Abstract

Continuous glucose monitoring (CGM) by suitable portable sensors plays a central role in the treatment of diabetes, a disease currently affecting more than 350 million people worldwide. Noninvasive CGM (NI-CGM), in particular, is appealing for reasons related to patient comfort (no needles are used) but challenging. NI-CGM prototypes exploiting multisensor approaches have been recently proposed to deal with physiological and environmental disturbances. In these prototypes, signals measured noninvasively (e.g., skin impedance, temperature, optical skin properties, etc.) are combined through a static multivariate linear model for estimating glucose levels. In this work, by exploiting a dataset of 45 experimental sessions acquired in diabetic subjects, we show that regularisation-based techniques for the identification of the model, such as the least absolute shrinkage and selection operator (better known as LASSO), Ridge regression, and Elastic-Net regression, improve the accuracy of glucose estimates with respect to techniques, such as partial least squares regression, previously used in the literature. More specifically, the Elastic-Net model (i.e., the model identified using a combination of and norms) has the best results, according to the metrics widely accepted in the diabetes community. This model represents an important incremental step toward the development of NI-CGM devices effectively usable by patients.

1. Introduction

Diabetes consists of a malfunction of the glucose-insulin regulatory system leading to the onset of long and short term complications, like retinopathy, neuropathy, and cardiovascular and hearth diseases, due to sustained blood glycaemic levels exceeding the normal range of 70–180 mg/dL [1]. According to the International Diabetes Federation, diabetes is estimated to currently affect more than 350 million people worldwide, and this number is rapidly growing [2]. Not surprisingly, in the last few decades, diabetes has received an increasing attention both for its social and economic implications [3].

Standard therapy of type 1 diabetes is based on a combination of diet, physical activity, and exogenous insulin injections, modulated on the basis of individual patient levels of glucose concentration in the blood. Accurate and frequent monitoring of glucose concentrations by portable sensor devices plays a crucial role in the diabetes therapy [4]. Self-monitoring blood glucose (SMBG) sensors require the collection of a blood sample by pricking the skin with a lancet device. An external pocket device is then used to analyze the blood and determine glucose concentration for instance by the glucose oxidase principle [5]. These sensors are uncomfortable for the patient and are typically used no more than 3-4 times per day. Such sparseness of sampling does not allow the observation of glucose dynamics and glucose excursions outside the safety range, and dangerous nocturnal hypoglycaemic swings are often not detected. To overcome these problems, portable continuous glucose monitoring (CGM) sensors, measuring blood glucose values every 1 to 5 minutes for up to 7 consecutive days, have been proposed in the early 2000s and are now of great interest for the tuning and optimization of diabetes therapy [6, 7]. In particular, online applications of CGM sensors include prevention of hyper-/hypoglycaemic events; see for example [8, 9], and closed-loop glucose control aimed at determining optimal automatic insulin infusion in the so-called artificial pancreas systems; see [1014]. Notably, dealing with these online applications requires facing nontrivial signal processing issues connected to denoising [15], calibration [16], and prediction [1721] (see [22, 23] for reviews) calling for the development of the so-called “smart” CGM sensor architecture [24].

Most of the CGM sensors requiring the placement of a small needle in the subcutaneous tissue use an enzyme-based glucose-oxidase electrode and thus are invasive, albeit minimally. To limit patients’ discomfort, in the last decade, several noninvasive continuous glucose monitoring (NI-CGM) sensors have been prototyped. Many different physical principles have been considered to pursue NI-CGM, but none of them has clearly outperformed the others so far; see for example [2527]. One major difficulty with NI-CGM is the fact that environmental (e.g., temperature) and physiological (e.g., sweat, blood oxygenation, etc.) processes act as perturbing factors and often allow blood glucose changes to be tracked only in highly controlled conditions (e.g., during in-hospital studies) [2830]. To tackle this issue, an approach gaining increasing attention in the last years is the multisensor approach to NI-CGM; that is, instead of focusing on a single physical principle, these devices resort to a combination of technologies. For instance, the GlucoTrack [31] exploits a mix of thermal, acoustic, and electromagnetic technologies and compares the three measurements, assuming they all reflect a glucose-related measurement. A different, yet recent, multi-sensor prototype [32] employs a combination of dielectric spectroscopy (DS) and mid-infrared-based sensors. A further example of multi-sensor device, proposed by Solianis Monitoring AG (Zürich, Switzerland, technology taken over by Biovotion AG, Zurich, Switzerland), embeds sensors of different nature for the biophysical characterisation of skin and underlying tissue in order to track glucose-related and perturbing factors separately [30, 33].

Multisensor approaches to NI-CGM require a model to connect the physical quantities measured by the sensors with blood glucose concentrations. For instance, in the Solianis Monitoring AG prototype (from now on named Solianis for the sake of readability) considered in this work, a multivariate linear regression model is used to combine ~150 signals recorded noninvasively for estimating a glucose concentration profile (see Section 3.1 for more details). As described in [33, 34], the linear regression model is identified on a dataset collected in a population of subjects and comprising multi-sensor channels measurements and reference blood glucose (RBG) values collected in parallel by a gold standard technique. Previous work [35] has shown that a regularised identification method, based on norm (least absolute shrinkage and selection operator—LASSO), provided a glucose profile more accurate than that of other models identified with approaches controlling complexity such as subset selection method or partial least squares (PLS). In the present work, by using the same dataset of 45 experimental sessions used in [35], we demonstrate that the accuracy of glucose estimates can be further improved by considering norm regularization (Ridge regression) and a combination of and norms (Elastic-Net—EN regression), providing a further incremental step toward the development of an NI-CGM effectively usable by diabetic patients.

2. Database

The database consists of 45 experimental sessions recorded from 6 type 1 diabetic subjects, during which plasma glucose was induced to vary according to different predetermined profiles to cause different hypo- and hyperglycaemic excursions. During each session, multi-sensor data and RBG were acquired in parallel, with a time sampling of 20 seconds and 10–15 minutes, respectively. The RBG samples were acquired by means of a SMBG device. The study was performed at the University Hospital Zurich according to the requirements of good clinical practice and was approved by the local ethical commission. More clinically related information can be found in [33].

For the analysis, the database was split into two data subsets of 22 and 23 experimental sessions, respectively. If the first data subset is used for identifying the models with the different techniques, the second is used to test the models over “unseen” multi-sensor data (1–>2) and vice versa (2–>1). For the sake of space, we will not discuss results of the application of the model on the same dataset used for their identification, and only model test results will be considered.

Multi-sensor data in the identification data subset undergoes a preprocessing step described in detail in [35]. Briefly, each multi-sensor channel is normalized to have zero mean and unitary standard deviation. These values are then used to standardize the same channels in the test data subset to permit simulation of real-time glucose monitoring. Moreover, the first RBG value available at the beginning of each recording session is used to calibrate the estimated glucose profiles by the model setting a baseline adjustment, to allow for a real-time consideration/implementation.

3. Methods

3.1. Glucose Determination by a Multisensor System

In stating the problem, we deliberately make only a brief description of the framework we are working on and we refer the reader to the quoted bibliography for details.

Rather than focusing on a single physical principle, the Solianis multi-sensor device resorts to a combination of technologies, embedded into the substrate in contact with the skin for the biophysical characterisation of the skin and underlying tissue in order to account for confounding factors, which can significantly deteriorate the accuracy of glucose measurements [30, 3638]. In particular, glucose-related signals are obtained from DS fringing field capacitive electrodes, with different geometrical properties, providing a spectrum of the frequency-dependent complex dielectric properties of skin and underlying tissue including blood, which can be easily parameterised by its magnitude and phase [39]. Environmental and physiological processes that can interfere with the measurements of the main glucose-related signals are measured with temperature, optical, humidity, accelerometer, and additional DS sensors incorporated into the same device substrate [30]. About 150 channels are thus provided by the multi-sensor device (Figure 1, left), which have to be properly combined through a mathematical model (Figure 1, middle) for estimating glucose concentrations in the blood (Figure 1, right). Since a mechanistic description comprising the physical principles, linking measured channels with physical/physiological processes, in particular, related to glucose changes, has not yet been fully developed, a “black-box” multivariate linear regression model is used.

Formally, if is the number of data points available and is the number of measured channels, the model to identify from the identification dataset is described as where represents the () target vector containing RBG values, is the () matrix collecting the multi-sensor channels, is the () vector containing the coefficients of the linear model, and is the () term representing the error.

The vector in (1) can be identified by minimizing the cost function given by the residual sum of squares (RSS): measuring the distance between data and model predictions. Since this cost function has a quadratic form, a closed-form solution, the so-called ordinary least squares (OLS) estimate, can be obtained. In the case under consideration, OLS suffers from overfitting due to the high dimensionality of the measurement space and to the correlation between subsets of input channels (well visible in the channels showed in Figure 1, left). Recent work [34] showed that further improved results can be obtained by exploiting subset selection techniques. Then, further work [35] investigated the use of other methods controlling “model complexity,” including PLS (widely used in chemometrics and related fields dealing with spectroscopy data) and the LASSO regularization technique. It has been shown that regularization techniques, in particular, the LASSO, outperform PLS since it sets many coefficients to zero being less sensitive to occasional noise in the multi-sensor channels.

Formally speaking, regularized model identification techniques add a term to the cost function of (2), leading to where is a function of reflecting complexity. Depending on the form and on the parameters of in (2), the resulting model will present different well-known features, as will be briefly reviewed in the following section.

Thus, the minimizing (3) is identified establishing a trade off between adherence to the data and model complexity, usually by a cross-validation procedure as better discussed next.

3.2. Model Identification by Regularisation Techniques
3.2.1. Norm: LASSO

In the term of (3), the norm can be used. In the literature, this norm has been proposed in signal processing (under the name of basis pursuit) [40] and in statistics [41] for its main feature of inducing sparse solutions. Formally, in (3), the norm leads to define as the sum of the absolute values of the coefficients of the model multiplied by the scalar nonnegative parameter, hereafter referred to as regularization parameter for the sake of reasoning. Thus, the solution is found by minimizing and is known as the least absolute shrinkage and selection operator (LASSO), for its properties of shrinking many coefficients to zero selecting only few input variables. In particular, in our application documented in [35], the LASSO was shown to outperform PLS since it sets many coefficients to zero being less sensitive to occasional noise in multi-sensor channels. Equation (5) does not have a closed-form solution because the cost function is not differentiable when some coefficients are zero, and a plethora of methods have been developed to calculate approximate solutions numerically; see [42, 43] for reviews. In particular, an efficient technique for computing the LASSO solution is obtained by modifying the least angle regression algorithm [44]. With this technique, the parameter that has to be fixed represents the number of input variables allowed to enter the model, indicated with in Section 3.

Remark 1. The regularization parameter weighting the term in (5) is obtained by means of a standard procedure known as -fold cross-validation [45]. Briefly, the identification dataset is split into approximately equal parts. Then, the model is identified on parts and tested over the portion of data not considered for deriving the model, calculating the mean squared error (MSE): where and represent the true and estimated output, respectively, and represents the number of data points in the portion of data. This procedure is repeated times and for a range of values of the parameter that has to be determined. Then, the cross-validation curve is plotted, presenting the MSE as a function of the regularization parameter. Empirical evidence suggests choosing its value in correspondence with a clear drop of the cross-validation curve.

3.2.2. Norm: Ridge Regression

The norm involves the penalization of the sum of squares of the coefficients of the model multiplied by a scalar nonnegative parameter: The so-called Ridge regression solution, from now on ridge, is thus given by The quadratic nature of the cost function in (7) entails a closed-form solution for dependent on the parameter : Also, in this case, the regularization parameter can be fixed by means of cross-validation. According to [45], as an indication of the model complexity, the degrees of freedom (df) can be used: where the are the singular values of . Thus, to determine the regularization parameter by cross-validation, the MSE is plotted against df in (10).

3.2.3. Norms: Elastic Net-Regression

The so-called Elastic-Net regression, from now on EN, resorts to a weighted sum of the two previously described norms, defining as where weighs the contribution of the two norms while sets the amount of regularization [46]. Hence, the EN model parameters are obtained solving the following: Problem (12) does not have a closed-form solution, and several numerical algorithms have been proposed to compute an approximate solution (some of them are simple adaptations of those developed for the LASSO [46]). The algorithm that has been used in this work is the one based on cyclical coordinate descent originally developed for the LASSO [47] and adapted to problem (12) following [4850]. The parameters and are determined by cross-validation, following a procedure similar to that of Remark 1.

3.3. Criteria for Model Assessment

The accuracy of estimated glucose profiles in the model test is measured through a set of indexes widely used in the diabetes research community. In particular, we consider the root mean squared error (RMSE) the mean absolute difference (MAD), indicating how much estimated glucose values are lower or higher than the reference, and the mean absolute relative difference (MARD), which characterizes the relative errors (in %) of the estimated glucose: where and , for , are, respectively, the reference RBG samples and the glucose estimates provided by the models (all the experimental sessions are simultaneously considered). Finally, a popular method used in the diabetes community to judge on the point accuracy of glucose sensors is the error grid analysis (EGA) proposed by Clarke and coworkers [51]. The area where estimated glucose by the model and RBG values are displayed as a scatterplot is broken down into five regions (labeled from A to E). Zone A represents those glucose values within 20% of the RBG values and so on. The most dangerous situations are those where estimated glucose values fall into zones C/D/E because, from a clinical point of view, they will lead to unnecessary or even wrong and potentially dangerous treatments. An evolution of EGA developed for CGM sensors is the continuous EGA (CEGA) that also measures the accuracy of estimated glucose trends by creating a grid which is broken down into regions labeled from AR to ER; see [52] for details.

To give an idea of the values of these indexes for the current state-of the-art, minimally invasive (needle-based) commercial CGM sensors, a recent study [11] reported mean MARD levels ranging from 11.8 to 20.2% and a percentage of data points in the region ranging from 98.9 to 96.9 under controlled conditions, comparing CGM measurements to gold standard blood glucose sampling.

4. Results

4.1. Regularization Parameter Determination by Cross-Validation

The first step in the analysis is setting the values for in (5), (8), and (12) and for in (12). These were determined by finding where the cross-validation curve presents a clear drop in slope, as explained in Remark 1. Figure 2 shows the values obtained when identification data subset “part 1” is considered, and the red cross, together with a vertical red dashed line for visualization purpose, highlights the selected “optimal” value (cross-validation plots for identification data subset “part 2” are not showed for the sake of space). Specifically, a -fold cross-validation strategy has been applied, with , for having a good compromise between bias and variance of the estimated error [45]. The left subplot shows the error curve as a function of the number of latent variables for the LASSO model, indicating a drop of the cross-validation curve around 15. The choice of the regularization parameter for ridge followed a similar approach, with the cross-validation curve (middle panel) presenting a drop when the degree of freedom, defined in (10), approximately equals 50, corresponding to . Similarly for EN, the ending part of the drop in the error curve can be noticed for (Figure 2(c)), corresponding to . For EN, different cross-validation curves for different values of were examined. The most reasonable choice seemed to be that obtained with . Indeed, this combination of parameters is the one providing a good trade off between the and norms allowing a reasonable compromise for the EN model to be achieved. A value of can suggest that, although it is important to shrink channel weights to zero in order to lower the probability of occasional jumps or spikes entering the model, allowing a grouping effect over correlated predictors is also important for a more robust estimation of glucose profiles.

4.2. Model Test

Figure 3 shows examples of estimated blood glucose concentration profiles (continuous line) versus reference RBG samples (open bullets) for two representative experimental sessions. Visual inspection suggests that the profiles provided by the EN model (bottom panels) outperform, in terms of accuracy, those provided by LASSO (top) and Ridge (middle). The same observation applies for the examples depicted in Figure 4, which represents a more challenging situation because of the presence of wider disturbances, as witnessed by the spikes and jumps affecting the representative multi-sensor raw channels displayed in an additional fourth row of panels. This qualitative observation is supported from the analysis of the MARD values obtained for the representative sessions in Figure 3, that is, 16.7%, 16.9%, and 16.5% (experimental session depicted in the left panels) and 12.7%, 11.2%, and 9.1% for the LASSO, Ridge, and EN models, respectively. When occasional noise is affecting some of the multi-sensor channels, the MARD values slightly worsen, as can be seen for the experiment depicted in Figure 4(a) presenting MARD of 53% for the LASSO, 24.5% for Ridge, and 20.6 for EN. However, the EN model still provides blood glucose estimation profiles more accurately than Ridge and LASSO. This is confirmed, in general, by Table 1, which shows the aggregate results over the test data subset.

By analysing the results in more detail, the LASSO model seems to have a slight advantage in estimating glucose trends (last column of Table 1). The reason is twofold: first, the regularization performed by the norm prevents the model coefficients from assuming large values thus predicting glucose profiles that are more flat than the other models, as it happens for example in Figures 3(b) and 4(b); second, channels more sensitive to noise that contain also glucose-related information are considered by Ridge and EN exploiting the effect of the norm but are less probable to be selected by LASSO, thus yielding smoother estimates. This fact can clearly be seen from Figure 4, where artifacts are present (e.g., in channel no.2) for the session of left data and in channel no.3 for the session of the right data.

Interestingly, the LASSO model seems more robust than the other models to these jumps in the data, preserving the glucose profile with elevated smoothness and reasonably accurate trend. Indeed the norm shrinks many coefficients to zero allowing an easier interpretation of the results with a reduced number of original variables, representing the strongest effects, considered important for estimating glucose profiles. This is a typical feature of the LASSO to act as a variable selection method. If, from one side, smoother estimates of glucose profiles are obtained with the shrinking properties of the LASSO, sometimes this can lead to biased estimates (see Figure 4(a)).

The Ridge model is identified minimizing the RSS cost function subject to a bound on the norm of the coefficients. This norm does not have the ability of inducing sparseness on the coefficients of the multivariate linear regression model; thus a parsimonious model is not identified and all the predictors are kept in the model. This might cause the estimated glucose profiles by the Ridge model to be sensitive to occasional spikes or jumps in the multi-sensor channels, as can be seen in Figure 4(b), where the Ridge model is the one more sensitive among the three. However, estimated glucose profiles by the Ridge model show accuracy indicators slightly better than those of LASSO. This might indicate that (a) channels discharged by the norm because sensitive to occasional spikes or jumps actually contain useful glucose-related information and (b) that retaining information from all the input channels may help in compensating noisy channels thanks to the grouping effect induced by the norm. Thus, it is reasonable to expect that a combination of the and norms could identify a model sharing both properties of sparseness and grouping effect.

Indeed, as mentioned before, the EN model results outperform those of Ridge and LASSO allowing a reasonably accurate estimation of the glucose profile concentrations also when occasional noise is affecting some multi-sensor channels (see Figure 4), presenting lower MARD values than LASSO and Ridge. Thus, EN is the model presenting the best indicators and is only slightly worse than LASSO in accuracy for glucose trends (see CEGA results). Moreover, its clinical accuracy in terms of Clarke error grid, with a percentage of points within the of 92.3 (see Table 1), is substantially close to that of minimally invasive devices, spanning from 98.9 to 96.9 [53].

The good results obtained by the EN model are likely due to the combination of the and norms, giving to this model both the advantages of LASSO and Ridge. Indeed, a limitation of the LASSO is that, if there is a group of correlated variables, then it tends to select only one variable from the group and does not care which one is selected, thus lacking the ability of revealing grouping information. On the opposite, the norm allows all coefficients to enter the model, resulting in more sensitive to noisy channels. Thus, the norm shrinks channel weights to zero (eliminating multi-sensor channels not useful for predicting glucose), while the norm encourages a grouping effect (automatically including whole groups into the model once one channel among them is selected). This combination results in indicators outperforming those of the other models and in estimated glucose profiles with a good trade off between sparseness of the model coefficients and robustness due to the grouping effect (see, e.g., Figure 4(a)).

Model test results when data subset “part 2” is used for model identification and data subset “part 1” for model test are comparable with those in Table 1 (not shown here for the sake of space).

5. Conclusions

In diabetes management, tight monitoring of glycaemic levels by CGM sensors is important for avoiding both long and short term complications related to hyper-and hypoglycaemic excursions. NI-CGM devices are potentially more appealing than the minimally invasive sensors based on needle electrodes, but their development is challenging for several technological and methodological reasons. In the last years, the idea of embedding sensors of different nature within the same device in order to obtain a better biophysical characterisation of the skin and underlying tissues has gained particular attention to develop NI-CGM. In these multi-sensor approaches, a model linking the measured multi-sensor channels to glucose is needed, together with a set of techniques that can be used to identify its parameters. In this work, we investigated the use of regularisation-based methods to identify the linear regression model employed in the multi-sensor device for NI-CGM proposed in [30]. Results on 45 experimental sessions indicate that the EN model generally outperforms the other models: thanks to the combination of the and norms, it allows to take the advantage of the LASSO—shrinking many model weights to zero being more robust to possible occasional jumps or spikes occurring on the multi-sensor data—and of the Ridge model—averaging the contribution of correlated channels allowing a more robust estimation of glucose profiles. With respect to the previous sensor literature, where PLS represents the current state of the art (see [34, 54, 55] to mention just a few), we showed that EN can become very useful with multi-sensor data. While retaining information from a group of variables (as PLS does), EN also automatically selects those channels representing the strongest effects, giving more insight into the specific problem at hand.

To conclude, in this work, we showed that further increased point accuracy can be obtained through suitable techniques for the identification of the multivariate model, representing an important incremental step towards the development of NI-CGM devices. While most of the accuracy indices of Table 1 have not yet reached a fully comparable level with those of current enzyme-based needle sensors [53], glucose trends estimated by the considered NI-CGM device exhibit an acceptable accuracy (last column of Table 1). This result could be potentially important in the treatment of diabetes since the glucose trend can be valid adjunctive information to complement standard SMBG devices that measure glucose by fingerstick, for example, helping the diabetic patient in preventing the occurrence of critical events, such as hypoglycaemia, by exploiting the dynamic risk concept recently developed in [56].