Next Article in Journal
Integrating Cosmic Microwave Background Readings with Celestial Navigation to Enhance Deep Space Navigation
Previous Article in Journal
An Automatic Lie Detection Model Using EEG Signals Based on the Combination of Type 2 Fuzzy Sets and Deep Graph Convolutional Networks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Semi-Supervised Autoencoder for Chemical Gas Classification with FTIR Spectrum

1
School of Electrical Engineering, Korea Advanced Institute of Science and Technology, Daejeon 34141, Republic of Korea
2
Chem-Bio Technology Center, Advanced Defense Science and Technology Research Institute, Agency for Defense Development, Daejeon 34186, Republic of Korea
*
Author to whom correspondence should be addressed.
Sensors 2024, 24(11), 3601; https://doi.org/10.3390/s24113601
Submission received: 1 April 2024 / Revised: 13 May 2024 / Accepted: 31 May 2024 / Published: 3 June 2024
(This article belongs to the Section Chemical Sensors)

Abstract

:
Chemical warfare agents pose a serious threat due to their extreme toxicity, necessitating swift the identification of chemical gases and individual responses to the identified threats. Fourier transform infrared (FTIR) spectroscopy offers a method for remote material analysis, particularly in detecting colorless and odorless chemical agents. In this paper, we propose a deep neural network utilizing a semi-supervised autoencoder (SSAE) for the classification of chemical gases based on FTIR spectra. In contrast to traditional methods, the SSAE concurrently trains an autoencoder and a classifier attached to a latent vector of the autoencoder, enhancing feature extraction for classification. The SSAE was evaluated on laboratory-collected FTIR spectra, demonstrating a superior classification performance compared to existing methods. The efficacy of the SSAE lies in its ability to generate denser cluster distributions in latent vectors, thereby enhancing gas classification. This study established a consistent experimental environment for hyperparameter optimization, offering valuable insights into the influence of latent vectors on classification performance.

1. Introduction

Chemical warfare agents represent highly toxic and perilous substances capable of inducing severe effects on the human body even in minute quantities. Many toxic chemical agents remain gaseous, colorless, and odorless at room temperature, posing challenges in promptly addressing chemical gas attacks or leaks. Hence, the early identification of and response to chemical gases are imperative tasks [1,2,3]. Additionally, the accurate classification of the specific chemical gas present in the atmosphere is crucial, as the approach to handling these gases varies with their types. Recognizing the significance of swift responses to chemical threats, researchers have focused on leveraging Fourier transform infrared (FTIR) spectroscopy [4,5,6,7,8,9,10,11,12]. This technique allows for the remote and non-contact analysis of materials. Through FTIR spectroscopy, the unique absorption or emission patterns in specific bands of the long-wavelength infrared region enable the identification of material types based on their molecular structures. Nevertheless, the observed spectrum is susceptible to background signals, light scattering, and noise inherent to equipment, making the clear identification and differentiation of chemical agents challenging. To address these challenges, recent studies have explored the application of machine learning and deep learning for FTIR sensing.
Yu et al. [13] employed an SVM to classify six classes, including five types of gases and a class that did not contain the other gases. They applied a preprocessing process, including converting the radiance spectrum to brightness temperature and noise removal [8] and baseline removal methods [8], achieving an enhanced performance by employing an SVM trained on preprocessed data. In a subsequent study [14], they added a process to remove the influence of background signals identified using principal component analysis (PCA) to the preprocessing method of the previous study and used this preprocessed data to train an SVM. The SVM trained with this preprocessed data achieved a higher performance than the correlation coefficients after removing background signatures [15], adaptive subspace detectors [16], and the SVM trained with preprocessed data in previous studies [13]. Nam et al. [17] classified gases using an SVM, discerning the presence or absence of classified gases based on SVM scores. Kim et al. [18] utilized deep neural networks and convolutional neural networks, outperforming the SVM as a comparative model. The ability of FTIR spectroscopy to identify unique spectral patterns based on material molecular structures extends beyond chemical gases to various materials. Ongoing research explores performance enhancement using machine learning and deep learning [19,20,21,22,23].
Given that measured spectra contain overlapping peaks not only from the specific band of the target material but also from background and surrounding materials, methods like feature selection and extraction have the potential to enhance the discrimination accuracy [24]. While methods using PCA have been studied [21], the use of deep learning autoencoders, expressing data characteristics as latent vectors, has gained traction. Jo et al. [24] classified agricultural products using an autoencoder and SVM, demonstrating an improved performance. The autoencoder was trained using the spectrum dataset, and the resulting encoder was employed as a feature extractor to train the SVM, exhibiting a superior performance compared to alternative feature extraction techniques like PCA or local linear embedding. Fine et al. [25] employed an autoencoder to utilize features with removed redundant information and noise. The features generated by the trained encoder were then employed in MLP training to classify functional groups. Chen et al. [26] proposed a model combining the encoder of a trained autoencoder and a classifier to classify orchid genotypes. Although many studies have employed autoencoders for extracting spectral features [27,28,29], the frameworks utilizing autoencoders in this manner have limitations, as the autoencoder and the model trained in a supervised manner are typically trained separately. While the autoencoder is trained to ensure that the reconstruction value produced through the encoder and decoder matches the input, the encoder obtained in this process may capture the semantics of the input data effectively, but this feature may not be optimal for supervised learning. To address these limitations, researchers have turned to semi-supervised autoencoders, combining both the autoencoder and supervised learning models for simultaneous training [30,31]. The semi-supervised autoencoder represents a unified framework that integrates an autoencoder with a task-specific supervised learning model attached to the latent vector of the autoencoder. Within this integrated framework, the autoencoder is trained to express not only the features for reconstruction from the input data but also the features for the attached supervised learning model, ultimately enhancing the performance of the attached task-specific model [32,33,34,35,36,37].
In this study, we propose a semi-supervised autoencoder for classifying chemical gases using FTIR spectra. The proposed model comprises a structure where a neural network-based classifier is attached to the latent vector, which is the encoder output of the autoencoder, training the classifier and autoencoder simultaneously. The classification performance of the proposed model was evaluated using nine types of FTIR spectra collected in a laboratory. A comparative analysis with existing machine learning methods, deep neural networks, and methods using an autoencoder as a feature extractor was conducted. The experimental conditions, including hyperparameter optimization and statistical performance analysis, were consistent across all models, providing insights into the impact of the latent vectors on the classification performance. Section 2 details the experimental data, the proposed model, and the hyperparameter optimization process. Section 3 presents the experimental setup and classification performance. Finally, Section 4 provides the conclusion.

2. Materials and Methods

2.1. Data Source

We employed FTIR spectral data collected from a chemical gas brightness temperature spectrum measurement system established at the Agency for Defense Development of the Republic of Korea. The measurement system comprises a custom-built remote chemical detection device (MSCAD, miniaturized standoff chemical agent detector) for capturing FTIR spectra, a gas chamber for collecting injected gas, and a black body used as a background. The MSCAD equipment has a field of view of 18 mrad, positioned at a distance of 3 m from the blackbody. The lens effective aperture is 15 mm. The chamber measures 2 m in length, with both the MSCAD equipment and the blackbody positioned at a distance of 0.5 m from the chamber. The process of collecting gas spectra was as follows: first, we ensured that the gas chamber window was within the field of view of the MSCAD measurement equipment. We aligned the MSCAD–chamber–blackbody setup so that the MSCAD ray passed through the gas chamber window and reached the black body. Subsequently, we set the temperature of the air inside the gas chamber and the temperature of the black body and injected chemical gas into the chamber. After introducing the gas, we waited for 20 s to allow mixing with the internal air and stabilization and then measured the brightness temperature spectrum of the gas at 35 Hz for approximately 20 to 40 s using the MSCAD device. The aforementioned process was conducted for nine types of chemical gases (cyclosarin (GF), lewisite (L), methanol, sarin (GB), nerve agent (VX), nitrogen mustard (HN3), soman (GD), sulfur mustard (HD), and tabun (GA)) with data collected at various concentrations for each gas by controlling the amount of substances injected into the chamber. For all gases except VX, the gas temperature inside the chamber was maintained at 40 ° C , while the blackbody was set at 30 ° C , resulting in a temperature difference of 10 ° C between the blackbody background and the gas. In the case of VX, the gas temperature inside the chamber was adjusted to 50   ° C , while the blackbody temperature remained at 40 ° C , maintaining the same 10 ° C temperature difference between the background and the gas. For security reasons related to the ongoing development of the spectrum measurement system, the names of the gases were pseudonymized as A–I.
To transform the collected spectrum into a dataset for training a neural network, we conducted a data preprocessing procedure and split the dataset into two subsets for training and testing datasets. We considered the spectrum range from the 739 cm 1 to 1260 cm 1 band, with a resolution of 1.6 cm 1 for the MSCAD measurement device. Therefore, the dimension N i of the input measurement spectrum x R N i was 327. The spectrum x underwent a 16-frame moving average sequentially to remove noise. In the preprocessing process, min–max normalization and zero-mean normalization were applied sequentially to adjust the scale and bias of the dataset. Additionally, as the measured time varies for each experimental condition, 100 data points were randomly selected for each condition set in the experiment. The label y R N C is a one-hot vector indicating the gas class to which the input x belongs, and N C was 9. Depending on the concentration in the dataset, high-concentration data were considered a training dataset, and low-concentration data were considered a test dataset. The higher the concentration of chemical gases, the higher the value in the unique peak band of the gas, allowing for the clear identification of the gas type in such cases. However, due to the potential fatal effects of chemical gases even in small amounts, accurate classification is necessary even in low-concentration gas situations. Therefore, we considered low-concentration data as the test dataset to evaluate applicability in a real environment. Table 1 shows the injected amounts and the number of data points for each gas included in the training dataset and test dataset.

2.2. Proposed Method

An autoencoder typically consists of an encoder that transforms input data into a low-dimensional latent vector and a decoder that reconstructs the input from the latent vector [38,39]. The autoencoder was trained in an unsupervised manner without labels. During the training step, latent vectors are expected to extract important features for reconstructing the same input data. This characteristic makes the autoencoder a representative model for representation learning. In various spectroscopy fields, researchers have explored using latent vectors for supervised tasks [24,25,26,27], employing the trained encoder as a feature extractor. However, since the process of training an autoencoder and the subsequent training of a task-specific model using latent vectors are disjoint, even if latent vectors capture features for input data reconstruction, these representations may lack features for label prediction in the subsequent task. Studies on semi-supervised autoencoders have studied to improve in performance by integrating an autoencoder with a task-specific model on the latent vector and training two models simultaneously, addressing the limitations of distinct learning phases and showing promising results.
From this perspective, we propose a deep neural network using a semi-supervised autoencoder (SSAE) to classify chemical gases based on FTIR spectra. The proposed model is a combination of an autoencoder and a classifier for gas classification. It receives the preprocessed spectrum x, generates the reconstruction x ^ R N i , and predicts the label y ^ R N C indicating the type of gas. The model structure comprises an encoder, decoder, and classifier, as illustrated in Figure 1. The encoder receives the input x and generates a latent vector z R N 2 through dense layers with nodes N 1 and N 2 . The nodes that have passed through the N 2 dense layer and subsequent activation function are considered the latent vector. The decoder, receiving z, reconstructs x ^ through dense layers with nodes N 1 and N i . The classifier, receiving z, predicts the probability y ^ through dense layers with nodes N 3 and N C . Batch normalization and leaky ReLU are employed after all dense layers, except the N 2 dense layer and output layers of the decoder and classifier. After the N 2 dense layer, a sigmoid activation function is used to set the value of the latent vector between 0 and 1. The decoder output layer uses a tangent hyperbolic activation function due to the input value range between 1 and 1. Dropout is applied to prevent overfitting.
We trained the proposed model using the Adam optimizer, and the training objective was to minimize the following loss function:
L ( x , x ^ , y , y ^ ) = 1 N B i = 1 N B 1 α x i x ^ i 2 2 + α c = 1 N C log exp y ^ i , c j = 1 N C exp y ^ i , j y i , c ,
where N B is the number of data points included in the batch, and N C is the number of gas classes to be classified. As evident in the above equation, the loss function of the proposed model is a combination of the mean squared error, representing the reconstruction error of the autoencoder, and the cross-entropy loss, representing the classification error. The model was trained to minimize both errors simultaneously. The weight of each reconstruction error and classification error was determined with α R .

2.3. Hyperparameter Optimization

The performances of deep learning models are significantly influenced by the chosen hyperparameters [40,41]. We optimized the hyperparameters through Bayesian optimization. Bayesian optimization defines a prior probability distribution for hyperparameters and an objective function representing the performance of the algorithm. Based on the objective function value for the previous parameters, the next parameters to be tried were calculated to optimize the objective function. This process was repeated for a predefined number of explorations to find the optimal hyperparameters. The Bayesian optimization method utilizes the results of previously searched parameters, enabling high-performance hyperparameter optimization compared to grid search and random search methods. A detailed description of Bayesian optimization is provided by Wu et al. [42].
The hyperparameter optimization process was accomplished through the following steps. First, hyperparameters were randomly selected, and a model using these hyperparameters was trained through 3-fold cross-validation. At this point, the data ratio for each class in each fold was set to be the same as the ratio of data for each class in the train dataset. Since our goal was to achieve high-performance gas classification, the cross-validation loss was calculated as the average cross-entropy loss of validation folds. In examining the cross-validation loss indicated by the selected hyperparameters, new parameters to be searched next were calculated within a predefined search space. In repeating this process, the hyperparameters that showed the lowest cross-validation loss among the searched hyperparameters were selected as the optimal hyperparameters.
The proposed model has seven hyperparameters: the number of neurons in each layer N 1 , N 2 , N 3 , dropout ratio D 1 , D 2 , weighting factor α , and learning rate η . Autoencoders are generally structured such that the number of neurons decreases as the encoder passes through each layer, and the number of neurons increases as the decoder passes through each layer. Considering this, the search space of N 1 is from 64 to 256, and that of N 2 is 8 to 64. The range of N 3 is the same as that of N 2 . The values of D 1 and D 2 ranged from 0 to 0.3, α ranged from 0.1 to 0.9, and η ranged from 0.01 to 0.0001 on a logarithmic scale.

3. Results

3.1. Experiment Setting

The performance of the proposed model was evaluated on the test dataset after being trained on the entire train dataset using the optimal hyperparameters selected through a hyperparameter optimization process. The number of hyperparameter combinations explored was 100, and the batch size and epochs in the training step were set to 1024 and 100 epochs, respectively. Evaluation metrics commonly used in classification performance analysis, such as the f1-score, precision, and recall, were employed. Despite introducing a hyperparameter optimization process for a detailed performance analysis of the proposed model, variations in performance may occur due to the selected optimal hyperparameters, influenced by the layer initialization value, the indexes of the folds divided in the cross-validation, and the randomness during the learning process. To enhance the general performance analysis of the model, a random state variable was set to determine the randomness in the training and hyperparameter optimization processes for reproducibility. Experiments were conducted by changing this random state five times, and the values were calculated as the average values of the evaluation metric values from the five trials.
For a comparative analysis, four conventional machine learning models (GNB, KNN, RF, SVM) and two deep learning models for classification (MLP, MLP-AE) were considered. The RF, SVM, KNN, and GNB models are commonly used for solving classification tasks using spectral data and served as performance comparison models for the deep learning models. Except for GNB, which has no hyperparameters, RF, SVM, and KNN underwent the same hyperparameter optimization process as described above. The hyperparameter search area for each model was determined by considering previous studies [43,44]. MLP and MLP-AE are deep learning models with structures similar to that of the proposed model. MLP-AE shares the same structure as the SSAE, but it first trains an autoencoder consisting of an encoder and a decoder, and then uses the latent vector generated by the trained encoder to train the classifier. As in previous studies [24,25,26,27], MLP-AE is a deep learning model that utilizes an autoencoder for feature extractor. MLP is a deep learning model for classification consisting of dense layers that incorporates an encoder and a classifier without a decoder, using leaky ReLU instead of a sigmoid as an activation function after the N 2 dense layer. MLP-AE and MLP have a total of six hyperparameters that are the same as those of the SSAE, except for α , and the search spaces are set to be the same as that of the SSAE. For all comparison models, the average performance values were calculated by changing the random state in the same manner as for the SSAE.

3.2. Performance Analysis

Table 2 summarizes the experimental results of the seven models for all gas classes in the test dataset. The evaluation metric values presented in the table were calculated as a weighted sum, taking into account the number of data points included in each class. The results indicate that the proposed SSAE model achieved the highest performance, surpassing 0.9 in all metrics, followed by MLP and MLP-AE. Given that the test dataset comprised low-concentration spectra, achieving a performance higher than 0.9 suggests that the proposed model can deliver robust classification even in scenarios with small amounts of gas in real environments. While there have been numerous instances of using trained autoencoders as feature extractors in previous research, the experimental results here show that MLP-AE, utilizing the extracted features, exhibits the lowest performance among deep learning models. This suggests that models using a semi-supervised autoencoder, which encourages feature extraction for classification, or general deep learning classifiers consisting of dense layers can achieve superior performances compared to models using a separately trained autoencoder as a feature extractor. Among the machine learning models, KNN demonstrated the highest performance, but still exhibited a lower performance compared to the deep learning based models.
Table 3, Table 4 and Table 5 present the f1-score, precision, and recall scores of the models for each gas. The SSAE exhibited the highest f1-score for all gases except gases A and F. RF achieved the highest f1-score for gas A, with only a 0.005 difference from the SSAE, but demonstrated a lower performance than the SSAE for the other gases. Particularly noteworthy is the significantly superior performance of the SSAE compared to RF in gases B, C, and D, with the most substantial difference of 0.413 observed in gas C. Also, GNB exhibited the highest f1-score for gas F, surpassing the SSAE by 0.029, but GNB performed less effectively than the SSAE for other gases and even exhibited a score as low as 0.409 in gas C. The SSAE demonstrated high performance in most cases for precision and recall, except for in gases A, C, and F in precision and gas D in recall. However, even in cases where the SSAE did not achieve the highest performance, the difference is only marginal, up to 0.138 from the model that attained the highest performance in those instances. Consequently, the proposed SSAE consistently demonstrates the best performance, whether through analyzing the performance for all gases collectively or individually for each gas.

3.3. Embedding Analysis

Our model demonstrated a higher classification performance than MLP-AE and MLP. To gain a more detailed understanding of why the classification performance of the SSAE is superior, we conducted an analysis of the latent vectors, which represent the output of the encoder where the input spectrum data are embedded in low dimensions. We employed the t-SNE method to reduce latent vectors to two dimensions and visualize them, examining the distribution of the latent vectors. Additionally, we calculated a silhouette score to numerically evaluate whether the latent vectors effectively represent characteristics for classification. The silhouette score indicates how closely the data corresponding to each class are clustered together and how well the clusters of each class are separated. Within a value between 1 and 1, a score closer to 1 signifies a higher clustering performance. The visualization and silhouette score calculation using t-SNE were used to analyze both the raw spectrum and latent vectors reduced to low dimensions with MLP, MLP-AE, and the SSAE on the test dataset.
Table 6 summarizes the silhouette scores, with the scores for the deep learning models representing the average value over five iterations of changing the random state. The SSAE exhibited the highest value of 0.6083 among the raw spectra and other methods. This indicates that the latent vectors generated by the SSAE encoder form well-defined clusters among the same classes, affirming the highest classification performance in the test dataset. The raw spectrum exhibited the lowest value of 0.4053, while MLP-AE showed a higher value than the raw spectrum, confirming the efficiency of using the trained encoder as a feature extractor. However, since MLP-AE demonstrated a lower score than MLP and the SSAE, this suggests that the features expressed by the latent vector are weighted more toward expressing features for reconstruction rather than features for classification. This provides an explanation for the lower classification performance observed in MLP-AE compared to MLP and SSAE.
Figure 2 visually represents the raw spectrum and latent vectors of MLP, MLP-AE, and the SSAE. Opaque points denote data from the train dataset, while clear points represent data from the test dataset. In the case of raw spectrum appearing in Figure 2a, points of the same class are not densely clustered, and gases A and B appear intermixed, which is also confirmed for gases C and D. Notably, gas H exhibits significant separation between some high-concentration training dataset points and low-concentration test dataset points. Conversely, the latent vectors of the deep learning models mostly form high-density clusters for the same classes, and the training and test data points are closely located. However, for MLP-AE, it can be observed that the test data points for gases C and D are partly mixed, whereas MLP and the SSAE demonstrate relatively clear distinctions between those points. This discrepancy is attributed to the fact that the autoencoder of MLP-AE is trained without labels for gas classification, visually illustrating why its classification performance is lower than those of MLP and the SSAE. These analyses provide evidence supporting the claim that the proposed SSAE model outperforms other models, as revealed by the visualization of latent vectors and silhouette clustering scores.

4. Conclusions

In this paper, we introduced the SSAE model for classifying gases present in FTIR spectra. The SSAE utilizes a semi-supervised autoencoder framework structure, jointly training an autoencoder for unsupervised input reconstruction and a classifier for supervised gas classification. In this combined framework, the autoencoder is trained not only to capture the semantics of the input data but also to represent features crucial for classification. The proposed model was trained on high-concentration gas spectra obtained in the laboratory, and its classification performance was evaluated on low-concentration gas spectra. In a statistical experimental environment, the SSAE demonstrated a superior classification performance compared to those of classical machine learning models and deep learning based classification models. The factors that enabled the SSAE to achieve the highest classification performance were confirmed by analyzing the distribution of latent vectors embedding the input data. The SSAE exhibits the ability to generate denser cluster distributions in latent vectors compared to deep learning classifiers that utilize a learned autoencoder or are composed of dense layers. This capability contributes to its high performance in classification, a conclusion supported by silhouette scores indicating the clustering performance and visualizations of the latent vectors.

Author Contributions

Conceptualization, H.-D.J., S.K. and H.N.; methodology, H.-D.J.; software, H.-D.J.; validation, H.-D.J., S.K. and H.N.; formal analysis, H.-D.J.; investigation, H.-D.J., S.K. and H.N.; resources, H.-D.J., S.K. and H.N.; data curation, H.-D.J., S.K. and H.N.; writing—original draft preparation, H.-D.J. and S.K.; writing—review and editing, D.E.C.; visualization, H.-D.J.; supervision, D.E.C.; project administration, D.E.C.; funding acquisition, D.E.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Agency For Defense Development Grant funded by the Korean Government (UI220039ZD).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available upon request from the corresponding author. The data are not publicly available due to legal restrictions.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

Abbreviations

The following abbreviations are used in this manuscript:
GNBGaussian naive Bayes;
KNNK-nearest neighbor;
SVMSupport vector machine;
RFRandom forest;
MLPMulti-layer perceptron;
MLP-AEMulti-layer perceptron with trained autoencoder;
SSAESemi-supervised autoencoder.

References

  1. Wu, J.; Qu, G.; Yan, L.; Wang, R.; Guo, P.; Yang, Y.; Li, X. Efficient Removal of Representative Chemical Agents by Rapid and Sufficient Adsorption via Magnetic Graphene Oxide Composites. Appl. Sci. 2023, 13, 10731. [Google Scholar] [CrossRef]
  2. Kumar, V.; Kim, H.; Pandey, B.; James, T.D.; Yoon, J.; Anslyn, E.V. Recent advances in fluorescent and colorimetric chemosensors for the detection of chemical warfare agents: A legacy of the 21st century. Chem. Soc. Rev. 2023, 52, 663–704. [Google Scholar] [CrossRef] [PubMed]
  3. Jang, H.D.; Park, J.H.; Chang, D.E.; Seo, H.S.; Nam, H. Deep Learning Algorithm with Residual Blocks for Chemical Gas Concentration Estimation. J. Inst. Control. Robot. Syst. 2023, 29, 556–561. [Google Scholar] [CrossRef]
  4. Manolakis, D.; Golowich, S.; DiPietro, R.S. Long-Wave Infrared Hyperspectral Remote Sensing of Chemical Clouds: A focus on signal processing approaches. IEEE Signal Process. Mag. 2014, 31, 120–141. [Google Scholar] [CrossRef]
  5. Harsanyi, J.; Chang, C.I. Hyperspectral image classification and dimensionality reduction: An orthogonal subspace projection approach. IEEE Trans. Geosci. Remote Sens. 1994, 32, 779–785. [Google Scholar] [CrossRef]
  6. Segal-Rosenheimer, M.; Dubowski, Y.; Jahn, C.; Schäfer, K.; Gerl, G.; Linker, R. Feasibility study of detection of hazardous airborne pollutants using passive open-path FTIR. In Proceedings of the Advanced Environmental, Chemical, and Biological Sensing Technologies VII, Orlando, FL, USA, 5–9 April 2010; Vo-Dinh, T., Lieberman, R.A., Gauglitz, G., Eds.; International Society for Optics and Photonics, SPIE: Orlando, FL, USA, 2010; Volume 7673, p. 76730G. [Google Scholar] [CrossRef]
  7. Shi, Z.; Huang, M.; Qian, L.; Han, W.; Zhang, G.; Lu, X. Unmanned Helicopter Airborne Fourier Transform Infrared Spectrometer Remote Sensing System for Hazardous Vapors Detection. Appl. Sci. 2024, 14, 1367. [Google Scholar] [CrossRef]
  8. Harig, R.; Matz, G. Toxic cloud imaging by infrared spectrometry: A scanning FTIR system for identification and visualization. Field Anal. Chem. Technol. 2001, 5, 75–90. [Google Scholar] [CrossRef]
  9. Yu, H.G.; Kim, W.; Park, D.J.; Chang, D.E.; Nam, H. Design of a Cooperative Chemical Agent (CA) Detection Algorithm with the Hyperspectral Imaging System. In Proceedings of the 2021 21st International Conference on Control, Automation and Systems (ICCAS), Jeju, Republic of Korea, 12–15 October 2021; pp. 2112–2115. [Google Scholar]
  10. Li, D.; Cui, F.; Wang, A.; Li, Y.; Wu, J.; Qiao, Y. Adaptive Detection Algorithm for Hazardous Clouds Based on Infrared Remote Sensing Spectroscopy and the LASSO Method. IEEE Trans. Geosci. Remote Sens. 2020, 58, 8649–8664. [Google Scholar] [CrossRef]
  11. Yu, H.G.; Lee, J.H.; Park, D.J.; Chang, D.E.; Nam, H. Design of a hard expectation-maximization-based normalized matched filter (EM-NMF) for the detection of chemical warfare agents under background contamination. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2020, 13, 2940–2952. [Google Scholar] [CrossRef]
  12. Yu, H.G.; Lee, C.S.; Park, D.J.; Chang, D.E.; Nam, H. Design of cooperative matched filter for detection of chemical agents. Electron. Lett. 2021, 57, 216–218. [Google Scholar] [CrossRef]
  13. Yu, H.G.; Lee, J.H.; Kim, Y.C.; Park, D.J. Intelligent detection algorithm of hazardous gases for FTIR-based hyperspectral imaging system using SVM classifier. In Proceedings of the Algorithms and Technologies for Multispectral, Hyperspectral, and Ultraspectral Imagery XXIII, Anaheim, CA, USA, 9–13 April 2017; SPIE: Anaheim, CA, USA, 2017; Volume 10198, pp. 75–83. [Google Scholar]
  14. Yu, H.G.; Lee, J.H.; Park, D.J.; Nam, H.W.; Park, B.H. Efficient detection algorithm of chemical warfare agents for FTIR-based hyperspectral imagery using SVM classifier. In Proceedings of the Imaging Spectrometry XXII: Applications, Sensors, and Processing, San Diego, CA, USA, 19–23 August 2018; Silny, J.F., Ientilucci, E.J., Eds.; International Society for Optics and Photonics, SPIE: San Diego, CA, USA, 2018; Volume 10768, p. 107680H. [Google Scholar] [CrossRef]
  15. Lee, J.H.; Yu, H.G.; Park, D.J.; Park, B.H.; Kim, J.H. Characterization of Hazardous Gases Using an Infrared Hyperspectral Imaging System. Instrum. Sci. Technol. 2015, 43, 469–484. [Google Scholar] [CrossRef]
  16. Manolakis, D.; Siracusa, C.; Shaw, G. Hyperspectral subpixel target detection using the linear mixing model. IEEE Trans. Geosci. Remote Sens. 2001, 39, 1392–1409. [Google Scholar] [CrossRef]
  17. Nam, H.; Kim, J.S.; Kim, H.J.; Lee, J.H.; Kang, Y.I.; Park, B.H. Development of a radiative transfer model for the determination of toxic gases by Fourier transform–infrared spectroscopy with a support vector machine algorithm. Instrum. Sci. Technol. 2019, 47, 264–277. [Google Scholar] [CrossRef]
  18. Kim, Y.C.; Yu, H.G.; Lee, J.H.; Park, D.J.; Nam, H.W. Hazardous gas detection for FTIR-based hyperspectral imaging system using DNN and CNN. In Proceedings of the Electro-Optical and Infrared Systems: Technology and Applications XIV, Warsaw, Poland, 11–14 September 2017; Huckridge, D.A., Ebert, R., Bürsing, H., Eds.; International Society for Optics and Photonics, SPIE: Warsaw, Poland, 2017; Volume 10433, p. 1043317. [Google Scholar] [CrossRef]
  19. Hesham, A.; Zeyad, L.; ElZahraa, F.; ElGamal, A.; Mohammed, P.; Sakr, M.; Sabry, Y.M. Deep Learning Enabling Analysis of Exhaled Breath Using Fourier Transform Spectroscopy in the Mid-Infrared. In Proceedings of the 2021 Tenth International Conference on Intelligent Computing and Information Systems (ICICIS), Cairo, Egypt, 5–7 December 2021; pp. 124–129. [Google Scholar] [CrossRef]
  20. Liu, X.; Yan, Y.; Xu, S. Multi-Scale Convolutional Neural Networks for the Quantitative Analysis of Multi-Component Gases. In Proceedings of the 2023 2nd International Conference on Robotics, Artificial Intelligence and Intelligent Control (RAIIC), Mianyang, China, 11–13 August 2023; pp. 395–399. [Google Scholar] [CrossRef]
  21. Farooq, S.; Del-Valle, M.; dos Santos, M.O.; dos Santos, S.N.; Bernardes, E.S.; Zezell, D.M. Rapid identification of breast cancer subtypes using micro-FTIR and machine learning methods. Appl. Opt. 2023, 62, C80–C87. [Google Scholar] [CrossRef]
  22. Enders, A.A.; North, N.M.; Fensore, C.M.; Velez-Alvarez, J.; Allen, H.C. Functional Group Identification for FTIR Spectra Using Image-Based Machine Learning Models. Anal. Chem. 2021, 93, 9711–9718. [Google Scholar] [CrossRef] [PubMed]
  23. Das, S.; Paramane, A.; Kumari, S.; Chatterjee, S. Deep Learning Aided Classification of Ageing Condition of Natural Ester Oils Using FTIR Analysis. In Proceedings of the 2023 IEEE 3rd International Conference on Sustainable Energy and Future Electric Transportation (SEFET), Bhubaneswar, India, 9–12 August 2023; pp. 1–6. [Google Scholar] [CrossRef]
  24. Jo, S.; Sohng, W.; Lee, H.; Chung, H. Evaluation of an autoencoder as a feature extraction tool for near-infrared spectroscopic discriminant analysis. Food Chem. 2020, 331, 127332. [Google Scholar] [CrossRef]
  25. Fine, J.A.; Rajasekar, A.A.; Jethava, K.P.; Chopra, G. Spectral deep learning for prediction and prospective validation of functional groups. Chem. Sci. 2020, 11, 4618–4630. [Google Scholar] [CrossRef] [PubMed]
  26. Chen, Y.; Chen, Y.; Feng, X.; Yang, X.; Zhang, J.; Qiu, Z.; He, Y. Variety Identification of Orchids Using Fourier Transform Infrared Spectroscopy Combined with Stacked Sparse Auto-Encoder. Molecules 2019, 24, 2506. [Google Scholar] [CrossRef] [PubMed]
  27. Yang, S.B.; Moreira, J.; Li, Z. Predicting crude oil properties using fourier-transform infrared spectroscopy (FTIR) and data-driven methods. Digit. Chem. Eng. 2022, 3, 100031. [Google Scholar] [CrossRef]
  28. Tung, P.Y.; Sheikh, H.A.; Ball, M.; Nabiei, F.; Harrison, R.J. SIGMA: Spectral interpretation using gaussian mixtures and autoencoder. Geochem. Geophys. Geosyst. 2023, 24, e2022GC010530. [Google Scholar] [CrossRef]
  29. De Oliveira, J.P.G.; Bastos-Filho, C.J.A.; Oliveira, S.C. Chemical sample classification using autoencoder-based spectroscopy. In Proceedings of the 2022 SBFoton International Optics and Photonics Conference (SBFoton IOPC), Recife, Brazil, 13–15 October 2022; pp. 1–5. [Google Scholar] [CrossRef]
  30. Gogna, A.; Majumdar, A. Semi Supervised Autoencoder. In Proceedings of the Neural Information Processing, Kyoto, Japan, 16–21 October 2016; Hirose, A., Ozawa, S., Doya, K., Ikeda, K., Lee, M., Liu, D., Eds.; Springer: Cham, Switzerland, 2016; pp. 82–89. [Google Scholar]
  31. Haiyan, W.; Haomin, Y.; Xueming, L.; Haijun, R. Semi-supervised autoencoder: A joint approach of representation and classification. In Proceedings of the 2015 International Conference on Computational Intelligence and Communication Networks (CICN), Jabalpur, India, 12–14 December 2015; pp. 1424–1430. [Google Scholar]
  32. Gogna, A.; Majumdar, A.; Ward, R. Semi-supervised stacked label consistent autoencoder for reconstruction and analysis of biomedical signals. IEEE Trans. Biomed. Eng. 2016, 64, 2196–2205. [Google Scholar] [CrossRef] [PubMed]
  33. Chai, Z.; Song, W.; Wang, H.; Liu, F. A semi-supervised auto-encoder using label and sparse regularizations for classification. Appl. Soft Comput. 2019, 77, 205–217. [Google Scholar] [CrossRef]
  34. Yin, W.; Li, L.; Wu, F.X. A semi-supervised autoencoder for autism disease diagnosis. Neurocomputing 2022, 483, 140–147. [Google Scholar] [CrossRef]
  35. Wu, X.; Zhang, Y.; Cheng, C.; Peng, Z. A hybrid classification autoencoder for semi-supervised fault diagnosis in rotating machinery. Mech. Syst. Signal Process. 2021, 149, 107327. [Google Scholar] [CrossRef]
  36. Said, M.; Wahba, A.; Khalil, D. Semi-supervised deep learning framework for milk analysis using NIR spectrometers. Chemom. Intell. Lab. Syst. 2022, 228, 104619. [Google Scholar] [CrossRef]
  37. Plana Rius, F.; Philipsen, M.P.; Mirats Tur, J.M.; Moeslund, T.B.; Angulo Bahón, C.; Casas, M. Autoencoders for Semi-Supervised Water Level Modeling in Sewer Pipes with Sparse Labeled Data. Water 2022, 14, 333. [Google Scholar] [CrossRef]
  38. Hinton, G.E.; Salakhutdinov, R.R. Reducing the Dimensionality of Data with Neural Networks. Science 2006, 313, 504–507. [Google Scholar] [CrossRef] [PubMed]
  39. Caterini, A.L.; Chang, D.E. Deep Neural Networks in a Mathematical Framework; Springer: Cham, Swizerland, 2018. [Google Scholar]
  40. Yang, L.; Shami, A. On hyperparameter optimization of machine learning algorithms: Theory and practice. Neurocomputing 2020, 415, 295–316. [Google Scholar] [CrossRef]
  41. Jang, H.D.; Park, J.H.; Nam, H.; Chang, D.E. Deep neural networks for gas concentration estimation and the effect of hyperparameter optimization on the estimation performance. In Proceedings of the 2022 22nd International Conference on Control, Automation and Systems (ICCAS), Jeju, Republic of Korea, 27 November–1 December 2022; pp. 15–19. [Google Scholar] [CrossRef]
  42. Wu, J.; Chen, X.Y.; Zhang, H.; Xiong, L.D.; Lei, H.; Deng, S.H. Hyperparameter optimization for machine learning models based on Bayesian optimization. J. Electron. Sci. Technol. 2019, 17, 26–40. [Google Scholar]
  43. Rodriguez-Galiano, V.; Sanchez-Castillo, M.; Chica-Olmo, M.; Chica-Rivas, M. Machine learning predictive models for mineral prospectivity: An evaluation of neural networks, random forest, regression trees and support vector machines. Ore Geol. Rev. 2015, 71, 804–818. [Google Scholar] [CrossRef]
  44. Dash, R.K.; Nguyen, T.N.; Cengiz, K.; Sharma, A. Fine-tuned support vector regression model for stock predictions. Neural Comput. Appl. 2023, 35, 23295–23309. [Google Scholar] [CrossRef]
Figure 1. Network architecture for SSAE, which consists of encoder, decoder, and classifier.
Figure 1. Network architecture for SSAE, which consists of encoder, decoder, and classifier.
Sensors 24 03601 g001
Figure 2. Visualization of two-dimensional reduced (a) raw spectrum and latent vectors generated by (b) MLP, (c) MLP-AE, and (d) the SSAE on the test dataset through t-SNE.
Figure 2. Visualization of two-dimensional reduced (a) raw spectrum and latent vectors generated by (b) MLP, (c) MLP-AE, and (d) the SSAE on the test dataset through t-SNE.
Sensors 24 03601 g002
Table 1. Injected amounts and the number of samples for each gas in the training and test datasets.
Table 1. Injected amounts and the number of samples for each gas in the training and test datasets.
Chemical GasTrain DatasetTest Dataset
Amounts ( µ L) Samples Amounts ( µ L) Samples
A6, 8, 10, 154002, 4200
B6, 8, 103002, 4200
C6, 8, 12, 154002, 4200
D6, 8, 12, 15, 205002, 4200
E60, 90, 12030030100
F60, 80, 120, 16040020, 40200
G40, 6020020100
H12, 162004, 8200
I201002, 4, 6, 8400
total 2800 1800
Table 2. Performance evaluation scores of models on test dataset.
Table 2. Performance evaluation scores of models on test dataset.
Modelf1-ScorePrecisionRecall
GNB0.77050.87050.7844
KNN0.85320.88940.8584
SVM0.74040.81340.7718
RF0.78090.87740.7875
MLP0.89020.91870.8938
MLP-AE0.87090.90950.8758
SSAE0.91160.93130.9133
Table 3. F1-score of models for each gas on test dataset.
Table 3. F1-score of models for each gas on test dataset.
ModelABCDEFGHI
GNB0.82320.81300.47900.65460.74900.89080.52940.63481.0000
KNN0.81900.71040.77350.80411.00000.83230.99690.74121.0000
SVM0.80940.51330.36170.69480.98120.74000.79290.65701.0000
RF0.89180.59360.47390.61860.96580.73230.98690.74191.0000
MLP0.86950.80140.85370.88570.99900.84151.00000.76810.9960
MLP-AE0.82750.73430.77500.84730.99320.85890.99270.80261.0000
SSAE0.88680.84370.88880.90891.00000.86101.00000.81521.0000
Table 4. Precision of models for each gas on test dataset.
Table 4. Precision of models for each gas on test dataset.
ModelABCDEFGHI
GNB0.74491.00001.00000.48770.59880.80321.00001.00001.0000
KNN0.69501.00000.85530.74161.00000.71301.0001.00001.0000
SVM0.69270.80000.80000.54420.97860.59480.80001.00001.0000
RF0.93590.96340.97740.46450.97180.57800.98391.00001.0000
MLP0.77280.99880.98430.80280.99800.72651.00001.00000.9923
MLP-AE0.70621.00000.99120.74180.98690.75281.0001.00001.0000
SSAE0.79781.00000.97940.84811.00000.75641.00001.00001.0000
Table 5. Recall of models for each gas on test dataset.
Table 5. Recall of models for each gas on test dataset.
ModelABCDEFGHI
GNB0.92000.68500.31500.99501.00001.00000.36000.46501.0000
KNN0.99800.55390.70800.88001.00001.00000.99400.58901.0000
SVM0.97400.37800.23500.97000.98401.00000.78600.50491.0000
RF0.85200.43090.31300.92690.96001.00000.99000.59001.0000
MLP0.99900.67990.75390.98801.00001.00001.00000.62401.0000
MLP-AE1.00000.58200.64400.99301.00001.00000.98600.67101.0000
SSAE1.00000.73200.81700.98201.00001.00001.00000.68901.0000
Table 6. The silhouette scores for the raw spectrum and latent vectors generated by MLP, MLP-AE, and the SSAE on the test datatset.
Table 6. The silhouette scores for the raw spectrum and latent vectors generated by MLP, MLP-AE, and the SSAE on the test datatset.
Raw SpectrumMLPMLP-AESSAE
silhouette score0.40530.55900.46530.6083
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

Jang, H.-D.; Kwon, S.; Nam, H.; Chang, D.E. Semi-Supervised Autoencoder for Chemical Gas Classification with FTIR Spectrum. Sensors 2024, 24, 3601. https://doi.org/10.3390/s24113601

AMA Style

Jang H-D, Kwon S, Nam H, Chang DE. Semi-Supervised Autoencoder for Chemical Gas Classification with FTIR Spectrum. Sensors. 2024; 24(11):3601. https://doi.org/10.3390/s24113601

Chicago/Turabian Style

Jang, Hee-Deok, Seokjoon Kwon, Hyunwoo Nam, and Dong Eui Chang. 2024. "Semi-Supervised Autoencoder for Chemical Gas Classification with FTIR Spectrum" Sensors 24, no. 11: 3601. https://doi.org/10.3390/s24113601

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