must be restricted to values between 0 and 1.
A popular variant of autoencoder is the variational autoencoder (VAE), which, instead of an encoding vector, produces a set of independent Gaussian distributions [35]. We looked at VAEs but have not produced any significant results.

2.3. Data

The datasets used in the following experiments—TREX13, PondEx09, PondEx10—were collected using the mechanism described in [36]. A tower containing sensors and electronics moved along a 20 m rail system. The electronics included 10 cm receiver, at 5 cm/s to allow for Synthetic Aperture Sonar processing of acoustic backscattering measurements. The tower transmitted a 6 ms pulse sliding from 1 to 30 kHz twice per second as it moved along the rail. The beam width was greater than 40 . The receiver had 6 vertically aligned apertures used to determine the return angle of the backscattering.
The data collection occurred at distances of 5, 10, 15, 20, 25, 30, 35 and 40 m from the center of the rail for a variety of targets at a minimum of nine orientations for each target, 80 to 80 in 20 increments. A few targets have data for the remaining 100 to 100 orientations as well.
The TREX13 dataset [16] has data for 27 diverse targets with a total of 810 time/cross range measurements. The PondEx datasets [37] total 117 time/cross range measurements for 13 targets.
The measurements are 2D arrays of floating point values, with a time axis and a position axis associated with measurements along the 20 m rail. When visualized, the backscattering response takes a shape similar to that of a smile, as shown in Figure 1, so the data is called smile data. The smile data has very large array size, but the smiles generally fit within a 1024 × 1024 element bounding box, so it is cropped to that size. 1024 is a power of 2, allowing up to 10 decimations in convolutional layers. The cropped data has its Fourier transform calculated along the time axis to produce raw frequency data.
The network implementations only handle real-valued data, so the data excluded the phase information, leaving the magnitude only. The magnitude of the frequency domain of a real signal is always symmetrical, so the training data only used half of the frequency data, leaving 512 (frequency) by 1024 (position) sized data arrays.
Visualization of magnitude frequency data is typically done in decibel form, so we also used this representation of the data as a training set, but the results were not useful.

2.4. Networks

We used the Python libraries Tensorflow [38] and Keras [39] to implement and train our networks through standard back propagation [40], using the Keras default for many metavariables. Further refinement of the metavariables is for future works.
We chose the Parametric ReLU (PReLU) [41] nonlinearity as our activation between convolutional layers as opposed to the more typical Rectified Linear Unit (ReLU) and Leaky ReLU [42]. PReLU is similar to Leaky ReLU, but the slope of the response to negative values is a trainable parameter–thus any activation can learn any range of activations between a typical ReLU to an identity, and beyond in both directions, as the problem needs.
The Glorot Uniform [43] weight initializer and the standard Adam optimizer [44], both with default parameters excluding learning rate, set and refined the learnable values weights.
The networks themselves are convolutional autoencoders acting upon 1024 × 512 input images, wherein an image in this application domain is described in Section 2.3. Our encoders have seven convolutional layers that each have 4 × 4 kernels, a stride factor of 2, and a batch normalization before a PReLU as the nonlinear activation.
The first layer reduces the 1024 × 512 input to a single channel 512 × 256 intermediate image, and subsequent layers each double the number of channels from their input to output.
After the seven convolutional layers, the 64 channels of 8 × 4 images are unwrapped into a single 2048 element vector. A neuron layer reduces this vector to a vector of the desired latent dimension. In the case of the basic autoencoder, there is no no-linearity applied to this layer, but for the sparse autoencoder a sigmoid function is applied to constrain the values between 0 and 1.
The decoder is close to an inverse network structure, with batch normalization followed by a PReLU functioning as a unit. It starts with a neuron layer that has batch normalization and a PReLU after it. The output of the neuron layer is reshaped into 64 channels of 8 × 4 images and fed into a sequence of transposed convolutional layers. The transposed convolutional layers have 4 × 4 kernels and a stride of 2 like in the encoder, and also have batch normalization followed by PReLU activations. The layers halve the number of channels from their input to output, except for the final layer which maintains a single channel. This final layer does not have batch normalization or PReLU applied to its output. The full network can be seen in Figure 2.
The integer ratio between the stride and kernel size is meant to reduce checkerboard noise patterns in the network output, as explored in [45]. Batch normalization is used increase the allowable learning rate magnitudes for training.

2.5. Training Process

The general training process is as follows:
  • Preprocess the data as described in Section 2.3.
  • Initialize a new autoencoder with random weights and the correct parameters.
  • Begin training the autoencoder on the full data set for 100 epochs (unless otherwise specified), using MSE loss.
  • Ensure that the loss value is beneath a threshold after one third of the training epochs have completed. Cancel the training and return to Step 2 if it is not.
  • Save the final weights of the trained network.
  • Repeat from Step 2 until ten sets of weights are saved.
When performing cross-validation, the process is slightly different. Rather than train on the full data set, five folds of the data set are isolated and the process from Step 2 and onward is performed on the each of the five complementary sets to the excluded folds. So, for each of the folds, the data not in the fold is used to train a network ten times.

2.6. Evaluation

To obtain an encoding, an input is put through just the encoder half of the network. The decoder does not provide any use in this case.
Reconstruction accuracy across a reserved fold in cross-validation is estimated by calculating MSE between each data element in the fold and the reconstruction the network produces for it. Analysis is performed on the reconstruction accuracies across the whole fold.

3. Results

3.1. Learning Rate Determination

The optimal learning rate of our network and optimizer combination was unknown, so we tested a set of learning rates with various latent space dimensions. The learning rate tested were the set 1.0 , 0.1 , , 10 6 was and the latent space dimensions were 2 , 4 , 6 , , 16 . A comparison of the results of training five networks for 50 epochs each with each combination of the sets facilitated a choice.
As seen in Figure 3, the rate 0.001 shows the earliest descents of the loss, so we used it for the rest of the experiments. The difference between the basic autoencoder and the sparse autoencoder is small enough that the 0.001 learning rate is still adequate.
In the data for the chosen learning rate, there is a noticeable poor performing level of convergence at around the Mean Squared Error of 0.02 . This likely indicates a set of sub-optimal local minima in the gradient equation for the encoding of the dataset, and a successful training of the network on the dataset must avoid this gradient trap. Trained encodings that end up at this or the higher convergence value are not useful, so any occurrence was rerun with different randomized initial weights.

3.2. Latent Space Dimension Comparison

We trained ten successful encodings for every even valued latent dimension from 2 to 30 for both the basic autoencoder and the sparse autoencoder. The MSEs of the dataset on the final learned encoding spaces and the best performing encoding space through the training process are shown in Figure 4. The same data is provided for the sparse encodings in Figure 5. The trend in MSE values as the latent encodings increase in dimension is similar looking to the reciprocal function for all four of the plots. By inspection, 16 appears to be a dimension choice that balances between high accuracy and low dimensionality.
For the sixteen dimension (16D) case, we give the reconstruction error for all combinations of measurement distance and angle in Table 1. Every measurement from the TREX13 data was put through the ten successfully trained 16D autoencoders and had a reconstruction error calculated. In comparison, the reconstruction errors for the 2D case are noticeably worse, as shown in Table 2.

3.3. 2D Encoding Space

16D encoding spaces are very hard to visualize, so for the sake of demonstration, 2D encodings will be evaluated. Figure 6 contains the full TREX13 dataset as encoded by the best performing of the ten 2D encodings from the previous section. The colors of the points match according to the measurement distances they are associated with in the dataset. Clearly there is a densely packed region of the encoding space where most of the dataset lies, with a few outlying points that tend to be in the 30 m–40 m measurement range.
The dense group is shown with more clarity in Figure 7. Here it is clear that the data measured at different distances forms bands that follow apparent contours from the 5m range to the 25 m range while the 30 m to 40 m data is not well separated, which is likely because the network cannot learn features that distinguish data measured at that far of a distance (low of a resolution). It is curious that the encodings at the other five distances appear in the correct order, and this was consistent across all of the learned 2D encoding spaces, see Figure 8.
Due to the nature of the data, it is very easy to distinguish measurement distance based on the size of the zero padding. It makes sense that the lowest dimension encoding is able to at minimum distinguish most of the distances. Additionally, different distances need features to be learned at different scales, so different parts of the network are activated, thus they have different encodings.
The outlying encodings are consistent between the learned encoding spaces. Figure 9 shows that the four encodings farthest from the mean are the same in the best and worst performing of the encoding spaces. The significance that all four are of target type 1, which is a DEU trainer, will not be explored here.
We also evaluated the sparse 2D encoding spaces; the best performing space is shown in Figure 10. Here we see that the sparsity constraint restricted the values of the encoding vectors to between 0.0 and 0.1 . This forced the previous outlier points to stay closer to the mean value. Additionally, the contours of distance measurement groups are not as separated.

3.4. Performance on Unseen Data

With a choice of 16 dimensions in the encoding space, we ran 5-fold cross-validation on the TREX13 dataset. The results of this are shown in Figure 11 for the autoencoder and sparse autoencoder.
Across both autoencoders, there were five successfully trained encodings for each of the five folds. The mean squared reconstruction error for each individual magnitude Fourier transform in the training sets and test sets was obtained for each of the 25 total encoding spaces.
Interestingly, the set of these error values fits a logarithmic scale despite the networks being trained on non-decibel data. In Figure 11, the base-10 logarithms of the reconstruction errors are visualized with violin plots, with the y axes corrected to show the values on a logarithmic scale. Violin plots are similar to box plots with the benefit of fitting multi-modal data far better. Vertically they show an estimated density, in this case the density estimation kernel is 10 % of the dataset. The violin plots were generated with an external Matlab library [46]. In these violins the data presented by a box plot is also shown; the white circles represent the medians and the gray vertical lines show the inner quartile range. The dark colored horizontal lines indicate the mean values.
The experiments with the autoencoders and sparse autoencoders used the same training and testing folds, which explains why the shapes of the violins associated with the folds look similar between the two networks. Looking at just the means and medians of the violins, in some of the folds the test set outperforms the training set. This shows that the learned encoding spaces generalize well to unseen data in the same dataset.
The logarithm of the reconstruction performance of the PondEx dataset is also shown in this figure with yellow violins. The 25 autoencoders and sparse autoencoders encoded the entire previously untouched PondEx dataset, producing similar looking violins regardless of fold. Notably, the PondEx data appears to achieve better reconstruction than the data that the autoencoders were trained on. The PondEx data was obtained in a controlled pool environment, whereas TREX13 was recorded in the ocean.
The PondEx data only contains data collected at either 5 and 10 m, so the distances that produced the largest error values in Table 1 are not represented. This suggests a skew due to selection bias. In Table 3 and Table 4, we can see that while the PondEx data performs fairly well at 5 m it does significantly worse than the training data at the 10 m range.

4. Conclusions

The goal of ongoing work is a disentanglement of the littoral channel noise from the true sonar response. We have demonstrated significant progress towards encoding underwater backscattering data. The precursory results show promise by demonstrating clustering correlated to measurement distance and the ability to generalize to unseen data. The optimal dimensionality of the encoding space has been estimated to be around sixteen dimensions, Section 3.2, and the efficacy of this latent vector rank was validated through cross validation in Section 3.4. While the basic convolutional autoencoders show promise, the sparse autoencoders do not provide tangible benefits yet in training efficiency or accuracy, but a constrained encoding space may yet prove useful in the future.
Future steps may involve developing alternatives to convolutional neurons or adding additional preprocessing steps to the data before training on it. Complex neural networks are one path that can be explored; this would include the phase information from the frequency domain data that is currently excluded at the cost of increasing the amount of operations needed per layer during both the forward and back propagation steps of the training.
The Keras default values were used in many places. In addition to fundamental changes like neuron type, variations on the metavariables may improve reconstruction accuracy and reduce training time. Candidates for this include: adjusting the network to have double the kernel count in the encoders so that the network is more symmetrical, using the He method [41] for weight initialization, changing which activation functions to use–PReLU may add unneeded complexity over a ReLU, varying kernel shape and size as relevant features may not be square, and fluctuating the number of convolutional layers in both the encoder and decoder.

Author Contributions

Conceptualization, A.S.G.; Methodology, T.J.L.; Software, T.J.L.; Validation, T.J.L.; Formal analysis, T.J.L.; Investigation, T.J.L.; Resources, A.S.G.; Writing—original draft, T.J.L. and M.B.; Writing—review & editing, A.S.G. and M.B.; Visualization, T.J.L.; Supervision, A.S.G.; Project administration, A.S.G.; Funding acquisition, A.S.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by The Office of Naval Research grant number N00174-20-1-0016.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
PCAPrincipal Component Analysis
ReLURectified Linear Unit
PReLUParametric Rectified Linear Unit
MSEMean Squared Error
ATRAutomated Target Recognition
GANGenerative Adversarial Network
VAEVariational Autoencoder
UXOUnexploded Ordinance

References

  1. Robinson, M.; Fennell, S.; DiZio, B.; Dumiak, J. Geometry and topology of the space of sonar target echos. J. Acoust. Soc. Am. 2018, 143, 1630–1645. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Fei, T.; Kraus, D.; Berkel, P. A new idea on feature selection and its application to the underwater object recognition. Proc. Meet. Acoust. 2012, 17, 070071. [Google Scholar] [CrossRef] [Green Version]
  3. Langner, F.A.; Jans, W.; Knauer, C.; Middelmann, W. Advantages and disadvantages of training based object detection/classification in SAS images. Proc. Meet. Acoust. 2012, 17, 070075. [Google Scholar] [CrossRef] [Green Version]
  4. Sawas, J.; Petillot, Y. Cascade of boosted classifiers for automatic target recognition in synthetic aperture sonar imagery. Proc. Meet. Acoust. 2012, 17, 070074. [Google Scholar] [CrossRef] [Green Version]
  5. Steiniger, Y.; Groen, J.; Stoppe, J.; Kraus, D.; Meisen, T. A study on modern deep learning detection algorithms for automatic target recognition in sidescan sonar images. Proc. Meet. Acoust. 2021, 44, 070010. [Google Scholar] [CrossRef]
  6. Warakagoda, N.; Midtgaard, Ø. Retrieval of similar targets in synthetic aperture sonar images with deep learning. Proc. Meet. Acoust. 2020, 40, 070013. [Google Scholar] [CrossRef]
  7. Bianco, M.J.; Gerstoft, P.; Traer, J.; Ozanich, E.; Roch, M.A.; Gannot, S.; Deledalle, C.A. Machine learning in acoustics: Theory and applications. J. Acoust. Soc. Am. 2019, 146, 3590–3628. [Google Scholar] [CrossRef] [Green Version]
  8. Williams, D.P.; España, A.; Kargl, S.G.; Williams, K.L. A family of algorithms for the automatic detection, isolation, and fusion of object responses in sonar data. Proc. Meet. Acoust. 2021, 44, 070022. [Google Scholar] [CrossRef]
  9. El Bergui, A.; Quidu, I.; Zerr, B.; Solaiman, B. Model based classification of mine-like objects in sidescan sonar using the highlight information. Proc. Meet. Acoust. 2012, 17, 070072. [Google Scholar] [CrossRef]
  10. Sammelmann, G.S. Simulation, Beam-forming, and Visualization of Bistatic Synthetic Aperture Sonar; Government Report OMB 0704-0188; DTIC: Fort Belvoir, VA, USA, 2010. [Google Scholar]
  11. Bays, M.J.; Shende, A.; Stilwell, D.J.; Redfield, S.A. A solution to the multiple aspect coverage problem. In Proceedings of the 2011 IEEE International Conference on Robotics and Automation, Shanghai, China, 9–13 May 2011; pp. 1531–1537. [Google Scholar] [CrossRef]
  12. Kriminger, E.; Cobb, J.T.; Príncipe, J.C. Online active learning for automatic target recognition. IEEE J. Ocean. Eng. 2014, 40, 583–591. [Google Scholar] [CrossRef]
  13. Du, X.; Seethepalli, A.; Sun, H.; Zare, A.; Cobb, J.T. Environmentally-adaptive target recognition for SAS imagery. In Detection and Sensing of Mines, Explosive Objects, and Obscured Targets XXII; International Society for Optics and Photonics: Bellingham, WA, USA, 2017; Volume 10182, p. 101820I. [Google Scholar]
  14. Stack, J. Automation for underwater mine recognition: Current trends and future strategy. In Detection and Sensing of Mines, Explosive Objects, and Obscured Targets XVI; International Society for Optics and Photonics: Bellingham, WA, USA, 2011; Volume 8017, p. 80170K. [Google Scholar]
  15. Dobeck, G.J. Algorithm fusion for automated sea mine detection and classification. In Proceedings of the MTS/IEEE Oceans 2001. An Ocean Odyssey. Conference Proceedings (IEEE Cat. No. 01CH37295), Honolulu, HI, USA, 5–8 November 2001; Volume 1, pp. 130–134. [Google Scholar]
  16. Kargl, S. Acoustic Response of Underwater Munitions near a Sediment Interface: Measurement Model Comparisons and Classification Schemes; Technical Report; Washington University Seattle Applied Physics Lab: Washington, DC, USA, 2015. [Google Scholar]
  17. Kargl, S.G.; Williams, K.L.; Thorsos, E.I.; Lopes, J.L. Synthetic aperture sonar simulations of cylindrical targets. J. Acoust. Soc. Am. 2009, 125, 2733. [Google Scholar] [CrossRef]
  18. Misiuk, B.; Brown, C.J.; Robert, K.; Lacharité, M. Harmonizing multi-source sonar backscatter datasets for seabed mapping using bulk shift approaches. Remote Sens. 2020, 12, 601. [Google Scholar] [CrossRef] [Green Version]
  19. Hefner, B.T.; Jackson, D.R.; Ivakin, A.N.; Wendelboe, G. Physics-based inversion of multibeam sonar data for seafloor characterization. J. Acoust. Soc. Am. 2013, 134, 4240. [Google Scholar] [CrossRef]
  20. Mukherjee, K.; Gupta, S.; Ray, A.; Phoha, S. Symbolic analysis of sonar data for underwater target detection. IEEE J. Ocean. Eng. 2011, 36, 219–230. [Google Scholar] [CrossRef] [Green Version]
  21. Ballard, D.H. Modular Learning in Neural Networks. In Proceedings of the AAAI Conference on Artificial Intelligence, Seattle, WA, USA, 13–17 July 1987. [Google Scholar]
  22. Schmidhuber, J. Deep learning in neural networks: An overview. Neural Netw. 2015, 61, 85–117. [Google Scholar] [CrossRef] [Green Version]
  23. Goodfellow, I.; Bengio, Y.; Courville, A. Deep Learning; Adaptive Computation and Machine Learning; MIT Press: Cambridge, MA, USA, 2017; Chapter 14; pp. 493–516. [Google Scholar]
  24. Chen, X.; Duan, Y.; Houthooft, R.; Schulman, J.; Sutskever, I.; Abbeel, P. InfoGAN: Interpretable Representation Learning by Information Maximizing Generative Adversarial Nets. In Advances in Neural Information Processing Systems 29; Lee, D., Sugiyama, M., Luxburg, U., Guyon, I., Garnett, R., Eds.; Neural Information Processing Systems Foundation, Inc.: Red Hook, NY, USA, 2016; pp. 2180–2188. [Google Scholar]
  25. Radford, A.; Metz, L.; Chintala, S. Unsupervised Representation Learning with Deep Convolutional Generative Adversarial Networks. arXiv 2016, arXiv:1511.06434. [Google Scholar]
  26. Salimans, T.; Goodfellow, I.; Zaremba, W.; Cheung, V.; Radford, A.; Chen, X. Improved Techniques for Training GANs. In Proceedings of the NIPS 2016, Conference on Neural Information Processing Systems, Barcelona, Spain, 5–10 December 2016; Curran Associates Inc.: Red Hook, NY, USA, 2016; pp. 2234–2242. [Google Scholar]
  27. Chen, J.L.; Summers, J.E. Deep neural networks for learning classification features and generative models from synthetic aperture sonar big data. Proc. Meet. Acoust. 2016, 29, 032001. [Google Scholar] [CrossRef] [Green Version]
  28. Karl Pearson, F.R.S. LIII. On lines and planes of closest fit to systems of points in space. Philos. Mag. J. Sci. 1901, 2, 559–572. [Google Scholar] [CrossRef] [Green Version]
  29. Bishop, C.M. Pattern Recognition and Machine Learning; Information Science and Statistics; Springer Science+Business Media: New York, NY, USA, 2006; Chapter 12; pp. 559–599. [Google Scholar]
  30. Plaut, E. From Principal Subspaces to Principal Components with Linear Autoencoders. arXiv 2018, arXiv:1804.10253. [Google Scholar]
  31. Goodfellow, I.; Bengio, Y.; Courville, A. Deep Learning; Adaptive Computation and Machine Learning; MIT Press: Cambridge, MA, USA, 2017; Chapter 5; pp. 115–117. [Google Scholar]
  32. Ioffe, S.; Szegedy, C. Batch Normalization: Accelerating Deep Network Training by Reducing Internal Covariate Shift. In Proceedings of the 32nd International Conference on Machine Learning, Lille, France, 6–11 July 2015; Volume 37. [Google Scholar]
  33. Dumoulin, V.; Visin, F. A guide to convolution arithmetic for deep learning. arXiv 2016, arXiv:1603.07285. [Google Scholar]
  34. Ng, A. Sparse Autoencoder. Stanford CS294A Lecture Notes. Online. 2011. Available online: https://web.stanford.edu/class/cs294a/sparseAutoencoder_2011new.pdf (accessed on 1 November 2022).
  35. Doersch, C. Tutorial on Variational Autoencoders. arXiv 2016, arXiv:1606.05908. [Google Scholar]
  36. Williams, K.L.; Kargl, S.G.; Thorsos, E.I.; Burnett, D.S.; Lopes, J.L.; Zampolli, M.; Marston, P.L. Acoustic scattering from a solid aluminum cylinder in contact with a sand sediment: Measurements, modeling, and interpretation. J. Acoust. Soc. Am. 2010, 127, 3356–3371. [Google Scholar] [CrossRef] [PubMed]
  37. Kargl, S.G.; Williams, K.L. Full Scale Measurement and Modeling of the Acoustic Response of Proud and Buried Munitions at Frequencies from 1–30 kHz; Final Report; University of Washington, SERDP: Washington, DC, USA, 2012. [Google Scholar]
  38. Abadi, M.; Agarwal, A.; Barham, P.; Brevdo, E.; Chen, Z.; Citro, C.; Corrado, G.S.; Davis, A.; Dean, J.; Devin, M.; et al. TensorFlow: Large-Scale Machine Learning on Heterogeneous Systems. 2015. Available online: https://www.tensorflow.org/ (accessed on 1 November 2022).
  39. Chollet, F. Keras. 2015. Available online: https://keras.io (accessed on 1 November 2022).
  40. Goodfellow, I.; Bengio, Y.; Courville, A. Deep Learning; Adaptive Computation and Machine Learning; MIT Press: Cambridge, MA, USA, 2017; Chapter 6; pp. 197–217. [Google Scholar]
  41. He, K.; Zhang, X.; Ren, S.; Sun, J. Delving Deep into Rectifiers: Surpassing Human-Level Performance on ImageNet Classification. In Proceedings of the ICCV 2015, IEEE International Conference on Computer Vision, Santiago, Chile, 7–13 December 2015. [Google Scholar]
  42. Andrew, L.; Maas, A.Y.H.; Ng, A.Y. Rectifier Nonlinearities Improve Neural Network Acoustic Models. In Proceedings of the ICML Workshop on Deep Learning for Audio, Speech, and Language Processing; International Conference on Machine Learning, Atlanta, GA, USA, 16 June 2013. [Google Scholar]
  43. Glorot, X.; Bengio, Y. Understanding the difficulty of training deep feedforward neural networks. In Proceedings of the Thirteenth International Conference on Artificial Intelligence and Statistics, Sardinia, Italy, 13–15 May 2010; Teh, Y.W., Titterington, M., Eds.; Volume 9. [Google Scholar]
  44. Kingma, D.P.; Ba, J.L. ADAM: A method for stochastic optimization. In Proceedings of the ICLR 2015, International Conference on Learning Representations, San Diego, CA, USA, 7–9 May 2015; Volume 3. [Google Scholar]
  45. Odena, A.; Dumoulin, V.; Olah, C. Deconvolution and Checkerboard Artifacts. Distill 2016, 1, e3. [Google Scholar] [CrossRef]
  46. Bechtold, B. Violin Plots for Matlab. Release v0.1. Github Release. 2016. Available online: https://github.com/bastibe/Violinplot-Matlab/releases/tag/v0.1 (accessed on 1 November 2022).
Figure 1. A “smile” image of a 155 mm Howitzer with collar (Target 9) from the TREX13 dataset, shown in decibels. This measurement was taken at 20 m and with a 0 inclination.
Figure 1. A “smile” image of a 155 mm Howitzer with collar (Target 9) from the TREX13 dataset, shown in decibels. This measurement was taken at 20 m and with a 0 inclination.
Jmse 11 00021 g001
Figure 2. The structure of the convolutional autoencoder used in this paper. The encodings are output from the Identity (Id) block. After replacing the identity with a sigmoid activation, a sparsity constraint can be imposed on the network, resulting in a sparse convolutional autoencoder.
Figure 2. The structure of the convolutional autoencoder used in this paper. The encodings are output from the Identity (Id) block. After replacing the identity with a sigmoid activation, a sparsity constraint can be imposed on the network, resulting in a sparse convolutional autoencoder.
Jmse 11 00021 g002
Figure 3. Plots of the loss trend lines for every model trained for some learning rates. The line colors serve only to differentiate different runs. The data for the other learning rates looks similar to the 0.01 and 0.00001 cases so it is excluded.
Figure 3. Plots of the loss trend lines for every model trained for some learning rates. The line colors serve only to differentiate different runs. The data for the other learning rates looks similar to the 0.01 and 0.00001 cases so it is excluded.
Jmse 11 00021 g003
Figure 4. The Final and Minimum loss values across 10 successfully trained encodings for each tested latent space dimension. The colors differentiate adjacent columns.
Figure 4. The Final and Minimum loss values across 10 successfully trained encodings for each tested latent space dimension. The colors differentiate adjacent columns.
Jmse 11 00021 g004
Figure 5. The Final and Minimum loss values across 10 successfully trained sparse encodings for each tested latent space dimension. The colors differentiate adjacent columns.
Figure 5. The Final and Minimum loss values across 10 successfully trained sparse encodings for each tested latent space dimension. The colors differentiate adjacent columns.
Jmse 11 00021 g005
Figure 6. The best performing 2D encoding space, showing the encodings of the input data. These are color coded by the measurement distance of the images.
Figure 6. The best performing 2D encoding space, showing the encodings of the input data. These are color coded by the measurement distance of the images.
Jmse 11 00021 g006
Figure 7. A close up of the tight group of encodings in Figure 6.
Figure 7. A close up of the tight group of encodings in Figure 6.
Jmse 11 00021 g007
Figure 8. The dense part of the worst performing 2D encoding space.
Figure 8. The dense part of the worst performing 2D encoding space.
Jmse 11 00021 g008
Figure 9. A comparison of outlier encodings between the best (a) and worst (b) performing encoding spaces. The labels indicate target number, measurement angle, and measurement distance.
Figure 9. A comparison of outlier encodings between the best (a) and worst (b) performing encoding spaces. The labels indicate target number, measurement angle, and measurement distance.
Jmse 11 00021 g009aJmse 11 00021 g009b
Figure 10. The best performing sparse 2D encoding space. The encoding points of the same four outlier inputs labeled in previous figures are also labeled here for comparison.
Figure 10. The best performing sparse 2D encoding space. The encoding points of the same four outlier inputs labeled in previous figures are also labeled here for comparison.
Jmse 11 00021 g010
Figure 11. 5-fold cross validation across the TREX13 dataset, also performance on PondEx data.
Figure 11. 5-fold cross validation across the TREX13 dataset, also performance on PondEx data.
Jmse 11 00021 g011
Table 1. The autoencoder’s average normalized reconstruction errors with 16 latent dimensions. All values × 10 3 .
Table 1. The autoencoder’s average normalized reconstruction errors with 16 latent dimensions. All values × 10 3 .
5 m10 m15 m20 m25 m30 m35 m40 m
80 0.133 0.686 1.716 4.021 4.883 9.283 22.009 22.009
60 0.149 0.698 1.678 3.131 3.676 4.449 5.891 5.891
40 0.129 0.515 1.928 2.502 3.800 4.153 4.998 4.998
20 0.200 0.775 3.169 3.520 6.335 7.475 10.532 10.532
0 0.278 1.157 3.441 4.976 8.483 9.167 14.659 14.659
20 0.278 1.083 2.414 4.095 5.832 7.842 8.907 8.907
40 0.159 0.674 1.871 2.724 4.372 4.622 5.072 5.072
60 0.172 0.617 1.278 2.820 4.183 4.084 5.949 5.949
80 0.202 0.592 1.497 4.268 4.900 6.701 11.929 11.929
100 0.175 6.245 6.245
120 0.428 6.168 6.168
140 0.576 4.361 4.361
160 0.902 7.110 7.110
180 1.475 16.876 16.876
200 0.895 6.671 6.671
220 0.577 4.662 4.662
240 0.477 10.769 10.769
260 0.772 22.848 22.848
Table 2. The autoencoder’s average normalized reconstruction errors with 2 latent dimensions. All values × 10 3 .
Table 2. The autoencoder’s average normalized reconstruction errors with 2 latent dimensions. All values × 10 3 .
5 m10 m15 m20 m25 m30 m35 m40 m
80 0.190 1.283 2.948 8.062 10.700 20.616 29.097 29.097
60 0.204 0.864 1.956 4.725 5.783 7.227 8.019 8.019
40 0.174 0.653 2.280 3.859 5.277 7.539 6.473 6.473
20 0.264 1.106 6.529 6.471 15.699 16.945 30.716 30.716
0 0.463 1.984 6.934 9.838 15.538 24.862 34.056 34.056
20 0.399 1.833 3.815 7.150 12.524 18.244 23.496 23.496
40 0.214 0.825 2.468 4.635 7.978 8.916 7.018 7.018
60 0.257 0.787 1.537 3.891 5.698 5.559 12.711 12.711
80 0.312 1.210 2.070 8.857 13.441 17.535 25.996 25.996
100 0.269 10.247 10.247
120 0.551 7.455 7.455
140 0.667 5.502 5.502
160 1.532 9.789 9.789
180 2.448 54.332 54.332
200 1.668 8.022 8.022
220 0.754 5.612 5.612
240 0.609 16.081 16.081
260 0.956 29.829 29.829
Table 3. Average error values for the PondEx datasets across all cross-validation folds (basic convolutional autoencoder). All values 10 3 .
Table 3. Average error values for the PondEx datasets across all cross-validation folds (basic convolutional autoencoder). All values 10 3 .
80 60 40 20 0 20 40 60 80
5 m 0.202 0.172 0.254 0.412 0.307 0.275 0.129 0.164 0.189
10 m 1.537 1.034 1.205 3.662 2.588 3.383 1.376 1.459 2.305
Table 4. The low measurement distance data from Table 1. All values × 10 3 .
Table 4. The low measurement distance data from Table 1. All values × 10 3 .
80 60 40 20 0 20 40 60 80
5 m 0.133 0.149 0.129 0.200 0.278 0.278 0.159 0.172 0.202
10 m 0.686 0.698 0.515 0.775 1.157 1.083 0.674 0.617 0.592
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Linhardt, T.J.; Sen Gupta, A.; Bays, M. Convolutional Autoencoding of Small Targets in the Littoral Sonar Acoustic Backscattering Domain. J. Mar. Sci. Eng. 2023, 11, 21. https://doi.org/10.3390/jmse11010021

AMA Style

Linhardt TJ, Sen Gupta A, Bays M. Convolutional Autoencoding of Small Targets in the Littoral Sonar Acoustic Backscattering Domain. Journal of Marine Science and Engineering. 2023; 11(1):21. https://doi.org/10.3390/jmse11010021

Chicago/Turabian Style

Linhardt, Timothy J., Ananya Sen Gupta, and Matthew Bays. 2023. "Convolutional Autoencoding of Small Targets in the Littoral Sonar Acoustic Backscattering Domain" Journal of Marine Science and Engineering 11, no. 1: 21. https://doi.org/10.3390/jmse11010021

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop