Next Article in Journal
WB Score: A Novel Methodology for Visual Classifier Selection in Increasingly Noisy Datasets
Previous Article in Journal
Comparative Study of Musical Timbral Variations: Crescendo and Vibrato Using FFT-Acoustic Descriptor
Previous Article in Special Issue
Enhanced Performance Electrochemical Biosensor for Detection of Prostate Cancer Biomarker PCA3 Using Specific Aptamer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Steady vs. Dynamic Contributions of Different Doped Conducting Polymers in the Principal Components of an Electronic Nose’s Response

1
Department of Physics, University of Monastir Tunisia, Monastir 5000, Tunisia
2
Institute of Electronics, Microelectronics and Nanotechnology (IEMN, UMR 8520) | University Lille, CNRS, University Polytechnique Hauts-de-France , F-59000 Lille, France
*
Author to whom correspondence should be addressed.
Eng 2023, 4(4), 2483-2496; https://doi.org/10.3390/eng4040141
Submission received: 24 August 2023 / Revised: 15 September 2023 / Accepted: 20 September 2023 / Published: 22 September 2023

Abstract

:
Multivariate data analysis and machine learning classification have become popular tools to extract features without physical models for complex environments recognition. For electronic noses, time sampling over multiple sensing elements must be a fair compromise between a period sufficiently long to output a meaningful information pattern and sufficiently short to minimize training time for practical applications. Particularly when a reactivity’s kinetics differ from the thermodynamics in sensitive materials, finding the best compromise to get the most from the data is not obvious. Here, we investigate the influence of data acquisition to improve or alter data clustering for molecular recognition on a conducting polymer electronic nose. We found out that waiting for sensing elements to reach their steady state is not required for classification, and that reducing data acquisition down to the first dynamical information suffices to recognize molecular gases by principal component analysis with the same materials. Especially for online inference, this study shows that a good sensing array is not an array of good sensors, and that new figures of merit should be defined for sensing hardware using machine learning pattern recognition rather than metrology.

1. Introduction

Current advances in artificial intelligence and internet-of-things facilitate identifying new input layers as data generators to feed machine learning (ML) for information classification. Before this, electronic noses (eNoses) have nourished the fantasy for more than 60 years of emulating a biological sense to perform classification tasks that sensors can barely make with molecular patterns [1,2,3,4,5]. It is therefore very appropriate to wonder whether conventional approaches to selecting sensing hardware are adapted to information processing requirements, with pattern recognition having more standards nowadays.
Chemistry is no trivial physics: it is no 1D continuum nor spectrum, but volatile molecules compose a vast group of hundred billion countable objects [6,7]. Therefore, measuring all of them simultaneously with a restricted and well-chosen set of ultimately selective sensors is unrealistic. If such complexity suits better ML-supported recognition than quantification using metrological sensors, proposing clear sensing figures of merit (FoMs) that guide technologists to identify the right sets of materials is still needed in such context [8,9,10,11]. The main difficulty in identifying those lies in the fact that smart sensing arrays do not involve physical models at the opposite of metrological sensors: ML is purely mathematically driven, which does not help in identifying device physical features to generate qualitative data. Each output signal conditions individual components of input data, but the quality of the output data themselves relates to features that affect the performances of the classifier as a group of device properties. The relationship between the data transformation as a result and the physical properties of a chosen sensing hardware is not straightforward and largely depends on the association of a software classifier with a sensing hardware. For an eNose, conductometric sensing arrays are often used as a hardware [12,13,14,15], in conjunction with principal component analysis (PCA)/k-means clustering (k-means) as a classifier [5,16,17,18]. In such systems, time-dependent current signals are inputted by environmentally sensitive materials in an array. An information feature of a high-dimensional vector is composed of a single current signal for classification. PCA is a linear classifier as it involves only one matrix transformation from a dataset of input vectors to a set of scores. The PCA projects data on a subspace defined by a basis of the covariance matrix eigenvectors which have the highest eigenvalues. As a consequence, PCA aims to ignore the contributions of “silent” sensing elements, or the ones with activities that are uncorrelated to one another. Conventional sensors’ FoMs do not consider information correlation between sensing elements, as it relates strictly to physical properties of individual elements to be standalone: for instance, selectivity, sensitivity or time response [19]. The latter is an important FoM that usually determines if a technology is suitable for online application. If information speed is essential for telecommunication, it is not necessarily the case for the velocity of information carriers themselves. The implementation of recurrent networks for sensing elements can illustrate such fact, where the group dynamic of different slow sensing elements enables recognizing frequency-modulated signals [20]. Homogenizing the physical properties of the population of sensing elements can even degrade the classifier’s recognition performance, so when the dynamic of sensing devices carries information, a long-enough acquisition is crucial for classification [20]. However, time is also a resource to enrich a database to train a system: the slower the hardware, the longer the training, which ultimately conditions any classifier’s recognition performance. As a consequence, in case information lies in the dynamic of a hardware [21], a fair compromise must be found for data acquisition time to optimize training with enough meaningful data but reasonable training duration for environmental pattern recognition.
In the framework of recognizing volatile solvents with doped conducting polymers, we propose in this study to quantify how relevant it is to consider the acquisition time as a metrological FoM for effective molecular recognition with a conductometric nose using PCA/k-means. By using both static and dynamic descriptors, we show that a steady sensing element response is unnecessary for optimal classification, and that training duration can greatly be shortened for practical inference.

2. Experimental Section

Device Fabrication: The fabrication of sensing elements used in this study has been reported elsewhere [22]. Concentric spiral-shaped Au microelectrodes (channel length L = 400 nm, spiral electrodes length W such as W/L > 103 in a round cavity of 28 μm in diameter) were micro-fabricated by electron beam lithography in a cleanroom environment. The devices’ active area features a large W/L over a restricted area to maximize conductance sensitivity while minimizing total area of an array of them for system integration. Contact lines are passivated by a conformal CVD-processed Parylene C passivation layer (2 μm thickness), structured by an O2 dry etching to expose pads and the devices active area, so each device outputs an individual response for co-integration at small scale. The active materials were subsequently deposited on top of the sensing arrays by drop casting. Formulations of regio-regular P3HT and dopants solutions (composed of Fe(OTf)3, Bi(OTf)3, Cu(OTf)2, In(OTf)3, Al(OTf)3, Dy(OTf)3 and Ce(OTf)3) were performed in two different steps to minimize the device electrical property distribution due to material heterogeneity if deposited with a single formulation.
Electrical Characterization: To obtain raw data, the study exploited current versus time measurements of eight populations of materials, composed of three different devices for each material. Each of the 24 current versus time trace sizes 6899 data points per current trace, sampled every second. The current measurements were performed with an Agilent 4155 parameter analyzer in an air blow, passing through a 5 mL vial that contains 1 mL of a volatile solvent [22]. The system was calibrated to blow at a rate of 1 mL/s over a sensing array. The control of the vapor exposure was manually operated with pneumatic valves (asynchronous delays up to five seconds can be considered between the actual operation of the valves and the attributions of the labels on the current traces). Elementary sequences of gas exposures were set to three minutes for sensing elements to reach their steady state. Each volatile solvent exposure was followed by a purging sequence when the air flow passes through an empty vial to enable the response of the sensing arrays recover their response.
Data Analysis: Raw data currents were compiled and treated without filtering. Raw data curves have already been published [22] and are available on a public repository. Principal component analysis (PCA) was used as a linear classifier by mean of the online open access tool Clustvis [23]. The PCA data were scaled by unit variance and computed by singular value decomposition. All PCA data are available as Supplementary Information of this article. Additionally, 95% confidence ellipsoids were determined by k-means clustering with k = 3, from random values for the initial centroids (in each case, k-means clustering was run at least three times to ensure repeatability of the resulting clusters).

3. Results

3.1. Data Collection and Feature Extraction

To interpret the physical significance of information features for the quality of a recognition, the machine learning (ML) approach is preferred to the deep learning (DL) one (see Figure 1 for the overall approach). DL algorithms are different from ML ones mostly by their complexity to allow the end-user not to have to identify information descriptors from raw data a priori. In our case, we chose specifically to compare two different features α1 and α2 that are extensively used in conductometric eNoses as information descriptors: the relative resistance modulation α1 = R/R0 − 1 and the drift resistance α2 = Ṙ/R, with R being the Ohmic resistance of a conductometric sensing element measured at a given time during exposure to an environment of a specific class, Ṙ its first derivative of time and R0 its value right before it starts to be exposed to an environment of a specific class at t0. As α1 compares two states before and after being exposed to a class, this information descriptor is qualified as a “thermodynamic feature” and relates to how much the signal’s transport is affected by the presence of a given environment compared to a reference, through a given material. Instead, α2 uses a first-order derivative of time to quantify how fast or slow the signal decays or amplifies through a given material by the presence of a given environment: this information descriptor is qualified as a “kinetic feature”. Each feature is calculated for different scenarios of crossed-coupled environments/devices parameters, with a population of 24 different sensitive devices (composed of eight populations of sensitive materials measured on 3 different devices), exposed at 18 occurrences to different environments (six alternate permutations of 3 different solvent vapor exposures). The experimental details are fully described in a previous study [22], where the PCA was used exclusively with the feature α1 calculated exclusively for the last ten seconds of each three-minute environment exposure. Although the steady-state response for these materials was relatively fast compared to that of other materials for comparable tests [24,25,26,27], the dataset totals over 44 h of acquisition time assisted by manual operations to expose the different materials on different devices with different environments repeatedly to train the system. Applying the same approach to exposing n! times the n different vapor classes for three minutes would result in over a week, over a month and over seven months of total data acquisition, without interruption, to respectively recognize four, five and six different classes only. As reducing the number of materials, devices or exposures would have consequences of degrading the recognition, it is straightforward to investigate decreasing the elementary period for one single environment exposure to minimize the total acquisition duration. To assess the classifier performance for a reduced acquisition time, the same dataset is considered, where data are pruned to ten seconds per acquisition sequences after the exposure starts in order to make sure that the recognition performances do not depend on the acquisition itself, keeping the same number of datapoints to analyze in each case. Also, PCA/k-means clustering was systematically used on all the different case scenarios with the same normalization and initialization settings to not bias the clustering by user-dependent factors. The recognition after clustering uses the same rules and thresholds in each case to assess a data sample as successfully, uncertainly or unsuccessfully recognized. All the individual PCA analyses are provided as Supplementary Information.

3.2. Steady Resistance Modulation as Thermodynamic Information Feature

The first analysis was performed using α1 as an information descriptor (see Figure 2), as in the former studies using the same dataset [22,28]. The choice of this specific feature was informed by the hypothesis that the presence of volatile molecules modulates the polymer electrical conductance by interacting with its dopant and shifts its thermodynamic charge-transfer equilibrium [22]. As a thermodynamic property, the acquisition was set to three minutes per exposure to provide enough time for all devices to reach a quasi-steady state of conductance (steady output currents at a constant applied voltage). Here, from the whole dataset, 18 different sub-datasets were generated by segmenting the data to a concatenation of sequences of ten seconds after each exposure starts at t0 (10 datapoints per exposures, 180 datapoints per sub-datasets—see Figure 2a). For each exposure, the feature α1 takes into account two values: the resistance measured between the considered time interval and a reference resistance value R0 measured while unexposed just before gas exposure. In each case, PCA was performed for the different sub-datasets, with the individual variance for each PC of each PCA for the different acquisition times shown in Figure 2b. From these values, one can observe that most of the variance (>66%) is explained within the first two PCs, except for data generated during the first ten seconds after t0. It is also noticed that the variance for all 24 PCs does not change significantly for acquisitions occurring 30 s after exposure (Figure 2b, the color shade encodes the variance values in linear scale), suggesting that the feature α1 is not enriched for the data separation in (PC1,PC2) if exposures last longer than 30 s. As shown by their scores in the (PC1,PC2) projection (Figure 2c), the PCA on α1 during the first ten seconds of acquisition, denoted PCAα10–9, embeds most variability from acetone exposures on PC1 compared to two other clusters of data for water and ethanol, which are almost entirely contained in the acetone confidence ellipsoid. Already for PCAα110–19 (Figure 2d), acetone is completely separable from the other two gases, and a minor overlap remains between water and ethanol confidence ellipsoids. The final sequence PCAα1170–179 in Figure 2e shows an ideal clustering in the (PC1,PC2) projection, which was used in the previous study for gas recognition [22]. The loadings of PC1 and PC2 do not significantly depend on the acquisition sequences, except for the first ten seconds (Figure 2f,g). Excluding the first ten seconds, deviations overtime are however noticed: for instance, the contribution of Ce(OTf)3-doped devices increases monotonically in PC1α1, such as the contribution of pristine P3HT in PC2α1. This suggests that materials that better facilitate the (PC1,PC2) data separation but are expressed slower than others do not necessarily enhance the clustering. The previous study concluded on the exclusive basis of the PCAα1170–179 analysis; two and three materials were preferred for recognizing ethanol from acetone from water using α1 as and information descriptor [22]. Figure 2f,g show that such conclusion is highly acquisition-dependent (even for the same dataset considered) and that experiments performed with shorter time exposures might have concluded otherwise for the material selection. As such dependency on the sample acquisition time greatly conditions the preferred material choice for classification, it appears legitimate to investigate a dynamical feature such as α2 as a better information descriptor for gas recognition with the same raw current data.

3.3. Resistance Drift as Kinetic Information Feature

Drift resistance α2 in conductometric eNoses is less often used as an information descriptor, although it is worth noticing that most conventional techniques in analytical chemistry use dynamical information features as information descriptors, such as chromatography (retention time of molecules on a substrate) or surface-plasmon resonance (absorption/desorption kinetic constants of molecules on a substrate). In a material, drifts can often be associated to a memory effect that reversibly or irreversibly alters a structural property [29,30,31,32,33], rather than the transduction of environmental information in a device. In case such drift is irreversible, such behavior appears as a material instability due to ageing and is often associated as a drawback for the lifetime. However, since such reactivity on the material is environmentally induced per se, it is legitimate to wonder whether the information quality is preserved despite its effect on the information carrier itself. As an example, Kiselev et al. showed that such resistance drift (two-time median-resistance increase over four years) on a conductometric nose is totally decorrelated from the ability of the classifier to recognize a class, as the strong correlation between all sensing elements’ resistance drift does not bias linear discriminant analysis [34]. Suppose that part of the useful information to recognize gases lies in the diffusion kinetics of the volatile molecules through the doped polymers, or the adsorption/desorption kinetic constants of the molecular gases on the doped materials under flow, the drift resistance may be a relevant information descriptor as it takes into account the first derivative of time (Figure 3a). Replicating the same methodology as previously for relative resistance modulation, PCA on α2 as an information descriptor was performed for the 18 different sub-datasets. As for α1, the evolution of the variance for the individual PC depends on the acquisition time (Figure 3b), mainly at the beginning of the transient response (typically the first 20 s). A singular difference with α1 is the degradation of the data variance on the first two PCs with the increase in acquisition time (from >66% for PCAα1170–179 down to <28% for PCAα2170–179). This was expected as the information becomes more sensitive to the signal’s noise on reaching the steady state; therefore, an increase in acquisition time decreases the contribution of the molecularly specific response compared in the overall information. We notice also in Figure 3b that significantly less variance is contained in PC2α2 compared to that in PC2α1, and more importantly that the variance in PC2α2 is far less dependent on the acquisition time than that in PC2α1. This suggests that despite the quality of the recognition, α2 seems to separate data mostly because of one single parameter (if a linear model is considered), while α1 seems to depend mostly on two parameters of a linear model. On the quality of the data separation (Figure 3c–e), similar trends in the (PC1,PC2) diagrams are observed depending on the acquisition time: in the first ten seconds of acquisition time after t0, acetone data are better separated from the rest, followed by the ethanol data. Confidence ellipsoids seem to be always overlapping one another; however, we observe that even until three minutes after time t0, the three volatile compounds are still recognizable, despite the residual dynamic of the sensing elements. It is also noticed that PCAα2170–179 (Figure 3e) clusters are more homogeneous than the ones of PCAα1170–179 (Figure 2e): smaller clusters within each ellipsoids can be distinguished in PCAα1170–179, which are attributed to experimental biases of the six repeated exposures for each class. The dependency of PCAα2 loadings on acquisition time displays different trends than the one of PCAα1 loadings (Figure 3f,g). For the first 20 s, all sensing elements except the ones coated with pristine P3HT and Ce(OTf)3-doped P3HT seem to have comparable contributions in PC1α2, and the Ce(OTf)3-doped P3HT-coated devices mostly constitute PC2α2–if t − t0 < 20 s. From 20 s after t0, Ce(OTf)3-doped P3HT-coated devices become the most significant in PC1α2, while PC2α2 loadings seem highly random if t − t0 < 20 s. This closer look on the PC1α2 and PC2α2 loadings highlights two different dynamics that successfully help in identifying the volatile compounds: before 20 s, “fast” and a “slow” dynamics yield the array response to recognize all three classes from PC1α2 and PC2α2. After 20 s, the fast dynamic of all sensing elements are faded away, leaving a minor contribution to the Ce(OTf)3-doped P3HT devices exclusively.

3.4. Steady vs. Dynamical Resistances for Gas Recognition

To qualitatively compare the total variance distributions over the different PCs with either α1 or α2 as information descriptors, the scree plots are displayed on two different graphs and show the effect of acquisition time after t0 for both descriptors (Figure 4a,b). Both of them confirm the trend of the variance with acquisition time for all principal components (not PC1 and PC2 exclusively). As Figure 4a shows, the increase in acquisition time after exposure at t0 tends to diminish the variance contained in the lowest PC in the case of considering α1 as an information descriptor, while it tends to increase the variance of the lower PC in the case of considering α2 as an information descriptor in Figure 4b. We observe that this diminishing rate is different for both descriptors as displayed in Figure 4a,b, which might come from bias between the linear ad hoc model computed by the PCA and the actual physical model that governs the dependency of the information descriptors on the different sensing elements. The position of the elbows on both plots (indicated by arrows on the scree plots in Figure 4a,b) confirms that most of the useful variance for PCAα1 is contained within the first three PCs, while only the first two for PCAα2. Particularly, for the drift resistance α2, this elbow is highly acquisition-dependent, and one may appreciate a richer model with more PCs to account for a shorter acquisition time. The variance analysis of PC1 and PC2 for both PCAα1 and PCAα2 shows an optimal acquisition time that contains most of the data variance in both PC1 and PC2 independently from the chosen information descriptor (Figure 4c). The histogram displayed in Figure 4c shows that if the acquisition time is set at the transient period between 10 and 19 s after t0, both information features can gather at least 70% of the total variance in two different biparametric linear models. As both information features show opposed trends in the PC1 + PC2 variance evolution with acquisition time, it was expected to observe an acquisition time optimum for both PCAs independently from the quality of the data clustering. To quantify the recognition performances of the sensing array with PCAα1 and PCAα2, the confidence ellipsoids determined by k-means clustering were used as a threshold to declare whether the given data belonged to a given class or not, defining seven classes (as displayed in Figure 4d). The relative area of the overlap for the confidence ellipsoids was evaluated for the different acquisition times and displayed in two different histograms for each descriptor in Figure 4d–f. For α1 (Figure 4e), one can observe that ellipsoids do not overlap from t − t0 > 20 s, while in the case of α2 (Figure 4f), confidence ellipsoids always overlap one another at any acquisition time. However, it should be noticed that the overlap in each case is only moderate: less than 20% of the total area of the ellipsoids. This suggests that recognition tests may be higher in the case where α1 is used as an information descriptor than for α2. To verify this, recognition tests were performed on the same sub-datasets using the 95% confidence ellipsoids as a threshold for classification (Figure 4g–i). Detection is considered as “positive” if the actual class of the PCA score is exclusively contained within the corresponding confidence ellipsoid, “uncertain” if the actual class of the PCA score is not exclusively contained within the corresponding confidence ellipsoid, and “negative” if the actual class of the PCA score is excluded from the corresponding confidence ellipsoid. On the one hand for α1 (Figure 4h), one can observe that the positive detection rate 99.4 ± 0.6% for acquisition time is higher than 20 s after t0. On the other hand for α2 (Figure 4i), one can observe that the positive detection rate >80% for acquisition time is higher, falling between 20 s and 149 s after t0. However, we noticed that using α2 as an information descriptor enables substantially better recognition than α1 for a short acquisition time below 10 s after t0: typically, 36% positive and 38% uncertain using α2, while 14% positive and 53% uncertain using α1. Overall, this statistical study shows that the different materials used in the array allow recognizing volatile organic compounds because of the relative resistance modulation (α1) rather than the resistance drift (α2), and it points out the thermodynamics of the current modulations rather than the kinetics involved in the process.

4. Discussion

Although this study focuses specifically on doped conducting polymers, the approach is generic to any conductometric eNose and could be used for different classes of conducting materials used as transducers in an array. Despite the fact that α1 and α2 are frequently used as information descriptors for such application, our study proposes to investigate which feature gathers most of the useful information for an eNose. However, we want to highlight that for practical applications, such a choice may rather be conditioned by the application than by the quality of the data per se. Specifically for the resistance modulation α1, the feature needs to routinely set a reference resistance R0 when the environment is considered “neutral” between each data sampling. This therefore requires one to have influence over the sample exposure to periodically regenerate the device from stimuli: it is therefore practically more convenient to use α1 as an information descriptor in cases where analytes can be exposed to a system in a supervised manner (to evaluate fragrance batch quality in an analytical laboratory, for instance). Such thermodynamic feature seems practically inadequate to extract for environment recognition in a case where the system is permanently exposed without periodic regeneration. For the drift resistance α2, the feature does not depend on a buffered reference. However, a steady value of the sampling period must be set to measure the dynamic: it is therefore a more adequate feature for online inference for environment recognition, on the condition that the sampling period is set a posteriori given the dynamic of the pattern to recognize. It is therefore a more adequate feature to recognize stimuli that are governed by a rhythm (for instance, pollution cycles in urban environments). In this study, the selection of these features α1 and α2 is not based on the practical convenience to compute them in a conductometric eNose application, but on the quality of the data they output for conducting polymers used in recognition tests, which conveniently enables computing both of them at will.
In such framework, PCA confirmed that the useful information in the current response of conducting polymer materials doped with different metal triflates and exposed to different volatile organic compounds lies specifically in the thermodynamics of the device resistance modulation and not its kinetics. The analysis of the relative resistance modulation α1 is more robust overtime when analyzed 20 s after each exposure at t0: the first two PCs explain at least two thirds of the data variance, allow optimal clustering without overlap, and enable 99.4 ± 0.6% recognition. Considering that all devices have the same geometry, mechanisms responsible for the modulation of the electrical resistivity are pointed out. As different doping effects were assessed by varying the nature of the dopant [22], the modulation of the doping yield by the interaction of different volatile molecules as electron-donating ligands on an electrophilic Lewis acid [35,36,37] is not invalidated by this analysis. The statistical approach does not allow discriminating between the effect on the charge carrier density or the charge carrier mobility for the different materials exposed to different gases. Despite the fact that material conductivity is a function of two independent physical properties (the charge carrier density and the charge carrier mobility), one cannot conclude that any of these properties are purely marking PC1α1 and PC2α1 since conductivity is the product of both and PCA scores are linear combinations of PCs. Such investigation could be carried out using the logarithmic value of the resistance modulation α3 = log(R/R0) as an information descriptor for the PCA in order to assess the individual contributions of the carrier mobility “μ” and carrier density “n” as principal components, since α3 = log(n0/n) + log(μ0/μ). Such validation would have to be confirmed by correlating the PCα3 loadings with a systematic study of carrier density and mobility assessment in another study that couples this approach with further material characterizations.

5. Conclusions

By statistical analysis, we evidenced that the relative resistance modulation of a doped conducting polymer array is a better information descriptor than the drift resistance, to recognize volatile organic compounds with different metal-triflate-doped polythiophene sensing elements. This shows that the information carrier for the molecular recognition using these materials in an electronic nose is rather linked to the thermodynamic equilibrium that yields the doped conducting polymers’ conductivity with its environment, rather than kinetic limitations which would select molecular targets based on their drift diffusion through different doped polymers. Despite this result suggesting to wait for a complete steady state for each sensing element response, we evidenced that the thermodynamic properties for recognizing molecular gases lie already in the transience of the material response. Despite the materials reaching a fast quasi-steady response within three minutes of environment exposures, principal component analysis and k-means clustering show that the recognition is already higher than 99% after only 20 s of gas exposures. This suggests that the training time to recognize a molecular environment can be greatly optimized for practical applications, even in the case where conducting materials in an electronics nose show a slow response. Applied to practical recognition with an eNose, this approach of finding a preliminary optimum for data acquisition before a test would greatly increase information throughputs for using eNose for chemical quality assessments by reducing the training duration while preserving most of the data quality to train an eNose to recognize various volatile molecular patterns.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/eng4040141/s1.

Author Contributions

Conceptualization, S.P.; Methodology, S.P.; Validation, W.H.A.; Formal analysis, W.H.A. and A.B. (Aicha Boujnah); Investigation, W.H.A., A.B. (Aicha Boujnah) and S.P.; Writing–original draft, W.H.A. and S.P.; Writing–review & editing, A.B. (Aicha Boujnah), A.B. (Aimen Boubaker), A.K. and K.L.; Visualization, S.P.; Supervision, S.P.; Project administration, A.B. (Aimen Boubaker), A.K. and S.P.; Funding acquisition, A.B. (Aimen Boubaker), A.K. and S.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the French Research Agency (ANR) for the “Sensation” project, grant number ANR-22-CE24-0001-01. The APC was paid by the editor.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author, SP, upon reasonable request.

Acknowledgments

The authors thank the French Nanofabrication Network RENATECH for financial support of the IEMN cleanroom.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Persaud, K.C.; Pisanelli, A.M.; Evans, P. Medical diagnostics and health monitoring. In Handbook of Machine Olfaction: Electronic Nose Technology; WILEY-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2002; pp. 445–460. [Google Scholar] [CrossRef]
  2. Karakaya, D.; Ulucan, O.; Turkan, M. Electronic nose and its applications: A survey. Int. J. Autom. Comput. 2020, 17, 179–209. [Google Scholar] [CrossRef]
  3. Cheng, L.; Meng, Q.-H.; Lilienthal, A.J.; Qi, P.-F. Development of compact electronic noses: A review. Meas. Sci. Technol. 2021, 32, 062002. [Google Scholar] [CrossRef]
  4. Illahi, A.A.C.; Dadios, E.P.; Bandala, A.A.; Vicerra, R.R.P. Electronic Nose Technology and Application: A Review. In Proceedings of the 2021 IEEE 13th International Conference on Humanoid, Nanotechnology, Information Technology, Communication and Control, Environment, and Management (HNICEM), Manila, Philippines, 28–30 November 2021. [Google Scholar] [CrossRef]
  5. Liu, T.; Guo, L.; Wang, M.; Su, C.; Wang, D.; Dong, H.; Chen, J.; Wu, W. Review on Algorithm Design in Electronic Noses: Challenges, Status, and Trends. Intell. Comput. 2023, 2, 0012. [Google Scholar] [CrossRef]
  6. Ruddigkeit, L.; van Deursen, R.; Blum, L.C.; Reymond, J.-L. Enumeration of 166 Billion Organic Small Molecules in the Chemical Universe Database GDB-17. J. Chem. Inf. Model. 2012, 52, 2864–2875. [Google Scholar] [CrossRef]
  7. Bühlmann, S.; Reymond, J.-L. ChEMBL-likeness score and database GDBChEMBL. Front. Chem. 2020, 8, 46. [Google Scholar] [CrossRef]
  8. Hierlemann, A.; Gutierrez-Osuna, R. Higher-Order Chemical Sensing. Chem. Rev. 2008, 108, 563–613. [Google Scholar] [CrossRef]
  9. Peveler, W.J.; Yazdani, M.; Rotello, V.M. Selectivity and Specificity: Pros and Cons in Sensing. ACS Sens. 2016, 1, 1282–1285. [Google Scholar] [CrossRef]
  10. Parastar, H.; Kirsanov, D. Analytical Figures of Merit for Multisensor Arrays. ACS Sens. 2020, 5, 580–587. [Google Scholar] [CrossRef]
  11. Alsaedi, B.S.; McGraw, C.M.; Schaerf, T.M.; Dillingham, P.W. Multivariate limit of detection for non-linear sensor arrays. Chemom. Intell. Lab. Syst. 2020, 201, 104016. [Google Scholar] [CrossRef]
  12. Hatfield, J.; Neaves, P.; Hicks, P.; Persaud, K.; Travers, P. Towards an integrated electronic nose using conducting polymer sensors. Sens. Actuators B Chem. 1994, 18, 221–228. [Google Scholar] [CrossRef]
  13. Chiu, S.-W.; Tang, K.-T. Towards a Chemiresistive Sensor-Integrated Electronic Nose: A Review. Sensors 2013, 13, 14214–14247. [Google Scholar] [CrossRef]
  14. Park, S.Y.; Kim, Y.; Kim, T.; Eom, T.H.; Kim, S.Y.; Jang, H.W. Chemoresistive materials for electronic nose: Progress, perspectives, and challenges. InfoMat 2019, 1, 289–316. [Google Scholar] [CrossRef]
  15. Sierra-Padilla, A.; García-Guzmán, J.J.; López-Iglesias, D.; Palacios-Santander, J.M.; Cubillana-Aguilera, L. E-Tongues/Noses Based on Conducting Polymers and Composite Materials: Expanding the Possibilities in Complex Analytical Sensing. Sensors 2021, 21, 4976. [Google Scholar] [CrossRef]
  16. Bedoui, S.; Faleh, R.; Samet, H.; Kachouri, A. Electronic nose system and principal component analysis technique for gases identification. In Proceedings of the 10th International Multi-Conferences on Systems, Signals & Devices 2013 (SSD13), Hammamet, Tunisia, 18–21 March 2013. [Google Scholar] [CrossRef]
  17. Yin, Y.; Zhao, Y. A feature selection strategy of E-nose data based on PCA coupled with Wilks Λ-statistic for discrimination of vinegar samples. J. Food Meas. Charact. 2019, 13, 2406–2416. [Google Scholar] [CrossRef]
  18. Bakar, M.; Abdullah, A.; Mustafa, W.; Razali, Z.; Saidi, S.; Kader, M.; Aman, M. Electronic Nose Testing for Confined Space Application Utilizes Principal Component Analysis and Support Vector Machine. IOP Conf. Ser. Mater. Sci. Eng. 2020, 932, 012072. [Google Scholar] [CrossRef]
  19. Nagle, H.T.; Schiffman, S.S. Electronic Taste and Smell: The Case for Performance Standards [Point of View]. Proc. IEEE 2018, 106, 1471–1478. [Google Scholar] [CrossRef]
  20. Pecqueur, S.; Talamo, M.M.; Guérin, D.; Blanchard, P.; Roncali, J.; Vuillaume, D.; Alibart, F. Neuromorphic Time-Dependent Pattern Classification with Organic Electrochemical Transistor Arrays. Adv. Electron. Mater. 2018, 4, 1800166. [Google Scholar] [CrossRef]
  21. Pecqueur, S.; Vuillaume, D.; Crljen; Lončarić, I.; Zlatić, V. A Neural Network to Decipher Organic Electrochemical Transistors’ Multivariate Responses for Cation Recognition. Electron. Mater. 2023, 4, 80–94. [Google Scholar] [CrossRef]
  22. Boujnah, A.; Boubaker, A.; Kalboussi, A.; Lmimouni, K.; Pecqueur, S. Mildly-doped polythiophene with triflates for molecular recognition. Synth. Met. 2021, 280, 116890. [Google Scholar] [CrossRef]
  23. Metsalu, T.; Vilo, J. ClustVis: A web tool for visualizing clustering of multivariate data using Principal Component Analysis and heatmap. Nucleic Acids Res. 2015, 43, W566–W570. [Google Scholar] [CrossRef]
  24. Peng, G.; Trock, E.; Haick, H. Detecting Simulated Patterns of Lung Cancer Biomarkers by Random Network of Single-Walled Carbon Nanotubes Coated with Nonpolymeric Organic Materials. Nano Lett. 2008, 8, 3631–3635. [Google Scholar] [CrossRef]
  25. Kumar, S.; Pavelyev, V.; Mishra, P.; Tripathi, N. Thin film chemiresistive gas sensor on single-walled carbon nanotubes-functionalized with polyethylenimine (PEI) for NO2 gas sensing. Bull. Mater. Sci. 2020, 43, 61. [Google Scholar] [CrossRef]
  26. Kanaparthi, S.; Singh, S.G. Reduction of the Measurement Time of a Chemiresistive Gas Sensor Using Transient Analysis and the Cantor Pairing Function. ACS Meas. Sci. Au 2021, 2, 113–119. [Google Scholar] [CrossRef]
  27. Freddi, S.; Vergari, M.; Pagliara, S.; Sangaletti, L. A Chemiresistor Sensor Array Based on Graphene Nanostructures: From the Detection of Ammonia and Possible Interfering VOCs to Chemometric Analysis. Sensors 2023, 23, 882. [Google Scholar] [CrossRef]
  28. Boujnah, A.; Boubaker, A.; Pecqueur, S.; Lmimouni, K.; Kalboussi, A. An electronic nose using conductometric gas sensors based on P3HT doped with triflates for gas detection using computational techniques (PCA, LDA, and kNN). J. Mater. Sci. Mater. Electron. 2022, 33, 27132–27146. [Google Scholar] [CrossRef]
  29. Li, J.; Luan, B.; Lam, C. Resistance drift in phase change memory. In Proceedings of the 2012 IEEE International Reliability Physics Symposium (IRPS), Anaheim, CA, USA, 15–19 April 2012. [Google Scholar] [CrossRef]
  30. Paredes-Madrid, L.; Matute, A.; Bareño, J.O.; Vargas, C.A.P.; Velásquez, E.I.G. Underlying Physics of Conductive Polymer Composites and Force Sensing Resistors (FSRs). A Study on Creep Response and Dynamic Loading. Materials 2017, 10, 1334. [Google Scholar] [CrossRef]
  31. Chen, Y.; Sun, L.; Zhou, Y.; Zewdie, G.M.; Deringer, V.L.; Mazzarello, R.; Zhang, W. Chemical understanding of resistance drift suppression in Ge–Sn–Te phase-change memory materials. J. Mater. Chem. C 2020, 8, 71–77. [Google Scholar] [CrossRef]
  32. Pries, J.; Stenz, C.; Schäfer, L.; Gutsche, A.; Wei, S.; Lucas, P.; Wuttig, M. Resistance Drift Convergence and Inversion in Amorphous Phase Change Materials. Adv. Funct. Mater. 2022, 32, 2207194. [Google Scholar] [CrossRef]
  33. Müller, G.; Sberveglieri, G. Origin of Baseline Drift in Metal Oxide Gas Sensors: Effects of Bulk Equilibration. Chemosensors 2022, 10, 171. [Google Scholar] [CrossRef]
  34. Kiselev, I.; Sysoev, V.; Kaikov, I.; Koronczi, I.; Tegin, R.A.A.; Smanalieva, J.; Sommer, M.; Ilicali, C.; Hauptmannl, M. On the Temporal Stability of Analyte Recognition with an E-Nose Based on a Metal Oxide Sensor Array in Practical Applications. Sensors 2018, 18, 550. [Google Scholar] [CrossRef]
  35. Pecqueur, S. Lewis Acid-Base Theory Applied on Evaluation of New Dopants for Organic Light-Emitting Diodes; Friedrich-Alexander-Universitaet Erlangen-Nuernberg: Erlangen, Germany, 2014. [Google Scholar]
  36. Pecqueur, S.; Maltenberger, A.; Petrukhina, M.A.; Halik, M.; Jaeger, A.; Pentlehner, D.; Schmid, G. Wide Band-Gap Bismuth-based p-Dopants for Opto-Electronic Applications. Angew. Chem. Int. Ed. 2016, 55, 10493–10497. [Google Scholar] [CrossRef]
  37. Ferchichi, K.; Bourguiga, R.; Lmimouni, K.; Pecqueur, S. Concentration-control in all-solution processed semiconducting polymer doping and high conductivity performances. Synth. Met. 2020, 262, 116352. [Google Scholar] [CrossRef]
Figure 1. Classifier Structure for Solvent Vapor Recognition. (a), Data classifier used in this study for unsupervised classification by using principal component analysis and k-means clustering on dynamical datasets generated by sensing elements composed of different materials. (bd), Pictures of the data acquisition hardware at different scales: photograph of seven arrays at the scale where each material was deposited (b), microscopic picture of the active area of an array gathering 16 sensing elements (c), microscopic picture of a single sensing element featuring a sub-micrometric conducting channel (d). (c), Example of raw current signal output i(t) from one single sensing element coated with one specific material, exposed to different sequences of solvent vapor exposures. (f), Zoom-in on one time sequence when three different vapors affect the sensing elements conductance between purge sequences (zoom-in (e) trace marked by a dashed square window). (e,f), Blue, green and red sequences depict the sensing element current output when water, acetone and ethanol vapors are respectively blown on the sensing elements with compressed air. Grey sequences depict the sensing element current output when clean compressed air is blown on the sensing elements. (g), Resistance value R(t) calculated from the raw current signal output i(t). Parameters R-R0 and dR/dt for calculating the information features α1(t) and α2(t) are depicted as the absolute resistance modulation and the slope of the resistance variation over time, extrapolated from the values of ten resistance data points in each case.
Figure 1. Classifier Structure for Solvent Vapor Recognition. (a), Data classifier used in this study for unsupervised classification by using principal component analysis and k-means clustering on dynamical datasets generated by sensing elements composed of different materials. (bd), Pictures of the data acquisition hardware at different scales: photograph of seven arrays at the scale where each material was deposited (b), microscopic picture of the active area of an array gathering 16 sensing elements (c), microscopic picture of a single sensing element featuring a sub-micrometric conducting channel (d). (c), Example of raw current signal output i(t) from one single sensing element coated with one specific material, exposed to different sequences of solvent vapor exposures. (f), Zoom-in on one time sequence when three different vapors affect the sensing elements conductance between purge sequences (zoom-in (e) trace marked by a dashed square window). (e,f), Blue, green and red sequences depict the sensing element current output when water, acetone and ethanol vapors are respectively blown on the sensing elements with compressed air. Grey sequences depict the sensing element current output when clean compressed air is blown on the sensing elements. (g), Resistance value R(t) calculated from the raw current signal output i(t). Parameters R-R0 and dR/dt for calculating the information features α1(t) and α2(t) are depicted as the absolute resistance modulation and the slope of the resistance variation over time, extrapolated from the values of ten resistance data points in each case.
Eng 04 00141 g001
Figure 2. Analysis of the Relative Resistance Modulation as Information Descriptor. (a), Extraction of R0 (in pink, as average for ten resistance values recorded between 11 and 20 s before vapor exposure, starting at t0) and R values (in green as average for ten resistance values recorded between t and t + 10 s every 10 s after t0) to evaluate the feature α1 = (R − R0)/R0. (b), PC variance for different PCAs on α1, where R is measured at different time intervals [t;t + 10 s] after t0 (the tricolor gradient shades linearly between the minimum and maximum value of the whole table). (ce), PCA scores on α1, where R is measured at different time intervals [0 s;9 s] (c), [10 s;19 s] (d) and [170 s;179 s] (e) after t0 (confidence ellipsoids are set to 95% probability). (f,g), Squared loadings of PC1 (f) and PC2 (g) for different PCA on α1, where R is measured at different time intervals [t;t + 10 s] after t0 (the bicolor gradient shades linearly between the minimum and maximum value of the whole tables).
Figure 2. Analysis of the Relative Resistance Modulation as Information Descriptor. (a), Extraction of R0 (in pink, as average for ten resistance values recorded between 11 and 20 s before vapor exposure, starting at t0) and R values (in green as average for ten resistance values recorded between t and t + 10 s every 10 s after t0) to evaluate the feature α1 = (R − R0)/R0. (b), PC variance for different PCAs on α1, where R is measured at different time intervals [t;t + 10 s] after t0 (the tricolor gradient shades linearly between the minimum and maximum value of the whole table). (ce), PCA scores on α1, where R is measured at different time intervals [0 s;9 s] (c), [10 s;19 s] (d) and [170 s;179 s] (e) after t0 (confidence ellipsoids are set to 95% probability). (f,g), Squared loadings of PC1 (f) and PC2 (g) for different PCA on α1, where R is measured at different time intervals [t;t + 10 s] after t0 (the bicolor gradient shades linearly between the minimum and maximum value of the whole tables).
Eng 04 00141 g002
Figure 3. Analysis of the Resistance Drift as Information Descriptor. (a), Extraction of dR/dt (red slope linear regression over nine resistance values centered on R) and R values (in green) to evaluate the feature α2 = 1/R ∗ dR/dt. (b), PC variance for different PCAs on α2, where R is measured at different time intervals [t;t + 10 s] after t0 (the tricolor gradient shades linearly between the minimum and maximum value of the whole table). (ce), PCA scores on α2, where R is measured at different time intervals [0 s;9 s] (c), [10 s;19 s] (d) and [170 s;179 s] (e) after t0 (confidence ellipsoids are set to 95% probability). (f,g), Squared loadings of PC1 (f) and PC2 (g) for different PCAs on α2, where R is measured at different time intervals [t;t + 10 s] after t0 (the bicolor gradient shades linearly between the minimum and maximum values of the whole tables).
Figure 3. Analysis of the Resistance Drift as Information Descriptor. (a), Extraction of dR/dt (red slope linear regression over nine resistance values centered on R) and R values (in green) to evaluate the feature α2 = 1/R ∗ dR/dt. (b), PC variance for different PCAs on α2, where R is measured at different time intervals [t;t + 10 s] after t0 (the tricolor gradient shades linearly between the minimum and maximum value of the whole table). (ce), PCA scores on α2, where R is measured at different time intervals [0 s;9 s] (c), [10 s;19 s] (d) and [170 s;179 s] (e) after t0 (confidence ellipsoids are set to 95% probability). (f,g), Squared loadings of PC1 (f) and PC2 (g) for different PCAs on α2, where R is measured at different time intervals [t;t + 10 s] after t0 (the bicolor gradient shades linearly between the minimum and maximum values of the whole tables).
Eng 04 00141 g003
Figure 4. Comparison of Both Descriptors. (a,b), Scree plots for the PCA performed with different acquisition times, from the first ten seconds after t0 (“0–9” in blue) to the last ten seconds after t0 (“170–179” in red), performed with either the relative resistance modulation—α1 (a) or the drift resistance—α2 (b) as an information descriptor. The arrows indicate the reduced number of component to consider by the elbow method. (c) Explained variance for PC1 and PC2 with PCA performed for different acquisition times with either α1 or α2. (df), Overlap area of the three 95%-confidence ellipsoids in (PC1,PC2), for PCA performed for different acquisition times with either α1 or α2. The areas were estimated by discretizing the different regions where the three ellipsoids are segmented in the smallest square of 500 × 500 pixels. In (d) is displayed the example for the PCA with α2 for the “0–9” acquisition time, and the histograms of the ellipsoids relative area as a function of acquisition time for α1 and α2 are respectively displayed in (e,f). (gi), Detection tests of the PCA scores for PCA performed for different acquisition times with either α1 or α2. The test is considered positive, uncertain and negative, respectively when scores are exclusively inside, not-exclusively inside and outside the corresponding 95% confidence ellipsoid. In (g) is displayed the example for the PCA with α2 for the “0–9” acquisition time, and the graphs of the detection rate as a function of acquisition time for α1 and α2 are respectively displayed in (h,i).
Figure 4. Comparison of Both Descriptors. (a,b), Scree plots for the PCA performed with different acquisition times, from the first ten seconds after t0 (“0–9” in blue) to the last ten seconds after t0 (“170–179” in red), performed with either the relative resistance modulation—α1 (a) or the drift resistance—α2 (b) as an information descriptor. The arrows indicate the reduced number of component to consider by the elbow method. (c) Explained variance for PC1 and PC2 with PCA performed for different acquisition times with either α1 or α2. (df), Overlap area of the three 95%-confidence ellipsoids in (PC1,PC2), for PCA performed for different acquisition times with either α1 or α2. The areas were estimated by discretizing the different regions where the three ellipsoids are segmented in the smallest square of 500 × 500 pixels. In (d) is displayed the example for the PCA with α2 for the “0–9” acquisition time, and the histograms of the ellipsoids relative area as a function of acquisition time for α1 and α2 are respectively displayed in (e,f). (gi), Detection tests of the PCA scores for PCA performed for different acquisition times with either α1 or α2. The test is considered positive, uncertain and negative, respectively when scores are exclusively inside, not-exclusively inside and outside the corresponding 95% confidence ellipsoid. In (g) is displayed the example for the PCA with α2 for the “0–9” acquisition time, and the graphs of the detection rate as a function of acquisition time for α1 and α2 are respectively displayed in (h,i).
Eng 04 00141 g004
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

Haj Ammar, W.; Boujnah, A.; Boubaker, A.; Kalboussi, A.; Lmimouni, K.; Pecqueur, S. Steady vs. Dynamic Contributions of Different Doped Conducting Polymers in the Principal Components of an Electronic Nose’s Response. Eng 2023, 4, 2483-2496. https://doi.org/10.3390/eng4040141

AMA Style

Haj Ammar W, Boujnah A, Boubaker A, Kalboussi A, Lmimouni K, Pecqueur S. Steady vs. Dynamic Contributions of Different Doped Conducting Polymers in the Principal Components of an Electronic Nose’s Response. Eng. 2023; 4(4):2483-2496. https://doi.org/10.3390/eng4040141

Chicago/Turabian Style

Haj Ammar, Wiem, Aicha Boujnah, Aimen Boubaker, Adel Kalboussi, Kamal Lmimouni, and Sébastien Pecqueur. 2023. "Steady vs. Dynamic Contributions of Different Doped Conducting Polymers in the Principal Components of an Electronic Nose’s Response" Eng 4, no. 4: 2483-2496. https://doi.org/10.3390/eng4040141

Article Metrics

Back to TopTop