Next Article in Journal
Design of Meat Product Safety Information Chain Traceability System Based on UHF RFID
Previous Article in Journal
Temperature Dependence of the Beating Frequency of hiPSC-CMs Using a MEMS Force Sensor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Dynamic Bayesian Model for Detecting Obstructive Respiratory Events by Using an Experimental Model †

1
ESAII Department, Universitat Politècnica de Catalunya—BarcelonaTech (UPC), 08019 Barcelona, Spain
2
Institute for Bioengineering of Catalonia (IBEC-BIST), 08028 Barcelona, Spain
3
CIBER of Bioengineering, Biomaterials and Nanomedicine (CIBER-BBN), 28029 Madrid, Spain
*
Author to whom correspondence should be addressed.
This manuscript is extension version of the conference paper: Romero, D.; Jané, R. Detecting Obstructive Apnea Episodes using Dynamic Bayesian Networks and ECG-based Time-Series. In Proceedings of the 44th Annual International Conference of the IEEE Engineering in Medicine & Biology Society (EMBC’2022), Glasgow, UK, 11–14 July 2022.
Sensors 2023, 23(7), 3371; https://doi.org/10.3390/s23073371
Submission received: 6 February 2023 / Revised: 17 March 2023 / Accepted: 21 March 2023 / Published: 23 March 2023
(This article belongs to the Section Biomedical Sensors)

Abstract

:
In this study, we propose a model-based tool for the detection of obstructive apnea episodes by using ECG features from a single lead channel. Several sequences of recurrent apnea were provoked in separate 15-min periods in anesthetized rats during an experimental model of obstructive sleep apnea (OSA). Morphology-based ECG markers and the beat-to-beat interval ( R R ) were assessed in each sequence. These markers were used to train dynamic Bayesian networks (DBN) with different orders and feature combinations to find a good tradeoff between network complexity and apnea-detection performance. By using a filtering approach, the resulting DBNs were used to infer the apnea probability signal for subsequent episodes in the same rat. These signals were then processed using by 15-s epochs to determine whether epochs were classified as apneic or nonapneic. Our results showed that fifth-order models provided suitable RMSE values, since higher order models become significantly more complex and present worse generalization. A global threshold of 0.2 gave the best overall performance for all combinations tested, with A c c = 81.3%, S e = 69.8% and S p = 81.5%, using only two parameters including the RR and D s (R-wave downslope) markers. We concluded that multivariate models using DBNs represent a powerful tool for detecting obstructive apnea episodes in short segments, which may also serve to estimate the number of total events in a given time period.

1. Introduction

Recurrent apnea during a patient’s sleep is caused by repeated episodes of obstructive sleep apnea (OSA), resulting in prolonged exposure to intermittent hypoxia (IH). This chronic condition has been associated with some cardiovascular consequences, including among others, systemic hypertension, heart failure, coronary artery disease, and stroke [1,2,3]. At the same time, OSA patients usually present excessive daytime sleepiness and nonrestorative sleep, fatigue, impaired learning, and significant social problems due to poor mental performance.
Nowadays, overnight polysomnography (PSG) is considered the gold standard for OSA diagnosis and detection, which includes multiple physiological signals such as electroencephalogram (EEG), electrooculogram (EOG), electrocardiogram (ECG), electromyogram (EMG), and blood oxygen level (SpO 2 ), among others [1]. However, analysis of PSG recordings is very expensive, time consuming, and uncomfortable for the patient, limiting its accessibility and suitability for long-term home monitoring or ambulatory systems [2,3,4]. Recent efforts have been made to propose lighter systems with few sensors or a single channel, but they focus mainly on signal acquisition [5,6,7]. Likewise, other systems that analyze signals in real time often have insufficient accuracy for medical diagnosis [8]. However, in recent years many promising algorithms for offline OSA detection and diagnosis have been implemented with proven performance, showing great potential for their implementation on lighter hardware and wearable devices [9].
A variety of algorithms typically use a set of features extracted from PSG recordings and apply machine learning, neural network, or deep learning techniques for OSA detection or classification as reported in [9]. Among them, some use a single-source sensor for SpO 2 [10,11,12], respiration [13,14,15], or sound [16,17,18] for this purpose. Others, however, use ECG channels and have been demostrated to achieve the highest classification accuracy [9], followed by oxygen saturation, respiration, and sound-based methods. The ECG-derived algorithms apply different approaches based on ECG waveform analysis, including ECG signal decomposition through different wavelet transforms [19,20,21], variational mode decomposition [22], and empirical mode decomposition [23].
Methods based on the analysis of heart rate variability (HRV) and beat-to-beat ( R R ) interval time series [24,25] include temporal and spectral features, time frequency-based parameters, and some common complexity measures. Moreover, ECG-derived respiratory (EDR) signals obtained from breathing modulation on the T and R wave amplitudes are also combined with heart rate to reliably detect sleep apnea. Studies carried out by Penzel et al. [26], de Chazal et al. [27], and Méndez et al. [28] represent some examples that use features extracted from these EDR signals. The above methods rely on the fact that during apnea, the muscles of the respiratory system increase the mechanical effort to overcome the obstruction, thus affecting the surface ECG. This additional respiratory effort can impact other peripheral systems such as the cardiovascular system. In general, the patient’s respiratory information can be extracted from the fluctuations observed in the R R interval time series, namely HRV, and from the morphological changes caused by the respiratory modulation, by analyzing the EDR signals [29]. However, despite their proven higher accuracy, most algorithms using ECG signals are based on segmented intervals or epochs, most typically 1 min in duration, and many handcrafted features to train the models and perform the prediction task in an offline fashion. Although the ideal algorithm would involve automatic ECG feature extraction along with the use of shorter segments [30], even of varying size, the development of simpler models with finer time resolution could be a valid starting point for classifying shorter epochs or even a single apnea episode, as compared to the longer periods commonly used in epoch-based detection approaches.
The aim of this study was to estimate the occurrence probability of apnea episodes based on the effects caused by airway obstruction on various ECG markers. A preliminary version of this work has been reported in [31]. Specifically, we exploited the subtle changes that occur in heart rate and morphology-based ECG markers during individual episodes of apnea by using an experimental rat model of OSA. To this end, beat-to-beat time series of the aforementioned markers, together with apnea occurrence information, were used to train low-order dynamic Bayesian networks (DBNs) able to accurately detect short apnea-related epochs of only 15 s.

2. Materials and Methods

2.1. Population Data

Five healthy male Sprague–Dawley rats (mean weight: 437 ± 27 g) were intraperitoneally anesthetized with urethane (1 g/1 kg). The rats were connected to a system that controls a nasal mask with one tube open to the atmosphere and another connected to a positive pressure pump, thus preventing the animal from rebreathing. OSA episodes were simulated by obstructing the airways in the tubes by using electrovalves controlled by BIOPAC Systems (Goleta, CA, USA) equipment and its associated software. The electrodes and nasal mask were placed on the rats after the animals had been anesthetized and shaved.

2.2. OSA Model

Three recurrent apnea sequences were induced for 15-min intervals, preceded and followed by 15-min periods of normal breathing. A frequency (F) of 20, 40, or 60 apneas/hour was used for each sequence, simulating different levels of severity, while the duration of the apnea episodes was fixed at 15 s. The order of the apnea sequences as a function of F was randomized in each animal during the experiment. However, the number of sequences with the same F was evenly distributed within the population. Specifically, a total of 150 apneas of 15 s were provoked among all rats (30 apneas per rat), distributed over 15 separate recordings of 15 min each (three recordings per rat).
Several physiological signals were acquired with the BIOPAC system during spontaneous breathing and during the simulated apneas. These included respiratory flow, respiratory pressure, and electrocardiogram (ECG) signals (leads I and II) [32].

2.3. ECG Data Processing

The ECG signals were originally sampled at 1250 Hz, but then resampled at 10 kHz before further processing. This was performed to improve the time resolution for R-peak detection, since the heart rate range is significantly higher in rats (300–500 bpm) compared to humans [33]. The subsequent preprocessing steps included baseline drift attenuation via cubic spline interpolation [34], fourth-order bidirectional Butterworth low-pass filtering at 45 Hz to remove power-line interferences (50 Hz) and high-frequency noise, and, finally, automatic QRS complex detection [35] followed by a visual inspection to exclude noisy beats in noisy ECG segments.

2.3.1. Heart Rate Periods

The final locations of the R-wave peaks were determined from the nonfiltered (low-pass) ECG signals based on previously detected QRS complexes, and used to generate the R R interval time series analyzed in this study.

2.3.2. Morphology-Based ECG Markers

To enhance OSA detection, several ECG markers related to ventricular depolarization were also evaluated on nonfiltered ECG signals, thus providing relevant information about breathing influence on ECG waves morphology. These markers included the amplitude of the R and S waves ( R a , S a ), the difference between R a and S a ( R S d ), and the upstroke and downstroke slopes of the R wave, denoted by U S and D S , respectively. U S and D S were obtained as a result of fitting two straight lines over the ECG signal in specific segments within the QRS complex. Further details about the calculation of the QRS slope are described in [36].
The previously described markers were also used to obtain a combined marker by using principal components analysis (PCA). In brief, the beat-to-beat time series from the depolarization markers, R a ( t ) , S a ( t ) , D S ( t ) , U S ( t ) , and R S d ( t ) , were entered as inputs in the PCA algorithm. Then, a new signal projected onto the first component, defined here as Q R S p c a ( t ) , was obtained to emphasize the apnea-related influence on the ECG. Finally, all single and combined marker time series are fed into the DBNs as described below.

2.4. Bayesian Networks Basics

Bayesian networks (BNs) are a type of probabilistic graphical model (PGM) representing the conditional independences among random variables with directed acyclic graphs (DAGs) [37,38]. BNs that can manage both discrete and continuous variables simultaneously are known as the so-called hybrid BNs, while those BNs containing only discrete or continuous nodes are called discrete BNs and Gaussian BNs (GBN), respectively. Each node in the DAG has an associated conditional probability distribution (CPD) that defines the probability distribution of the node given its parents in the DAG. In general, for a BN with N variables X = { X 1 , , X N } , the joint distribution factorizes as
P ( X ) = i = 1 N P ( X i | p a ( X i ) ) ,
where p a ( X i ) denotes the configuration of the set of parents of X i in the network. Note that this factorization enforces several conditional independence statements. For instance, any variable in a Bayesian network is conditionally independent of the other variables of the model given its Markov blanket (mb), which is denoted as X i X | m b ( X i ) for short. Specifically, in a hybrid BN, the models are constructed as a set of conditional linear Gaussian (CLG) distribution models [39]. Here, discrete nodes are not allowed to have continuous parents. Likewise, the conditional distribution of each discrete variable X X D given its parents is a multinomial, while the conditional distribution of each continuous variable Z X C with discrete parents Z D X D and continuous parents Z C X C , is given by
f ( z | p a ( z ) = { z D , z C } ) = N ( z ; α ( z D ) + β ( z D ) T z C , σ 2 ( z D ) )
for all Z D X D and Z C X C , where α and β are the coefficients of a linear regression model of Z given its continuous parents. Note that this model can differ for each configuration of the discrete variables Z D . After fixing any configuration of the discrete variables X D , the joint distribution of any subset of the continuous variables X C is a multivariate Gaussian.

2.5. Dynamic Bayesian Networks

The obstructive apnea episodes affect several physiological variables from different interacting systems including the cardiac, respiratory, and neural systems. These variables present different responses with regard to amplitude variations and time durations, which may help to detect the apnea occurrence if their relationships can be modeled. In our study, these variables are defined as univariate time series that affect one another via unknown relationships. Therefore, we aim to model these relationships by using a DBN that combines both continuous and categorical variables, to approximate their temporal interactions during both normal respiration and apnea episodes.
Usually, DBNs assume that the underlying process they model represents a first-order Markovian, indicating that future states are independent of the past, given the present. However, in many biological processes, this particular assumption needs to be relaxed to include certain past information to improve model prediction of such future states.
In DBN, the time is discretized into slices for a given period. This period is usually determined by the frequency of the collected signals. For each time slice, there is a static BN that has parents in the previous slices in addition to the actual intraslice parents. In our study, each time slice is associated with each heartbeat occurrence. Here, the joint probability distribution accounts for all time slices from a certain time T,
p ( X 0 : T ) = p { X 0 , X 1 , , X T } = p ( X 0 ) t = 0 T 1 p ( X t + 1 | X 0 : t ) ,
where X t = { X 1 t , X 2 t , , X n t } represents all the nodes in a time slice t for t = 0 , 1 , , T . In this equation, it is required that all the previous time slices be taken into account to calculate the product. This can be simplified by using the Markov assumption [40]. On the other hand, the Markovian order defines the number of time slices required to confidently assume that the present is independent of the past. By applying a first-order Markovian network, the resulting joint probability distribution will be
p ( X 0 : T ) = p ( X 0 ) t = 0 T 1 ( X t + 1 | X t ) .
To compute the joint density distribution over time, the nodes in the current time slice and the parent nodes in the previous time slice only have to be considered. On the other hand, it is commonly assumed that the structure in a DBN is homogeneous over time, meaning that the network structure does not change with t. Because of this, the network structure including the intra- and interslice arcs and the parameters are replicated for each unrolled time step, which also means that only interslice arcs are permitted from t 1 to t. Increasing the Markovian order implies more arcs appearing from earlier lags to the present and thus a greater complexity when learning the network structure and its parameters.

2.5.1. Structure Learning of DBN

The structure of the DBNs can be learned in two steps: (1) the intraslice arcs (static structure) of the network are learned with the max–min hill-climbing (MMHC) algorithm, (2) followed by learning the interslice arcs (transition structure). MMHC is a hybrid learning method that searches for possible network structures with a local search and then directs the arcs and scores the networks with the Bayesian information criterion (BIC) [41]. The local search is performed around each node to find its potential parents and children (i.e., its Markov blanket), which makes it conditionally independent of the other nodes. Specifically, a modified version of the MMHC algorithm named dynamic MMHC (DMMHC) was applied [42], in which the Markovian order of the network can also be modified in order to adjust the autoregressive order as needed.
Since the DBNs created here are CLG, the following assumptions are made: (a) discrete nodes have a multinomial distribution, (b) continuous nodes with no discrete nodes as parents have a normal distribution, (c) continuous nodes with at least one discrete node as a parent have a mixture of normal distributions, (d) all normal distributions in each mixture have a separate mean and variance, with each component having an independent set of parameters, and (e) discrete nodes can have only discrete nodes as parents, while continuous nodes can have both continuous and discrete parents [43].

2.5.2. Parameter Learning in DBN

Once the network structure is learned, the next step consists of estimating the network parameters for each node. This step is performed based on the maximum likelihood estimator (MLE), whose specific form depends on the parents having each node and the assumed distribution for each variable. For continuous nodes with only continuous parents, the parameters of the local distribution are the regression coefficients associated with the parents and the standard deviation (SD) of the residuals [43]. In the case of continuous nodes with at least one discrete parent, the local distribution is modeled with a mixture of linear regressions by fitting a separate linear model for each data subset associated with the different values contained in the discrete parent. Finally, for discrete nodes, the conditional probabilities given their parents are estimated by using empirical frequencies. The same parameter-learning approach can be applied to DBNs once the structure is learned.

2.5.3. Inference in DBN

After learning the structure and parameters of a DBN, it is possible to make inferences about any unobserved nodes or system states by providing some evidence to the network. For example, we can predict the most likely state of the system (i.e., whether there is an apnea episode or not) over the current interval. The evidence provided by past time slices should be used to predict the next time slice for a given node, either continuous or discrete, but also to estimate the actual value for some unknown parameter. In the case of Markovian orders higher than one, the prediction task consists of providing some evidence of the observed nodes for all time slices except the current one, and predicting the state of the desired nodes at time t,
p ( X i , t + Δ | Y : , 0 : t ) ,
where X i , t + Δ represents the predicted variable at time t + Δ , and Y : , 0 : t represents all the evidence from t = 0 to T. There are other special cases of interest such as the filtering, p ( X i , t | Y : , 0 : t ) , and the fixed lag smoothing, p ( X i , t τ | Y : , 0 : t ) , which are particularly considered in this study [44]. In such cases, X i , t and X i , t τ represent the ith hidden variable at time t and t τ , respectively.

2.6. Data Preprocessing

All time series T S ( t ) evaluated for our study were preprocessed before feeding into the DBN. For each marker, T S f ( t ) , a detrended version was obtained by removing a 20th-order polynomial trend resulting from the original time series to attenuate slow oscillations that affect the detection rate and overall performance of the models. Since we are only interested in detecting apnea episodes, we do not care about the actual amplitude of the different markers, which can vary substantially between recordings. Figure 1 shows an example of this detrending process, emphasizing the effect caused by the apneas.

2.7. Model Training and Validation

2.7.1. Univariate Conditional Lag-Based Models

To train the DBNs, we started with the first-order Markovian model ( L = 1) by using only one time series T S f ( t ) at a time ( T S f = { R R , R a , S a , D S , U S , R S d , Q R S p c a } ) in combination with the available apnea information, A p = { 1 , 0 } . Subsequently, the Markovian order was increased up to L = 15 to obtain more complex but better predictive models.

2.7.2. Multivariate Conditional Lag-Based Models

A strategy similar to that described above was followed to train DBNs with more than one marker. Specifically, all paired combinations between the R R time series and the remaining ECG markers were used as input data for DBNs. In these cases, the apnea-detection rate can be significantly improved at the expense of more complex networks in terms of total number of nodes and arcs but with smaller orders. Finally, a model including all ECG markers together (excluding Q R S p c a ) and the R R time series was also trained and compared to all previous simpler models. Figure 2 shows three examples of DBNs obtained for different combinations of parameters and fixing the lag value to L = 3, where the leftmost slice is the earliest ( t 3 ) and the rightmost slice represents the present time ( t 0 ).
The models were trained by using each 15-min recording of recurrent apnea simulated in each rat. This approach makes it possible to find the sequence that provides the best predictive model for a given animal. Model quality was assessed during model training by using both RMSE and Bayesian information criterion (BIC), which was also used to select the optimal lag for each network.

2.7.3. Model Assessment

With the DBNs trained in each animal and sequence, the next step was to recursively compute the filtering density p ( x t | y 1 : t ) , where y can be either the R R , any ECG parameter, or any combination of markers, while x represents the apnea occurrence at time t. The resulting conditional probability time series was then smoothed by using a 40-beat moving average filter to reduce potential spurious peaks.
An example of this methodology is shown in Figure 3 for a particular segment of several apnea episodes. In this figure, the normalized time series of the R R and Q R S p c a markers are plotted with the actual apnea information in the top graphs, while the estimated apnea probability time series is shown before and after the smoothing step in the bottom graphs. Note that during apneas, the smoothed probability peaks are significantly larger and wider when using two parameters compared to the univariate approach. This implies a significant improvement in detection sensitivity.
The smoothed probability signal was then split into 15-s segments, or epochs, from which the RMS values were calculated. These RMS values were used to determine whether a segment was classified as normal or apneic. Finally, performance metrics including sensitivity ( S e ), specificity ( S p ), accuracy ( A c c ), and area under the ROC curve ( A U C ) were calculated for all models as follows.
  • Ideal record-specific analysis: DBNs were trained and tested on each 15-min recording of recurrent apnea (15 recordings of 60 epochs each) available for all rats. Performance measures were first averaged for each individual rat and then averaged across all rats.
  • Record-specific analysis: DBNs were trained on the first recording of 15-min from each rat, and tested on the remaining two recordings from the same rat. The results from the testing phase were then averaged among all rats.
Finally, the average performance metrics were considered to select the optimal threshold and lag of the final network configuration for apnea detection.

3. Results

3.1. Exploring the Optimal Order Range for the DBNs

Figure 4 represents an example of the R R time series before and after removing the trend for a particular sequence of 15 apneas. As can be seen, several peaks are clearly observed on the smoothed probability signal (right side, bottom panel), corresponding to the timing of the occurrence of the apneas. However, the smoothed signal obtained before normalization (left side, bottom panel) follows a similar trend to that of the R R time series, significantly attenuating or masking the apnea-related peaks analyzed in further steps.
Regarding the optimal order range, Figure 5 illustrates the RMSE values obtained as a function of the network order for the R R and R a markers in three particular recordings. For both markers, the RMSE values for the middle and right columns decrease rapidly from L = 1 to L = 5 and then remain at similar values up to L = 15. However, a slightly different pattern is observed for the example in the left column, where the RMSE values decrease slowly up to L = 10. In all cases, a very similar pattern is reflected for the three apnea sequences of each recording. Notably, depending on the rat analyzed, the sequences with smaller RMSE values along different order values may suggest the best one to train the final record-specific model. Several potential factors may affect the overall quality of the trained models, in particular the noise and the number of apnea episodes simulated in the time series.

3.2. Comparing Univariate Models: Heart Rate Fluctuations vs. Morphology-Based ECG Models

Having explored the suitable range of values for the network order in univariate models, the next step consists of training and testing DBNs by using each individual marker. Table 1 summarizes the results obtained during model assessment for all univariate models by using two different orders, L = { 5 , 15 } .
From this table, it is clearly observed that the R R marker outperformed the results obtained for all morphology-based ECG markers in terms S e and S p . In the case of L = 5, the S e value for R R doubled those corresponding to the ECG markers. When using L = 15, the S e values increased by approximately 20–35% among the different ECG markers and about 30% for the R R marker. However, Q R S p c a , R a , and U s presented S p and A c c estimates similar to those of R R . When testing the networks in the record-specific analysis by using the sequences not seen during training, the performance metrics decreased overall. In this case, the U S resulted in the best marker with S e = 43.3%, S p = 76.8%, and A c c = 74.7%.

3.3. Multivariate Models: Combining ECG Parameters with RR Interval

Each individual morphology-based ECG marker was included as an additional input data to obtain different paired combinations. For instance, Figure 6 shows the evolution of the networks’ structure size for the combination of R R and R a as a function of the network order. An almost linear increase in the total number of arcs is observed as the Markovian order increases. For the three examples shown in the figure, remarkable differences arise from L > 4 , but the smaller orders are quite similar. Note, however, that the number of nodes remains equal in the three sequences for all cases, while the number of arcs can vary significantly depending on the analyzed recording.
Figure 7 (left and middle panels) shows the estimates of A c c , S e , and S p obtained for two univariate models using R R and R a , and the corresponding paired combination (right panel) in the ideal record-specific analysis. Different lags and thresholds ( δ = { 0.1 , 0.2 , , 0.9 } ) were tested for the apnea-detection task in each epoch against their R M S estimates. As expected, the larger the threshold δ , the smaller the S e values in favor of higher S p estimates. Note that the A c c and S p values increased more rapidly up to thresholds of 0.3–0.5, depending on the lag L , but then remained fairly stable for larger values of δ . The curves obtained for S e by using the R a marker alone showed a worse performance compared to the R R marker, even for smaller δ and larger L values.
Although univariate R R -based models showed very good overall performance, smaller orders yielded similar results when combining the R R and R a markers (see Figure 7, right panel). The latter required only an order of L 3 to achieve performance metrics similar to those obtained with L = 10 by using the R R marker alone. Table 2 summarizes the average results obtained during model evaluation by using all paired combinations for the ideal record-specific analysis. In these cases, with L = 5, we obtained results that were comparable to those with L = 15, with DBNs that are significantly less complex. The estimates of the performance metrics ranged from 77.4–83.7% for S e , 78.5–80.2% for S p , and 80.3–81.8% for A c c , with S a and U S showing the best results. The results obtained when testing the paired combinations in the record-specific analysis decreased overall for all metrics, ranging from 57.2–70.6% for S e , 69.0–81.5 % for S p , and 70.7–82.1 % for A c c . In this analysis, the best markers to combine with R R were R S d with S e = 64.4%, S p = 79.7%, and A c c = 79.3%, and D S with S e = 68.9%, S p = 81.5%, and A c c = 81.3%.
Note that the Q R S p c a marker was also combined with the R R marker and had a similar performance to other combinations, but only in the ideal record-specific analysis. On the other hand, when all ECG markers were used together, the observed improvement in performance metrics was not relevant enough compared to the paired combinations with the R R marker. The performance metrics for this case in the record specific analysis for L = 5 were S e = 69.4%, S p = 78.4%, and A c c = 78.8%.

4. Discussion

In this study, we proposed a predictive model to detect obstructive apnea episodes by using heart rate information and morphological features obtained from a single ECG channel. The proposed model, based on DBN, was implemented and trained with ECG data obtained during an experimental model of obstructive sleep apnea in rats, where several sequences of recurrent apneas were simulated in a well-controlled protocol. In this sense, we explored different combinations of parameters and network configurations in order to (1) minimize the amount of prior information used to train the network, and (2) determine the best combination of physiological variables that will become part of the final structure. This was achieved by optimizing the network order (i.e., the number of previous heartbeats considered or lag) and searching for the best combination of features and optimal threshold that provides the best detection performance. Specifically, we analyzed the probability signal provided by the network and classified each 15-s epoch as apneic or normal breathing. The nature of our experimental protocol allowed us to use such short intervals, unlike the conventional 1-min epochs used in many patient studies, as well summarized in [9]. The epoch length can be reduced if necessary, since the final DBN is composed of only five consecutive heartbeats.
Several findings can be highlighted from the results obtained. First, the higher the Markovian order (lag) used for the trained DBNs, the better the sensitivity estimates, at the expense of higher model complexity. However, beyond a certain order, usually above 10, the performance did not improve further, since the number of arcs increases exponentially (see Figure 6). In fact, larger orders showed worse performance metrics when testing the model on data not seen during training. Secondly, when searching for the best predictive variables, it was found that the R R marker alone outperformed any single morphology-based ECG marker, especially in terms of sensitivity values. However, paired combinations including the R R marker seemed to significantly reduce the network order compared to those networks using a single marker. This means that adding more diverse information to the network allows one to obtain more sensitive models with reduced past information, supporting the importance of multivariate models in prediction problems. Furthermore, when attempting to combine all ECG markers using PCA to further increase apnea-detection rates, no significant improvement in performance metrics was observed compared to the paired combinations. Similar results were obtained when all ECG markers were combined as input variables, but with more smoothed performance curves when threshold and order were varied. A possible reason for this could be that all morphology-based ECG markers were associated with the depolarization phase. In this regard, adding other markers from the ECG, such as repolarization markers, or from another physiological signal, could improve the overall performance.
Another relevant aspect worth mentioning is the normalization process applied to the time series before training the dynamic Bayesian networks. As demonstrated in the Methods section, this process increased the chances of detecting apnea episodes in some sequences particularly affected by the amplitude drifts observed in the parameters. On the other hand, this normalization could help to obtain a more generalized model suitable for all cases, regardless of animal and stage, instead of a personalized model for each record. An analysis of this particular step has been discussed in [45] for a human study.
Despite the fact that the overall detector performance was approximately 80%, we believe it is good enough considering the limited number of variables used to train the DBNs. Nevertheless, depending on the rat analyzed, the detection performance for the record-specific analysis was in some cases higher than 90%. However, only the first recording (one-third of the data) was used to train the model with respect to each particular rat; it was then applied to the following two sequences of 15 min. Preliminary analysis, not presented here, showed that training DBNs with later sequences could successfully predict apnea episodes from earlier sequences. In fact, for some animals, the performance measures were better with the model trained with a different recording than those obtained with the model trained on the same recording. This fact suggests the importance of selecting the most appropriate segment to train the record-specific models. It could be extrapolated, for example, to the best PSG recording among several, if available for the same OSA patient, as happens in the Apnea ECG database of the PhysioNet [46], leading to a more robust and personalized model for continuous monitoring of patients. In clinical studies conducted in OSA patients where the epoch duration was shorter than 60 s (i.e., 30 s, 40 s), the results obtained by the proposed methods were 70% [47] and 88% [48] of accuracy using 40-s epochs and involving 21 and 6 subjects, respectively. In the case of 30-s epochs, the accuracy obtained was 75% with 100 involved records [49] and 89% with 12 subjects [50]. The methods that achieved the highest scores were actually developed from one or more respiration-based signals instead of ECG channels.
Furthermore, the number of previous heartbeats used in the training phase was significantly small. Here, the largest order used ( L = 15) sometimes does not exceed 3 s of the analyzed time series, due to the fast heart rhythm of the rats. Longer periods could be used if the original time series is resampled at a lower and fixed frequency, so that more information can be considered at once in the DBN. The latter might also reduce the noise observed in the raw probability time series of the network output. However, the end goal for future work should be a simple predictive network that may eventually be suitable for real-time applications in wearable or ambulatory monitoring devices.
Finally, many clinical studies performed in OSA patients have also used only ECG features to detect sleep apnea, even using a single channel [9]. These features typically include temporal and spectral markers derived from heart rate variability (HRV) analysis, as well as a variety of ECG markers derived from both depolarization and repolarization periods. The majority of these studies have used longer epochs, with 1-min epochs being the most common, to train complex multivariate models that typically run in an offline fashion. These longer epochs may contain multiple apneas of short duration and different types, making it difficult to detect and classify individual episodes and automatically estimate the Apnea–Hypopnea Index (AHI). The design of simpler models with a refined time resolution can be of interest for real-time scenarios. In particular, it can benefit their hardware implementation, thus having a great impact on home monitoring systems, wearable devices, and healthcare for OSA patients.

Limitations

Our study had some limitations, especially the small sample size, the fixed apnea duration, and the single event type. Our experimental model simulated only obstructive apnea episodes of 15 s duration, which is not the real case in OSA patients. A future step could be to develop and test similar models by using PSG recordings from OSA patients. In particular, they should include a wide range of disease severity, event duration and type, and occur at nonperiodic intervals. Further work is needed to achieve comparable results in humans, where the detection and diagnosis of OSA is much more difficult than in well-controlled experimental models. This may involve redefining the model and testing it on other different ECG leads, including cardiac repolarization features and EDR signals, to achieve similar performance metrics.

5. Conclusions

The proposed methodology, based on dynamic Bayesian networks, allowed us to obtain record-specific models for apnea segment detection by using short 15-s epochs and a single ECG channel. Although implemented on experimental data from an OSA model, the methodology represents a first step toward the development of a real-time monitoring tool for OSA patients. Successful implementation of this tool on smart health devices could be used to detect and classify abnormal events over several nights instead of a single night, allowing the patient’s disease progression to be followed. In addition, it could also estimate the number of events per hour as a potential surrogate for the AHI.

Author Contributions

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

Funding

This research has received funding from the European Union Horizon 2020 Research and Innovation Programme under the Marie Sklodowska–Curie grant agreement No. 846636. This work was also supported in part by the Universities and Research Secretariat from the Generalitat de Catalunya under Grant GRC 2021 SGR 01390 and by the Agencia Estatal de Investigación from the Spanish Ministry of Science, Innovation and Universities and the European Regional Development Fund, under the Grant PID2021-126455OB-I00.

Institutional Review Board Statement

The animal study protocol was approved by the Institutional Animal Care and Ethics Committee of the Hospital Clínic, Barcelona. All the experiments were performed by following the European Community Directive (2010/63/UE) and Spanish guidelines (RD 53/2013) for the use of experimental animals.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The authors are grateful to Puy Ruiz, from the IBEC, and Marta Torres, Isaac Almendros, Ramon Farré, and Daniel Navajas, from the Universitat de Barcelona, for their help in designing and implementing the animal model.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
OSAObstructive sleep apnea
DBNDynamic Bayesian networks
AHIApnea–Hypopnea Index
HRVHeart rate variability
PCAPrincipal component analysis
RMSERoot mean square error
ROCReceiver operative characteristic
MAPMaximum a posteriori
CPConditional probability
MLEMaximum likelihood estimator
MMHMMax–min hill-climbing
CLGConnditional linear Gaussian
BICBayesian information criterion
DAGDirected acyclic graphs
CPDConditional probability distribution
PGMProbabilistic graphical model
EDRECG-derived respiratory signal
PSGPolysomnography
IHIntermittent hypoxia

References

  1. Berry, R.B.; Brooks, R.; Gamaldo, C.E.; Harding, S.M.; Marcus, C.; Vaughn, B.V. The AASM Manual for the Scoring of Sleep and Associated Events. Rules, Terminology and Technical Specifications; American Academy of Sleep Medicine: Darien, IL, USA, 2012; Volume 176. [Google Scholar]
  2. Shouldice, R.B.; O’Brien, L.M.; O’Brien, C.; de Chazal, P.; Gozal, D.; Heneghan, C. Detection of obstructive sleep apnea in pediatric subjects using surface lead electrocardiogram features. Sleep 2004, 27, 784–792. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Marcos, J.V.; Hornero, R.; Nabney, I.T.; Álvarez, D.; Del Campo, F. Analysis of nocturnal oxygen saturation recordings using kernel entropy to assist in sleep apnea-hypopnea diagnosis. In Proceedings of the 2011 IEEE International Conference of the Engineering in Medicine & Biology Society (EMBC), Boston, MA, USA, 30 August–3 September 2011; pp. 1745–1748. [Google Scholar]
  4. Koley, B.L.; Dey, D. Real-time adaptive apnea and hypopnea event detection methodology for portable sleep apnea monitoring devices. IEEE Trans. Biomed. Eng. 2013, 60, 3354–3363. [Google Scholar] [CrossRef]
  5. Oliver, N.; Flores-Mangas, F. HealthGear: A real-time wearable system for monitoring and analyzing physiological signals. In Proceedings of the International Conference of Wearable and Implantable Body Sensor Networks (BSN), Cambridge, MA, USA, 3–5 April 2006. [Google Scholar]
  6. Bsoul, M.; Minn, H.; Tamil, L. Apnea MedAssist: Real-time sleep apnea monitor using single-lead ECG. IEEE Trans. Inf. Technol. Biomed. 2010, 15, 416–427. [Google Scholar] [CrossRef] [Green Version]
  7. Kelly, J.M.; Strecker, R.E.; Bianchi, M.T. Recent developments in home sleep-monitoring devices. Int. Sch. Res. Not. 2012, 2012, 768794. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Surrel, G.; Aminifar, A.; Rincón, F.; Murali, S.; Atienza, D. Online obstructive sleep apnea detection on medical wearable sensors. IEEE Trans. Biomed. Circuits Syst. 2018, 12, 762–773. [Google Scholar] [CrossRef] [PubMed]
  9. Mendoça, F.; Mostafa, S.S.; Ravelo-García, A.G.; Morgado-Dias, F.; Penzel, T. A review of obstructive sleep apnea detection approaches. IEEE J. Biomed. Health Inf. 2018, 23, 825–837. [Google Scholar] [CrossRef]
  10. Álvarez, D.; Hornero, R.; Abásolo, D.; Del Campo, F.; Zamarrán, C. Nonlinear characteristics of blood oxygen saturation from nocturnal oximetry for obstructive sleep apnoea detection. Physiol. Meas. 2006, 27, 399. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Marcos, J.V.; Hornero, R.; Álvarez, D.; Del Campo, F.; Aboy, M. Automated detection of obstructive sleep apnoea syndrome from oxygen saturation recordings using linear discriminant analysis. Med. Biol. Eng. Comput. 2010, 48, 895–902. [Google Scholar] [CrossRef]
  12. Morillo, D.S.; Gross, N. Probabilistic neural network approach for the detection of SAHS from overnight pulse oximetry. Med. Biol. Eng. Comput. 2013, 51, 305–315. [Google Scholar] [CrossRef]
  13. Caseiro, P.; Fonseca-Pinto, R.; Andrade, A. Screening of obstructive sleep apnea using Hilbert-Huang decomposition of oronasal airway pressure recordings. Med. Eng. Phys. 2010, 32, 561–568. [Google Scholar] [CrossRef]
  14. Gutiérrez-Tobal, G.C.; Hornero, R.; Álvarez, D.; Marcos, J.V.; Del Campo, F. Linear and nonlinear analysis of airflow recordings to help in sleep apnoea-hypopnoea syndrome diagnosis. Physiol. Meas. 2012, 33, 1261. [Google Scholar] [CrossRef] [PubMed]
  15. Avci, C.; Akbaş, A. Sleep apnea classification based on respiration signals by using ensemble methods. Biomed. Mater. Eng. 2015, 26, S1703–S1710. [Google Scholar] [PubMed] [Green Version]
  16. Karunajeewa, A.S.; Abeyratne, U.R.; Hukins, C. Multi-feature snore sound analysis in obstructive sleep apnea–hypopnea syndrome. Physiol. Meas. 2010, 32, 83. [Google Scholar] [CrossRef] [PubMed]
  17. Goldshtein, E.; Tarasiuk, A.; Zigel, Y. Automatic detection of obstructive sleep apnea using speech signals. IEEE Trans. Biomed. Eng. 2010, 58, 1373–1382. [Google Scholar] [CrossRef] [PubMed]
  18. Penzel, T.; Sabil, A. The use of tracheal sounds for the diagnosis of sleep apnoea. Breathe 2017, 13, e37–e45. [Google Scholar] [CrossRef]
  19. Rachim, V.P.; Li, G.; Chung, W.Y. Sleep apnea classification using ECG-signal wavelet-PCA features. Biomed. Mater. Eng. 2014, 24, 2875–2882. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Hassan, A.R.; Haque, M.A. An expert system for automated identification of obstructive sleep apnea from single-lead ECG using random under sampling boosting. Neurocomputing 2017, 235, 122–130. [Google Scholar] [CrossRef]
  21. Lin, R.; Lee, R.G.; Tseng, C.L.; Zhou, H.K.; Chao, C.F.; Jiang, J.A. A new approach for identifying sleep apnea syndrome using wavelet transform and neural networks. Biomed. Eng. Appl. Basis Commun. 2006, 18, 138–143. [Google Scholar] [CrossRef] [Green Version]
  22. Smruthy, A.; Suchetha, M. Real-time classification of healthy and apnea subjects using ECG signals with variational mode decomposition. IEEE Sens. J. 2017, 17, 3092–3099. [Google Scholar] [CrossRef]
  23. Hassan, A.R. Automatic screening of obstructive sleep apnea from single-lead electrocardiogram. In Proceedings of the 2015 International Conference on Electrical Engineering and Information Communication Technology (ICEEICT), Savar, Bangladesh, 21–23 May 2015; pp. 1–6. [Google Scholar]
  24. Ravelo-García, A.G.; Saavedra-Santana, P.; Juliá-Serdá, G.; Navarro-Mesa, J.L.; Navarro-Esteva, J.; Álvarez-López, X.; Wessel, N. Symbolic dynamics marker of heart rate variability combined with clinical variables enhance obstructive sleep apnea screening. Chaos Interdiscip. J. Nonlinear Sci. 2014, 24, 024404. [Google Scholar] [CrossRef]
  25. Pathinarupothi, R.K.; Vinaykumar, R.; Rangan, E.; Gopalakrishnan, E.; Soman, K.P. Instantaneous heart rate as a robust feature for sleep apnea severity detection using deep learning. In Proceedings of the IEEE-EMBS International Conference on Biomedical & Health Informatics (BHI), Orlando, FL, USA, 16–19 February 2017; pp. 293–296. [Google Scholar]
  26. Penzel, T.; Kantelhardt, J.W.; Bartsch, R.P.; Riedl, M.; Kraemer, J.F.; Wessel, N.; Garcia, C.; Glos, M.; Fietze, I.; Schöbel, C. Modulations of heart rate, ECG, and cardio-respiratory coupling observed in polysomnography. Front. Physiol. 2016, 7, 460. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. de Chazal, P.; Penzel, T.; Heneghan, C. Automated detection of obstructive sleep apnoea at different time scales using the electrocardiogram. Physiol. Meas. 2004, 25, 967. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Mendez, M.O.; Bianchi, A.M.; Matteucci, M.; Cerutti, S.; Penzel, T. Sleep apnea screening by autoregressive models from a single ECG lead. IEEE Trans. Biomed. Eng. 2009, 56, 2838–2850. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Varon, C.; Morales, J.; Lázaro, J.; Orini, M.; Deviaene, M.; Kontaxis, S.; Testelmans, D.; Buyse, B.; Borzée, P.; Sörnmo, L.; et al. A Comparative Study of ECG-derived Respiration in Ambulatory Monitoring using the Single-lead ECG. Sci. Rep. 2020, 10, 5704. [Google Scholar] [CrossRef] [Green Version]
  30. Ullah, H.; Heyat, M.B.B.; Akhtar, F.; Muaad, A.Y.; Ukwuoma, C.C.; Bilal, M.; Lai, D. An Automatic Premature Ventricular Contraction Recognition System Based on Imbalanced Dataset and Pre-Trained Residual Network Using Transfer Learning on ECG Signal. Diagnostics 2022, 13, 87. [Google Scholar] [CrossRef]
  31. Romero, R.; Jané, R. Detecting Obstructive Apnea Episodes Using Dynamic Bayesian Networks and ECG-Based Time-Series. In Proceedings of the 2022 IEEE International Conference on Engineering in Medicine & Biology Society (EMBC), Glasgow, UK, 11–14 July 2022. [Google Scholar]
  32. Romero, D.; Jané, R. Global and Transient Effects of Intermittent Hypoxia on Heart Rate Variability Markers: Evaluation Using an Obstructive Sleep Apnea Model. IEEE Access 2021, 9, 19043–19052. [Google Scholar] [CrossRef]
  33. Camm, A.J.; Malik, M.; Bigger, J.T.; Breithardt, G.; Cerutti, S.; Cohen, R.J.; Coumel, P.; Fallen, E.L.; Kennedy, H.L.; Kleiger, R.E.; et al. Heart rate variability: Standards of measurement, physiological interpretation and clinical use. Task Force of the European Society of Cardiology and the North American Society of Pacing and Electrophysiology. Circulation 1996, 93, 1043–1065. [Google Scholar]
  34. Jané, R.; Laguna, P.; Thakor, N.V.; Caminal, P. Adaptive baseline wander removal in the ECG: Comparative analysis with cubic spline technique. In Proceedings of the Computers in Cardiology (CinC), Durham, NC, USA, 11–14 October 1992; p. 143. [Google Scholar]
  35. Martínez, J.P.; Almeida, R.; Olmos, S.; Rocha, A.P.; Laguna, P. A wavelet-based ECG delineator: Evaluation on standard databases. IEEE Trans. Biomed. Eng. 2004, 51, 570–581. [Google Scholar] [CrossRef]
  36. Romero, D.; Ringborn, M.; Laguna, P.; Pueyo, E. Detection and quantification of acute myocardial ischemia by morphologic evaluation of QRS changes by an angle-based method. J. Electrocardiol. 2013, 46, 204–214. [Google Scholar] [CrossRef]
  37. Jensen, F.V. An Introduction to Bayesian Networks; UCL Press: London, UK, 1996; pp. 1–178. [Google Scholar]
  38. Pearl, J.; Russell, S. Bayesian Networks; The Handbook of Brain Theory and Neural Networks; MIT Press: Cambridge, MA, USA, 2003; pp. 157–160. [Google Scholar]
  39. Lauritzen, S.L.; Wermuth, N. Graphical models for associations between variables, some of which are qualitative and some quantitative. Ann. Stat. 1989, 17, 31–57. [Google Scholar] [CrossRef]
  40. Koller, D.; Friedman, N. Probabilistic Graphical Models: Principles and Techniques; MIT Press: Cambridge, MA, USA, 2009. [Google Scholar]
  41. Tsamardinos, I.; Brown, L.E.; Aliferis, C.F. The max-min hill-climbing Bayesian network structure learning algorithm. Mach. Learn. 2006, 65, 31–78. [Google Scholar] [CrossRef] [Green Version]
  42. Trabelsi, G.; Leray, P.; Ben Ayed, M.; Alimi, A.M. Dynamic MMHC: A local search algorithm for dynamic Bayesian network structure learning. In Proceedings of the International Symposium on Intelligent Data Analysis (IDA), London, UK, 17–19 October 2013; pp. 392–403. [Google Scholar]
  43. Scutari, M.; Denis, J.B. Bayesian Networks: With Examples in R; Chapman and Hall/CRC: New York, NY, USA, 29 July 2021. [Google Scholar]
  44. Murphy, K.P. Machine Learning: A Probabilistic Perspective; MIT Press: Cambridge, MA, USA, 2012. [Google Scholar]
  45. Mrudula, G.B.; Haritha, H.; Kumar, C.S.; Kumar, A.A.; Gopinath, G. Feature normalization for improving the performance of sleep apnea detection system. In Proceedings of the 15th IEEE India Council International Conference (INDICON), Coimbatore, India, 15–18 December 2018; pp. 1–5. [Google Scholar]
  46. Penzel, T.; Moody, G.B.; Mark, R.G.; Goldberger, A.L.; Peter, J.H. The apnea-ECG database. In Proceedings of the Computers in Cardiology (CinC), Cambridge, MA, USA, 24–27 September 2000. [Google Scholar]
  47. Lázaro, J.; Gil, E.; Vergara, J.M.; Laguna, P. Pulse rate variability analysis for discrimination of sleep-apnea-related decreases in the amplitude fluctuations of pulse photoplethysmographic signal in children. IEEE J. Biomed. Health Inf. 2013, 18, 240–246. [Google Scholar] [CrossRef] [PubMed]
  48. Ozdemir, G.; Nasifoglu, H.; Erogul, O. A time-series approach to predict obstructive sleep apnea (OSA) Episodes. In Proceedings of the 2016 2nd World Congress on Electrical Engineering and Computer Systems and Science, Budapest, Hungary, 16–17 August 2016. [Google Scholar]
  49. Haidar, R.; Koprinska, I.; Jeffries, B. Sleep apnea event detection from nasal airflow using convolutional neural networks. In Proceedings of the 24th International Conference of Neural Information Processing (ICONIP 2017), Guangzhou, China, 14–18 November 2017; pp. 819–827. [Google Scholar]
  50. Maali, Y.; Al-Jumaily, A. Automated detecting sleep apnea syndrome: A novel system based on genetic SVM. In Proceedings of the 2011 11th International Conference on Hybrid Intelligent Systems (HIS), Melacca, Malaysia, 5–8 December 2011; pp. 590–594. [Google Scholar]
Figure 1. The R R time series (in blue) before (top) and after (bottom) removing the trends and slow oscillations (in black). Red vertical lines mark the time instant of apneas onset and end.
Figure 1. The R R time series (in blue) before (top) and after (bottom) removing the trends and slow oscillations (in black). Red vertical lines mark the time instant of apneas onset and end.
Sensors 23 03371 g001
Figure 2. Examples of dynamic Bayesian networks obtained using four time slices ( L = 3). (a) Only the R R time series. (b) The combination of R a and R R time series. (c) All the ECG parameters ( R a , S a , R S d , U s , D s ) combined with the R R time series.
Figure 2. Examples of dynamic Bayesian networks obtained using four time slices ( L = 3). (a) Only the R R time series. (b) The combination of R a and R R time series. (c) All the ECG parameters ( R a , S a , R S d , U s , D s ) combined with the R R time series.
Sensors 23 03371 g002
Figure 3. (Top) Normalized R R and Q R S p c a time series (blue color) and apnea occurrence (red color) for a particular segment of four apnea episodes. (Bottom) Apnea probability estimates using a DBN with L = 10, before (gray color) and after smoothing (black color) using only the R R marker, and when adding the Q R S p c a marker.
Figure 3. (Top) Normalized R R and Q R S p c a time series (blue color) and apnea occurrence (red color) for a particular segment of four apnea episodes. (Bottom) Apnea probability estimates using a DBN with L = 10, before (gray color) and after smoothing (black color) using only the R R marker, and when adding the Q R S p c a marker.
Sensors 23 03371 g003
Figure 4. (Top) Original and normalized R R (blue color) and apnea occurrence (red color) for a particular sequence of 15 episodes. (Bottom) Apnea probability estimates (gray, raw values; black, average values) before and after the normalization process.
Figure 4. (Top) Original and normalized R R (blue color) and apnea occurrence (red color) for a particular sequence of 15 episodes. (Bottom) Apnea probability estimates (gray, raw values; black, average values) before and after the normalization process.
Sensors 23 03371 g004
Figure 5. RMSE values as a function of the Markovian order when training DBNs in three different recordings for R R (top) and R a (bottom) markers. Each color represents one particular sequence/recording of 15 min within the same rat having a different number of apnea episodes.
Figure 5. RMSE values as a function of the Markovian order when training DBNs in three different recordings for R R (top) and R a (bottom) markers. Each color represents one particular sequence/recording of 15 min within the same rat having a different number of apnea episodes.
Sensors 23 03371 g005
Figure 6. Complexity of the dynamic Bayesian networks as a function of lag ( L ) in three different recordings for the R R + R a combination. The blue line represents the number of nodes involved while the other colors indicate the total number of arcs obtained from the networks’ structure in each sequence/recording.
Figure 6. Complexity of the dynamic Bayesian networks as a function of lag ( L ) in three different recordings for the R R + R a combination. The blue line represents the number of nodes involved while the other colors indicate the total number of arcs obtained from the networks’ structure in each sequence/recording.
Sensors 23 03371 g006
Figure 7. Performance metrics as a function of threshold obtained for Bayesian networks trained with individual (left  R R , middle  R a ) and combined (right  R R + R a ) markers. Color intensity represents the lag ( L = 1:15), where the lighter tones refer to smaller values.
Figure 7. Performance metrics as a function of threshold obtained for Bayesian networks trained with individual (left  R R , middle  R a ) and combined (right  R R + R a ) markers. Color intensity represents the lag ( L = 1:15), where the lighter tones refer to smaller values.
Sensors 23 03371 g007
Table 1. Overall apnea detection performance obtained for the R R and ECG markers by using L = { 5 , 15 } and threshold equal 0.2.
Table 1. Overall apnea detection performance obtained for the R R and ECG markers by using L = { 5 , 15 } and threshold equal 0.2.
L = 5 L = 15
Se Sp Acc Se Sp Acc
Marker(%)(%)(%)(%)(%)(%)
R R 63.384.884.091.176.779.8
R a 35.280.277.166.779.579.8
S a 30.176.773.664.477.477.6
R S d 32.670.268.951.174.873.9
U s 29.675.872.765.279.079.1
D s 31.168.467.350.474.073.2
Q R S P C A 40.080.177.663.779.279.0
Table 2. Overall apnea-detection performance obtained for R R and all paired combinations by using L = { 5 , 15 } .
Table 2. Overall apnea-detection performance obtained for R R and all paired combinations by using L = { 5 , 15 } .
L = 5 L = 15
Se Sp Acc Se Sp Acc
Marker(s)(%)(%)(%)(%)(%)(%)
R R
   Training63.384.884.091.176.779.8
   Testing80.668.670.567.869.669.2
R R + R a
   Training81.179.581.390.478.480.6
   Testing58.374.374.060.069.170.2
R R + S a
   Training81.980.181.885.677.279.1
   Testing64.469.770.755.074.573.7
R R + R S d
   Training83.779.881.686.376.077.8
   Testing64.479.779.362.064.866.7
R R + U s
   Training83.080.282.189.377.780.3
   Testing70.669.071.263.371.873.0
R R + D s
   Training79.380.181.685.276.578.1
   Testing68.981.581.356.775.174.5
R R + P C A
   Training77.480.081.185.678.280.2
    Testing57.270.270.758.367.268.4
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

Romero, D.; Jané, R. Dynamic Bayesian Model for Detecting Obstructive Respiratory Events by Using an Experimental Model. Sensors 2023, 23, 3371. https://doi.org/10.3390/s23073371

AMA Style

Romero D, Jané R. Dynamic Bayesian Model for Detecting Obstructive Respiratory Events by Using an Experimental Model. Sensors. 2023; 23(7):3371. https://doi.org/10.3390/s23073371

Chicago/Turabian Style

Romero, Daniel, and Raimon Jané. 2023. "Dynamic Bayesian Model for Detecting Obstructive Respiratory Events by Using an Experimental Model" Sensors 23, no. 7: 3371. https://doi.org/10.3390/s23073371

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