• No results found

University of Groningen Ambulatory assessment of human circadian phase and related sleep disorders from heart rate variability and other non-invasive physiological measurements Gil Ponce, Enrique

N/A
N/A
Protected

Academic year: 2021

Share "University of Groningen Ambulatory assessment of human circadian phase and related sleep disorders from heart rate variability and other non-invasive physiological measurements Gil Ponce, Enrique"

Copied!
23
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

University of Groningen

Ambulatory assessment of human circadian phase and related sleep disorders from heart

rate variability and other non-invasive physiological measurements

Gil Ponce, Enrique

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from it. Please check the document version below.

Document Version

Publisher's PDF, also known as Version of record

Publication date: 2017

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

Gil Ponce, E. (2017). Ambulatory assessment of human circadian phase and related sleep disorders from heart rate variability and other non-invasive physiological measurements. University of Groningen.

Copyright

Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

Take-down policy

If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum.

(2)

39

2

H

UMAN CIRCADIAN PHASE ESTIMATION FROM

SIGNALS COLLECTED IN AMBULATORY

CONDITIONS USING AN AUTOREGRESSIVE MODEL

Gil EA, Aubert XL, Møst EIS, Beersma DGM. (2013), Human circadian

phase estimation from signals collected in ambulatory conditions

using an autoregressive model. Journal of Biological Rhythms 2013,

(3)

40

2.1 A

BSTRACT

Phase estimation of the human circadian rhythm is a topic that has been explored using various modeling approaches. The current models range from physiological to mathematical, all attempting to estimate the circadian phase from different physiological or behavioral signals. Here we have focused on the estimation of circadian phase from unobtrusively collected signals in ambulatory conditions using a statistically trained autoregressive moving average with exogenous inputs (ARMAX) model. Special attention has been given to the evaluation of heart rate inter-beat intervals (RR intervals) as a potential circadian phase predictor. Prediction models were trained using all possible combinations of RR intervals, activity levels, and light exposure, each collected over a period of 24 hours. The signals were measured without any behavioral constraints, aside from the collection of saliva in the evening to determine melatonin concentration which was measured in dim light conditions. The model was trained and evaluated using two completely independent datasets, with 11 and 19 participants respectively. The output was compared to the gold standard of circadian phase: dim light melatonin onset (DLMO). The most accurate model we found made use of RR intervals and light, and was able to yield phase estimates with a prediction error of 2±39 minutes (mean±SD) from the DLMO reference value.

2.2 I

NTRODUCTION

Human beings possess an internal master biological clock, located in the suprachiasmatic nuclei (SCN) region of the hypothalamus, which is responsible for the synchronization of all other secondary clocks in the body. The clock is advanced or delayed through environmental cues known as zeitgebers. Although many zeitgebers exist, by far the most influential one is light [1–3]. Through the exposure to different light levels during the night and day, a person becomes entrained to the external light/dark cycle [2,4,5]. The master clock relays this timing information to the secondary clocks to synchronize the various organs and influence physiological and behavioral signals such as core body and skin temperature, heart rate, hormone levels, and the sleep/wake cycle [6–12].

Due to its location in the brain, it is very difficult to assess the state of the circadian clock directly. Therefore, one must rely on signals closely coupled to the master circadian clock [13]. Due to the complexity of these signals and the intricate interaction between endogenous and exogenous factors, these signals are often subject to noise and masking effects which make the assessment of the circadian phase a challenging task. One of the most widely accepted measures of circadian phase is the dim light melatonin onset (DLMO) [5,14]. This requires the collection of saliva, blood or urine samples either during the evening or over an entire night. From this melatonin profile, the DLMO can be calculated using one of several described methods [14]. Controlling for masking effects requires monitoring the participants’ light exposure, activity level, and food intake. Another circadian phase marker is

(4)

41

2

core body temperature (CBT) [13], which requires the ingestion of a temperature

recording pill or the use of a rectal thermistor. This method is invasive and requires that the participants adhere to behavioral constraints to minimize masking effects [15,16]. The procedures are time consuming, expensive, and inconvenient to the participants.

In an attempt to solve this problem, models have been derived which try to minimize the burden on participants by using non-invasively collected signals. Some commonly used signals are activity levels, light exposure, [17–20] and skin temperature [21,22]. The drawback with these methods is that they require the collection of data over extended periods of time, in some cases up to one week, to obtain accurate results. Given a model that uses unobtrusively collected signals in ambulatory conditions measured over a short period of time, circadian phase estimation could be done inexpensively with little to no burden to the participant. This would lead to a more efficient and practical way of assessing the circadian phase of a person, and in turn, improve the efficiency of the diagnosis of circadian rhythm irregularities. The method could furthermore help in monitoring the effects of light or hormone therapy [4,23,24], or improve the timing of chronomedication [25,26].

A signal that follows a circadian cycle, yet is not used as a circadian phase marker, is heart rate variability (HRV). Heart rate variability has become a common measure of cardiovascular regulation by the autonomic nervous system (ANS) [27–31]. The balance between the sympathetic and parasympathetic nervous systems changes throughout the day and this is expressed in the HRV as a modulation in the length of RR intervals, as well as in the spectral power at frequencies associated to sympathetic and parasympathetic activation. Furthermore, the modulation of HR can be attributed to endogenous circadian factors or to exogenous factors such as physical exercise, stress and sleep. Several studies have aimed at separating the effects of sleep on HRV to isolate the circadian influence through sleep deprivation [32], forced desynchrony [33], or ultradian rhythms [34]. Physiologically, the circadian modulation of RR intervals originates at the SCN which connects to the paraventricular nucleus (PVN) of the hypothalamus via separate pre-sympathetic and pre-parasympathetic neurons [35]. The PVN is the region of the hypothalamus responsible for hormonal and autonomic control. Furthermore, the PVN is connected to the dorsal motor vagal nucleus which is involved in the parasympathetic activity of the heart and to the preganglionic sympathetic neurons in the intermediolateral cell column, which together regulate cardiac activity [34]. Using these pathways, the autonomic function of the heart is affected by the light-dark dependency of the SCN’s regulatory circadian signal.

Although the circadian rhythmicity of the HRV signal has been confirmed in a number of studies [34,36–38], the signal is not used as a circadian marker or as a circadian phase predictor due to its high proneness to masking effects. Both heart

(5)

42

rate and heart rate variability can be subject to masking effects from body posture, meals, physical and mental activity, sleep/wake cycle, or stress [31]. The high number of possible masking effects makes the circadian component of the signal difficult to assess. One way of demasking the signal is applying a constant routine [39], but this method induces the same restrictions and impracticalities mentioned earlier. Extracting the circadian component could be done mathematically by either discriminating the dynamics of masking signals from the dynamics of the circadian component or by making use of other signals in parallel which are known to affect or mask the HRV. An example of this technique would involve monitoring physical activity simultaneously with heart rate, and then use this information to demask the HRV signal. Coupled with information about the person’s specific activities and behavior at different times, the magnitude of possible masking effects could be more reliably assessed. Using the differences in the dynamics between the masking and the circadian components can be done either through the selected features, or through the type of model used. Extracting specific features from the HRV signal, either in the time or frequency domain, which are stable enough to withstand environmental or behavioral influences, could lead to a more reliable circadian assessment of said signal. From a modeling perspective, the type of model will have a significant effect on how the signal is processed and interpreted to arrive at an output. Consequently, the processing of the HRV signal, the features extracted, and the type of model used are all interesting aspects that can be explored. Here, we will look into the RR intervals over 24 hours, and will try to determine whether the heart rate inter-beat intervals can be used as a predictor of circadian phase. The focus of the present modeling approach will be based on statistical machine learning. In this approach, the inputs and outputs of the training dataset are known, and given a model structure, the model then learns the weights of the coefficients that best optimize the resulting output. This approach will be used to train an autoregressive moving average with exogenous inputs (ARMAX) model of low order.

2.3 M

ATERIALS AND

M

ETHODS 2.3.1 Subjects and Protocol

Data from two individual studies was used for the training and validation of the model. The first study consisted of 14 healthy participants (6 females and 8 males) with an age of 23.9±1.9 (mean±SD). The data were collected at the Chronobiology department of the University of Groningen [40]. Participants were healthy, without depressive complaints (Beck Depression Inventory-II, Dutch version, BDI-II_NL < 8; [41]) or sleeping problems, non-smokers, had limited caffeine and alcohol consumption, and had not taken part in shift work or traveled across two or more time zones in the last three months. The participants had a BMI of 21.8±2.4. Based on the habitual weighted mean mid-sleep (hMS) over one week using the Munich Chronotype Questionnaire (MCTQ) [42], the habitual mid-sleep had to be between

(6)

43

2

4:15 and 6:09. The average hMS of the participants was 5:00±0:35. Late chronotypes

were selected for purposes not relevant to this study. However, this data were also used in another study where this criterion was of relevance [40]. The study was approved by the medical ethical board of the University of Groningen and all participants signed an informed consent form.

Participants wore an Actiwatch 2 (Philips Respironics, Pittsburgh, USA) on their non-dominant wrist for the duration of the study, which monitored activity levels and light exposure in terms of illuminance (lux). The participants were instructed to leave the watch uncovered to ensure continuous light data. On day 3, the RR intervals were measured for 24 hours using the Equivital Sensory Electronics Module (Philips Respironics, Pittsburgh, USA). The RR interval is defined as the time between two successive R-peaks in an electrocardiogram. Following the RR interval data collection, a full nighttime saliva sampling was collected at one hour intervals. Sampling started 9 hours prior to mid-sleep and ended 5 hours post mid-sleep, yielding an overnight melatonin profile. The data were collected between July and August. This dataset, from now on referred to as dataset A, was used to train the model.

The second study consisted of 24 participants (12 females and 12 males) with an age of 27.1±3.8. The data were collected at Philips Research in Eindhoven. Participants were healthy without pulmonary, cardiac, or sleeping disorders, not taking over the counter medication, non-smokers, consumed less than 3 units of alcohol per week, had a BMI of 22.4±2.9, did not do more than 4 hours of physical exercise per week, consumed less than 350mg of caffeine per day [37], and had not taken part in shift work or traveled across two or more time zones in the last three months. No restrictions were imposed on the chronotype of the participants. The study was approved by the internal ethical board of Philips Research and all participants signed an informed consent form.

During a two week collection period, the participant wore an Actiwatch Spectrum (Philips Respironics, Pittsburgh, USA) on the non-dominant wrist at all times to measure light exposure, and activity rhythms. The participants were instructed to keep the watch uncovered to ensure continuous light data, and to press the event marker button when turning off the lights when going to bed and in the morning when getting up. Their sleep schedule was further monitored by keeping a sleep log [42]. An electrocardiogram (ECG) was recorded at a sampling rate of 256Hz for 30 hours continuously from day 1 to day 2 ensuring at least 24 hours of usable data would be recorded. The ECG monitor (TMSI, Oldenzaal, The Netherlands) was connected to the participant using a standard 3-lead ECG placement. On day 1, the participants were given 5 saliva samplers (Salivette, Sarstedt AG & Co, Nuembrecht, Germany). An hourly sampling schedule was agreed upon based on the person’s usual bedtime, counting back in 1 hour intervals. Compliance to the lighting conditions and the sampling times was assessed by having the participants press the

(7)

44

event marker button on the Actiwatch Spectrum at the time the saliva sample was taken. During the entire sampling period, starting one hour before the first sample, the participant wore blue light filtering glasses (LowBlueLights, Photonic Developments LLC, Walton Hills, USA). The same protocol was repeated on day 8. Each of these recording instances is regarded as session 1 and session 2 respectively. This data were collected between November and February. This dataset, from now on referred to as dataset B, was used to validate the model.

The saliva samples of both datasets were analyzed in one batch using the Bühlmann Direct Saliva Melatonin RIA (Bühlmann Laboratories AG, Schönenbuch, Switzerland). The intra-assay variance was 10.69% for the low controls and 10.03% for the high controls. Inter-assay variance was 11.30% for the low controls and 12.18% for the high controls. The functional sensitivity of the assay was 0.9pg/ml. The DLMO was calculated using a threshold of 3pg/ml instead of the 25% method for consistency with the second dataset which included only evening melatonin.

The characteristics of dataset A and dataset B were compared using Matlab version R2010b (The MathWorks, Natick, USA).

2.3.2 ARMAX Model

Time series models are used to identify and predict the future values of a time dependent signal based on statistical forecasting. Within the hierarchy of time series models, several model structures are available with varying degrees of complexity for different applications. Linear regression aims at modeling the relationship between an output variable and one or multiple input variables. Another time series model is the autoregressive (AR) model, in which the current value is a function of its past values given a time delay na. Continuing in this hierarchy, the moving average

(MA) is a model where the output is dependent on a weighted sum of the current and past values given a time delay nc. The autoregressive moving average (ARMA)

model is a combination of the two previous models. The ARMA model considers only stochastic trends in the signal, and is able to yield a more compact representation of processes than the AR or MA models individually [43–45]. ARMA models have been used in the context of biological systems to represent the dynamics of various systems [46]. In order to influence the behavior of the ARMA model, it is necessary to take into account external inputs. An extension of this model which does precisely that is the autoregressive moving average with exogenous inputs (ARMAX) model [47]. The ARMAX model contains the AR and MA parts as the ARMA, however, includes a linear combination of external time series (exogenous inputs). Much like the ARMA model, it is able to give very compact representations of the signal processes, yet offers the possibility to also consider deterministic trends.

(8)

45

2

𝐴(𝑞)𝑦(𝑡) = 𝐵(𝑞) [ 𝑢1(𝑡 − 𝑛𝑘) 𝑢2(𝑡 − 𝑛𝑘) ⋯ 𝑢𝑖(𝑡 − 𝑛𝑘) ] + 𝐶(𝑞)𝑒(𝑡) (1)

Where u(t-nk) are the delayed inputs, y(t) is the output, e(t) is the noise model, and

A,B,C are the model coefficients. The nk term corresponds to the delay and the q is

the backward shift operator. The variable e(t) is a white-noise disturbance generated by the algorithm. It is uncorrelated with the input signals, uncorrelated with itself, and has a constant variance over time. The model identifies long term trends in the input signals (exogenous inputs) which are combined linearly to produce the desired output. The current output depends on previous inputs and delayed inputs, dictated by the model parameters nb and nk respectively. As in AR

models, past output values also influence the current output, with a delay na

specified also by the model configuration. This implicit recursion takes into account past context beyond the actual order of the model. Incorporating the MA portion, the model includes a noise disturbance value, or stochastic white process, of order nc. All these parameters determine the value of the model coefficients A, B, C (2-4)

𝐴(𝑞) = 1 + 𝑎1𝑞−1+ 𝑎2𝑞−2+ ⋯ + 𝑎𝑛𝑎𝑞−𝑛𝑎 (2) 𝐵(𝑞) = [ 𝑏11+ 𝑏12𝑞−1+ ⋯ + 𝑏1𝑛𝑏𝑞 −𝑛𝑏+1 𝑏21+ 𝑏22𝑞−1+ ⋯ + 𝑏2𝑛𝑏𝑞 −𝑛𝑏+1 ⋯ 𝑏𝑖1+ 𝑏𝑖2𝑞−1+ ⋯ + 𝑏𝑖𝑛𝑏𝑞−𝑛𝑏+1] (3) 𝐶(𝑞) = 1 + 𝑐1𝑞−1+ 𝑐2𝑞−2+ ⋯ + 𝑐𝑛𝑐𝑞−𝑛𝑐 (4)

which are linearly combined to produce the output y(t). The values of the a, b, c weights were determined using an adaptive Gauss-Newton method [48]. The term B(q) can be considered a matrix of coefficients b with the number of columns determined by the nb parameter and the number of rows by the number of input

signals. As mentioned previously, the output of the model is influenced not only by the input and delayed input, but also by a delayed output. This relation can be illustrated in the flowchart depicted in Figure 2.1.

(9)

46

2.3.3 Model Parameters and Testing

In statistical machine learning, the goal is to construct a system that is able to learn to solve a problem given a set of example data containing inputs as well as their expected outputs. The example dataset can be used to select an appropriate model and to evaluate the performance that is expected from the model. The model selection can be done by training and validating the model using one of two methods. Given a large enough dataset, the data can be divided into a training set and a validation set. The model can then be trained using the training dataset, and then applied to the unseen data from the validation set. If the dataset is not large enough to create two datasets, the cross-validation technique can be used. A common form of this technique is the leave-one-out-cross-validation (LOOCV) [49]. In this method, the model is trained using all but one sample, and then the validation is carried out using that left out sample. This is then repeated by leaving out a different sample and training on the rest until all samples have been used for the validation. In order to carry out both the model selection and evaluation of the model performance, these methods need to be combined. One option is to have three datasets where the first one is used for training, the second for validation, and the third one for performance evaluation. A second option is to do a double-cross-validation, which is essentially a cross-validation where the data is divided into three subsets. The last possibility is using a combination of the two aforementioned techniques. The model selection can be carried out using the cross-validation technique applied on the training set and a second independent dataset is used for the performance evaluation. The model selection and performance evaluation in this study were carried out using this last method.

Prior to training the ARMAX model, the input signals were preprocessed and scaled. The activity and light data were sampled at 1 sample/minute. The RR intervals extracted from the ECG signal (256Hz) were naturally non-uniformly sampled; therefore the signal was resampled to match the sampling frequency of the other signals. A median filter, with a window size of 15 minutes, was applied to smooth the signals and remove noise. 24 hour segments of each signal were extracted, aligned and standardized. Standardization was done using equation (5).

𝑥𝑠𝑡𝑎𝑛𝑑= (𝑥 − 𝑥𝑚𝑒𝑎𝑛)⁄𝑥𝑠𝑑 (5)

Various processing schemes of the light trace were applied. First of all, clipping of the light intensity values at 1000lx was compared to no clipping. Furthermore, the light trace was processed using three different techniques and the performance in the model of each was compared, namely the light intensity as measured by the Actiwatch Spectrum in lux, using a logarithmic compression (6) [50], or using a power compression (7) [10,11]. These three light traces were individually used in the training of the models, and the performance of each was compared.

(10)

47

2

𝐼𝑝𝑜𝑤 = 𝐼(𝑡)0.5 (7)

The reference output signal was a 24 hour cosine with amplitude of 1, period of 24 hours, and a phase equal to the DLMO. When using a time series model such as the ARMAX, it is necessary that the input and output vectors are of the same length. Given that the input vector is a 24 hour segment, the output vector must also be comprised of 24 hours of data. Since the parameter of interest was phase, the simplest way to express phase as a circadian signal is with a cosine having this particular phase shift. Using higher harmonics to compose a melatonin profile would require a priori knowledge of representative frequencies, as well as amplitudes. This DLMO-coded cosine (8) was used as the output signal, where the maximum of the cosine corresponds to the DLMO of the participant.

𝑦(𝑡) = cos⁡(2𝜋𝑓𝑡 − 𝜑𝐷𝐿𝑀𝑂) (8)

The output of the ARMAX model was then fitted using a cosinor analysis, from which the maximum was obtained. This maximum was used as the DLMO prediction (DLMOpred), which was compared to the DLMO calculated from the melatonin

profile. Figure 2.2 provides a flowchart of the model.

Figure 2.2 Flowchart of DLMO prediction model

The ARMAX model was trained with dataset A and evaluated using dataset B. The best set of coefficients and delays was determined using LOOCV, by training on 10 participants and evaluating the results on 1. Having determined the best model parameters based on the Akaike information criterion (AIC) and standard deviation of the error, the model performance was evaluated using dataset B. Using these criteria, the best model in the training data was the one using RR intervals and light with an order configuration of na = 3, nb = 3 3, nc = 1, nk = 3 3. Building upon the

description of the ARMAX model, this model configuration can be interpreted as a third order model which looks at three points in each of the input signals (nb) and

(11)

48

three points in the output itself (na). The model includes a delay (dead time) of three

points for the input data (nk). There are two values for nb each with a value of 3,

indicating that the same number of points is taken into account from both input signal streams, each with the same dead time nk. The models were not retrained or

modified in any way; the signals from dataset B were simply processed in the same manner as those from dataset A. This was done using all possible input combinations of RR intervals, activity, and each of the three processed light signals with and without clipping, and the previously described DLMO-coded cosine output signal resulting in 27 ARMAX models of third order. All processing, modeling, and analysis was done using Matlab R2010b (The MathWorks, Natick, USA).

A sample of each of the input signals and the DLMO-coded cosine signal together with the model output can be found in Figure 2.3. The signals have been preprocessed as outlined before, standardized, and the light signal was power law compressed with no clipping.

Figure 2.3 Sample input and output signals, in normalized units

2.4 R

ESULTS

In dataset A, the heart rate measurement of three participants was excluded due to technical problems with the heart monitor. Out of the 48 recordings of dataset B, six were excluded due to technical problems with the ECG monitors, two participants failed to comply with the saliva sampling protocol, and four melatonin curves were excluded due to melatonin measurement errors.

(12)

49

2

There was no significant difference (p=0.72) between the mean DLMO values for

dataset A and B, 21:48±1:06 and 21:50±00:48 respectively. However in line with the inclusion criteria, there was a significant difference (p<0.01) in the habitual mid-sleep times. Participants in dataset A and B had an average mid-mid-sleep time of 5:06±00:47 and 3:49±00:45 respectively. Hence, the angle of entrainment between DLMO time and mid-sleep time was on average longer by about one hour for the subjects of dataset A compared to those in dataset B. In addition, there was a significant difference (p=0.01) between the mean age of both datasets, 23.7±1.8 and 26.7±4.0, but no significant difference (p=0.98) found between the mean BMI, 21.7± 2.5 and 21.8± 2.6 respectively.

Dataset B included two RR interval samples over 24 hours from each participant. Both instances were recorded in the same manner and there was no explicit element of intervention in the protocol that would lead to a difference in the DLMO values. However, during this one week period in between ECG recordings, the lighting conditions, sleep schedules, and essentially any other factors relevant to circadian timing could be different. Therefore, in order to eliminate the effects that these differences could have had on the comparison of phase estimations, the two sessions were randomized. This means that the two sessions were assigned to two groups randomly while ensuring that each participant was present in each group only once. The two sessions were analyzed separately and the average of the two is presented here. The estimate errors were not significant different (p<0.05) and therefore, they could be combined into one average error and one average standard deviation. The different scaling and light processing schemes were tested individually; however, only the results of the best performing scaling and light traces are shown. In order to give a comparison on the performance of the models, the two sessions for the validation dataset were reduced to include only the common participants resulting in a total of 16 participants, as opposed to 17 and 19 for the two sessions respectively. Nevertheless, the evaluation was also run using all participants and the results were not significantly different for any of the models (0.54≤p≤1.0)

(13)

50

Table 2.1 Average error (bias) defined as the difference between the measured DLMO and the predicted DLMO, and standard deviation of the error (SD error) for the ARMAX models using activity, RR intervals, and light. Pearson’s R coefficient indicating a significant correlation (p<0.05) are shown in bold.

Signals used in model Error, mean±SD

(minutes) Pearson's R p value

Light pow -37±56 0.548 0.03

Activity -08±81 0.307 0.47

RR Intervals +23±56 0.650 <0.01

Activity, RR Intervals -27±40 0.654 0.01

Activity, Light log +06±57 0.606 0.01

RR Intervals, Light pow +02±39 0.712 <0.01

Activity, RR Intervals, Light pow -05±39 0.722 <0.01

From the 27 models, only 7 models are presented since only the best performing scaling and light processing schemes are shown. The best performing models were selected based on the Akaike information criterion (AIC), resulting in models with varying levels of complexity. Scatter plots of the measured DLMO and the predicted DLMO for the 7 different models are shown in Figure 2.4. The third order ARMAX models yielded a compact representation of signal processes ranging from 7 to 15 coefficients depending on the number of signals used.

The average error can be thought of as a bias, or a calibration factor that can be applied uniformly to the outputs. On the other hand, the standard deviation of the error represents the error that can be expected among participants. Therefore, the standard deviation of the error provides a more representative evaluation of the performance of the model, as it shows the range of errors that can be expected.

(14)

51

2

Figure 2.4 Measured DLMO versus predicted DLMO for each subject using the 7 ARMAX models presented in Table 2.1. Only one of the two sessions is shown. The plot shows the linear fit and the

±95% confidence interval.

The best performing model was determined to be the RR interval and light model based on the standard deviation of the error and the Pearson’s correlation. Figure 2.5 shows the Bland-Altman plots [51–53] for this model applied to the two sessions. Bland-Altman plots are a standard method of comparing two measurements and determining the agreement between the two. In this case, the plots compare the output of the model to the reference values of DLMO. The systematic bias is displayed as a solid line, which in this case corresponds to +4 minutes for session 1 and 0 minutes for session 2 (an average bias of +2). The standard deviations of the error are also shown with dashed lines at 36 and 42 minutes for session 1 and 2 respectively (39 minutes on average). In addition, the 95% limits of agreement showing the magnitude of the error that can be expected for 95% of the population are shown with dash-dotted lines. This allows for easily assessing the presence of outliers.

(15)

52

a)

b)

Figure 2.5 Bland-Altman plots of the model with the lowest error and standard deviation of the error. The solid line labeled “Mean Diff” shows the mean difference between the predicted DLMO and the measured DLMO (bias), the dashed lines labeled “Mean Diff ± SD” show the mean plus/minus one

standard deviation, and the dashed lines labeled “Mean Diff ± 1.96*SD” show the 95% limits of agreement defined as the mean difference plus/minus 1.96*standard deviation. (a) Shows the results of the model for session 1 with an error of +4±36 minutes. (b) Shows

(16)

53

2

2.5 D

ISCUSSION

Phase estimation of the human circadian pacemaker is possible through the use of various physiological signals which rely strongly on the master circadian clock. However, in order to arrive at an accurate estimate without using DLMO, prolonged data collections are necessary. We presented a method that is able to predict human circadian phase based on only 24 hours of data recorded in ambulatory conditions. Models using combinations of activity, light, and RR intervals were presented, and it was found that the inclusion of RR intervals yielded the best performing models. RR intervals, or more broadly speaking HRV, are dependent on the autonomic nervous system which follows a circadian cycle. This circadian output signal is highly prone to masking effects, making it difficult to assess the pure circadian component of the SCN which is driving it. The ARMAX model structure presented here was able to extract circadian information from the different signals, particularly the RR interval signal, with only 24 hours of data. The masking effects from daily ambulatory activity did not degrade the result given the model’s moving average smoothing coupled with its autoregressive behavior. Through statistical machine learning, the ARMAX model determined the weight of each signal as well as the magnitude of the coefficients. In all models combining two or three signals, the RR interval signal was given the largest weight.

The gold standard of assessing circadian phase is the DLMO due to its close dependence on the circadian pacemaker and its robustness [5]. In the evaluation part of this study, two melatonin measurements were taken one week apart. One cannot assume that given ambulatory conditions and the lack of behavioral constraints, the DLMO would remain constant during this time. Nevertheless, as a test of the prediction power of using one DLMO to estimate a second DLMO, the first session was compared to the second session. It was found that this estimation yielded an error of -12±28 minutes (R=0.841, p<0.01). If the two DLMO values were assessed in successive days, one could assume that the correlation would probably be higher and the standard deviation of the difference would be smaller. However, using one DLMO value to estimate a second DLMO, without considering the data between the two sessions, is not ideal given the nature of ambulatory conditions and the lack of behavioral restrictions that could lead to shifts that cannot be easily assessed.

Systematically evaluating the results obtained from the different models, it can be seen that the models using individual signals yielded large standard deviation of the error values, with the RR intervals producing the lowest at 56 minutes (R=0.650, p<0.01). Combining signals resulted in an improvement of the results as would be expected from the inclusion of more information. However, the largest improvement was achieved from the combination of RR intervals and light. Light alone yielded a prediction error of +37±56 minutes (R=0.548, p=0.03) and RR intervals alone a prediction error of -23±56 minutes (R=0.650, p<0.01). When

(17)

54

combining the two signals, the prediction error dropped to +2±39 minutes (R=0.712, p<0.01). The combination of all three signals resulted in prediction errors very similar to those of the RR intervals and light model, with an error of -5±39 minutes (R=0.722, p<0.01). Therefore, the inclusion of activity did not provide the model with new information that could improve the estimates of circadian phase. To determine which of the signals provided redundant information, the covariance between signals was calculated. The covariance between the activity and light traces was 0.11, while the covariance between the activity and RR intervals was -0.68. This suggests that the information found in the activity signal was already well represented by the RR interval signal. Depending on the randomization of the two recording sessions, different standard deviations of the error were obtained. The average error was always constant. The randomization was iterated 400 times and the average standard deviation of the error along with the 95% confidence intervals of the standard deviation of the error were calculated. For the selected model, the 95% confidence interval was ±3 minutes, resulting in a range of standard deviations of the error of 36 to 42 minutes. Given the progression of the improvement, it seems that the RR intervals alone provided the bulk of the information for the phase estimation. Once the other signals were included, the estimates were shifted in time, yielding lower bias values and an improvement in the standard deviation of the errors. As previously mentioned, the RR intervals were given the largest weighting by the statistical model training. The other signals provided a way of fine-tuning the phase estimates given their lower weightings.

The results showed that the model with RR intervals and light yielded the best estimates of circadian phase. The circadian dynamics of heart rate variability have been assessed in various studies such as those by Scheer et al. 1999, van Eekelen et al. 2004, Vandewalle et al. 2007, and recently by Boudreau et al. 2012. However, its relevance to phase estimation has been underestimated. The results presented here show that meaningful circadian information can be extracted from a 24 hour segment of RR intervals, and that this information can be used to estimate circadian phase given the appropriate type of model is used. In addition, incorporating light into the model resulted in improvements of the phase estimates. However, using these signals alone did not reveal reliable phase estimations, despite the light’s position as the primary zeitgeber. This is expected as the light trace alone contains no information regarding the intrinsic period of the person or the dynamics of the response to light. As a common characteristic in previously mentioned circadian phase estimation models, it seems that it is rather the combined use of endogenous circadian signals such as heart rate variability, and an environmental entrainment signal such as light, that is able to yield reliable estimates of the state of the circadian clock.

2.5.1 Future Work

The current model which uses RR intervals and light to estimate circadian phase has been tested in normal ambulatory conditions. Given the interaction of the signals, it

(18)

55

2

would be interesting to test the effects of light therapy, and whether this can be

accurately modeled using an autoregressive model as presented here. In addition, other subject populations such as shift workers or patients with circadian or sleep disorders can be addressed to test the limits of the model and determine the model’s applicability to other conditions of interest.

Heart rate variability is a very rich signal that is often studied both in the time and in the frequency domain. This study has focused only on the temporal sequence of RR intervals. Therefore, further work can be carried out taking into account the spectral information, which is more common practice, especially in the cardiology domain. Although the current models showed the importance of the RR intervals as a possible predictor of circadian phase, the magnitude of the circadian component in this signal is still not clear. Therefore, a more in depth analysis of HRV from a chronobiological point of view could shed light into the mechanism by which the circadian pacemaker influences heart rate variability.

In addition to the signals tested here, other signals which can be collected non-invasively can be evaluated in future models. Skin temperature has been shown to possess a circadian rhythm [54,55] and has also been used in a linear regression model [21] and a non-linear neural network model [22] to estimate circadian phase. Nevertheless, when considering other signals, emphasis must be given to the use of signals collected non-invasively, in ambulatory conditions, and for periods of time in the order of 24 hours. This is a need that is still present and it remains an ongoing challenge to provide accurate circadian phase estimations readily and reliably, with as little patient burden as possible. Providing practical methods that are less time consuming and less expensive, while still delivering accurate results like the one we presented here, can help advance chronotherapeutic and chronomedicine applications.

2.6 A

CKNOWLEDGEMENT

This work was supported by the EU Marie Curie Network iCareNet under grant number 264738.

We thank Marijke Gordijn and Moniek Geerdink of the University of Groningen for the collection of dataset A and the analysis of the melatonin samples.

2.7 R

EFERENCES

[1] Czeisler CA, Kronauer RE, Allan JS, Duffy JF, Jewett ME, Brown EN, et al. Bright light induction of strong (type 0) resetting of the human circadian pacemaker. Science 1989;244:1328–33.

[2] Wirz-Justice A. How to measure circadian rhythms in humans. Medicographia 2007;29:84–90.

(19)

56

[3] Klerman EB, Rimmer DW, Dijk DJ, Kronauer RE, Rizzo JF 3rd, Czeisler CA. Nonphotic entrainment of the human circadian pacemaker. Am J Physiol 1998;274:R991–6.

[4] Lack L, Bramwell T, Wright H, Kemp K. Morning blue light can advance the melatonin rhythm in mild delayed sleep phase syndrome. Sleep Biol Rhythms 2007;5:78–80. doi:10.1111/j.1479-8425.2006.00250.x.

[5] Lewy AJ, Sack RL. The dim light melatonin onset as a marker for circadian phase position. Chronobiol Int 1989;6:93–102.

[6] Gompper B, Bromundt V, Orgül S, Flammer J, Kräuchi K. Phase relationship between skin temperature and sleep-wake rhythms in women with vascular dysregulation and controls under real-life conditions. Chronobiol Int 2010;27:1778–96. doi:10.3109/07420528.2010.520786.

[7] Van Someren EJW, Nagtegaal E. Improving melatonin circadian phase estimates. Sleep Med 2007;8:590–601. doi:10.1016/j.sleep.2007.03.012. [8] Jewett ME, Forger DB, Kronauer RE. Revised limit cycle oscillator model of

human circadian pacemaker. J Biol Rhythms 1999;14:493–9.

[9] Hofstra WA, de Weerd AW. How to assess circadian rhythm in humans: A review of literature. Epilepsy & Behavior 2008;13:438–44. doi:10.1016/j.yebeh.2008.06.002.

[10] Kronauer RE, Forger DB, Jewett ME. Quantifying human circadian pacemaker response to brief, extended, and repeated light stimuli over the phototopic range. J Biol Rhythms 1999;14:500–15.

[11] St Hilaire MA, Klerman EB, Khalsa SBS, Wright KP, Czeisler CA, Kronauer RE. Addition of a non-photic component to a light-based mathematical model of the human circadian pacemaker. J Theor Biol 2007;247:583–99. doi:10.1016/j.jtbi.2007.04.001.

[12] Levi FA. The circadian timing system: a coordinator of life processes: chronobiological investigations. Implications for the rhythmic delivery of cancer therapeutics. IEEE Eng Med Biol Mag 2008;27:17–9. doi:10.1109/MEMB.2007.907361.

[13] Klerman EB, Gershengorn HB, Duffy JF, Kronauer RE. Comparisons of the variability of three markers of the human circadian pacemaker. J Biol Rhythms 2002;17:181–93.

[14] Pandi-Perumal SR, Smits M, Spence W, Srinivasan V, Cardinali DP, Lowe AD, et al. Dim light melatonin onset (DLMO): a tool for the analysis of circadian phase in human sleep and chronobiological disorders. Prog

Neuropsychopharmacol Biol Psychiatry 2007;31:1–11.

doi:10.1016/j.pnpbp.2006.06.020.

[15] Weinert D. Circadian temperature variation and ageing. Ageing Res Rev 2010;9:51–60. doi:10.1016/j.arr.2009.07.003.

[16] Brown EN, Choe Y, Luithardt H, Czeisler CA. A statistical model of the human core-temperature circadian rhythm. Am J Physiol Endocrinol Metab 2000;279:E669–83.

(20)

57

2

[17] Jewett ME, Forger DB, Kronauer RE. Revised limit cycle oscillator model of

human circadian pacemaker. J Biol Rhythms 1999;14:493–9.

[18] Kronauer RE, Forger DB, Jewett ME. Quantifying human circadian pacemaker response to brief, extended, and repeated light stimuli over the photopic range. J Biol Rhythms 1999;14:500–15.

[19] St Hilaire MA, Klerman EB, Khalsa SBS, Wright KP Jr, Czeisler CA, Kronauer RE. Addition of a non-photic component to a light-based mathematical model of the human circadian pacemaker. J Theor Biol 2007;247:583–99. doi:10.1016/j.jtbi.2007.04.001.

[20] Mott C, Dumont G, Boivin DB, Mollicone D. Model-based human circadian phase estimation using a particle filter. IEEE Trans Biomed Eng 2011;58:1325– 36. doi:10.1109/TBME.2011.2107321.

[21] Kolodyazhniy V, Späti J, Frey S, Götz T, Wirz-Justice A, Kräuchi K, et al. Estimation of Human Circadian Phase via a Multi-Channel Ambulatory Monitoring System and a Multiple Regression Model. J Biol Rhythms 2011;26:55–67. doi:10.1177/0748730410391619.

[22] Kolodyazhniy V, Späti J, Frey S, Götz T, Wirz-Justice A, Kräuchi K, et al. An improved method for estimating human circadian phase derived from multichannel ambulatory monitoring and artificial neural networks. Chronobiol Int 2012;29:1078–97. doi:10.3109/07420528.2012.700669. [23] Kubota T, Uchiyama M, Suzuki H, Shibui K, Kim K, Tan X, et al. Effects of

nocturnal bright light on saliva melatonin, core body temperature and sleep propensity rhythms in human subjects. Neurosci Res 2002;42:115–22. [24] Smith MR, Cullnan EE, Eastman CI. Shaping the light/dark pattern for circadian

adaptation to night shift work. Physiol Behav 2008;95:449–56. doi:10.1016/j.physbeh.2008.07.012.

[25] Levi FA. Chronotherapeutics: the relevance of timing in cancer therapy. Cancer Causes and Control 2006;17:611–21.

[26] Levi FA OA. Circadian clocks and drug delivery systems: impact and opportunities in chronotherapeutics. Expert Opinion on Drug Delivery 2011;8:1535–41.

[27] Task Force of The European Society of Cardiology and The North American Society of Pacing and Electrophysiology. Heart rate variability: Standards of measurement, physiological interpretation, and clinical use. European Heart Journal 1996;17:354–81.

[28] Scheer FAJL, Van Doornen LJP, Buijs RM. Light and diurnal cycle affect autonomic cardiac balance in human; possible role for the biological clock. Auton Neurosci 2004;110:44–8. doi:10.1016/j.autneu.2003.03.001.

[29] Bilan A, Witczak A, Palusiński R, Myśliński W, Hanzlik J. Circadian rhythm of spectral indices of heart rate variability in healthy subjects. J Electrocardiol 2005;38:239–43.

[30] Freitas J, Lago P, Puig J, Carvalho MJ, Costa O, de Freitas AF. Circadian heart rate variability rhythm in shift workers. J Electrocardiol 1997;30:39–44.

(21)

58

[31] Hayano J, Sakakibara Y, Yamada M, Kamiya T, Fujinami T, Yokoyama K, et al. Diurnal variations in vagal and sympathetic cardiac control. Am J Physiol 1990;258:H642–6.

[32] Anders D, Vollenweider S, Cann J, Hofstetter M, Flammer J, Orgül S, et al. Heart-rate variability in women during 40-hour prolonged wakefulness. Chronobiol Int 2010;27:1609–28. doi:10.3109/07420528.2010.504317. [33] Hu K, Ivanov PC, Hilton MF, Chen Z, Ayers RT, Stanley HE, et al. Endogenous

circadian rhythm in an index of cardiac vulnerability independent of changes in behavior. PNAS 2004;101:18223–7. doi:10.1073/pnas.0408243101. [34] Boudreau P, Yeh WH, Dumont GA, Boivin DB. A circadian rhythm in heart rate

variability contributes to the increased cardiac sympathovagal response to awakening in the morning. Chronobiol Int 2012;29:757–68. doi:10.3109/07420528.2012.674592.

[35] Buijs RM, la Fleur SE, Wortel J, Van Heyningen C, Zuiddam L, Mettenleiter TC, et al. The suprachiasmatic nucleus balances sympathetic and parasympathetic output to peripheral organs through separate preautonomic neurons. J Comp Neurol 2003;464:36–48. doi:10.1002/cne.10765.

[36] Scheer FA, van Doornen LJ, Buijs RM. Light and diurnal cycle affect human heart rate: possible role for the circadian pacemaker. J Biol Rhythms 1999;14:202–12.

[37] Van Eekelen APJ, Houtveen JH, Kerkhof GA. Circadian variation in base rate measures of cardiac autonomic activity. Eur J Appl Physiol 2004;93:39–46. doi:10.1007/s00421-004-1158-6.

[38] Vandewalle G, Middleton B, Rajaratnam SMW, Stone BM, Thorleifsdottir B, Arendt J, et al. Robust circadian rhythm in heart rate and its variability: influence of exogenous melatonin and photoperiod. J Sleep Res 2007;16:148– 55. doi:10.1111/j.1365-2869.2007.00581.x.

[39] Aoyagi N, Ohashi K, Yamamoto Y. Frequency characteristics of long-term heart rate variability during constant-routine protocol. American Journal of Physiology - Regulatory, Integrative and Comparative Physiology 2003;285:171–6.

[40] Geerdink M, Beersma D, Hommes V, Gordijn M. Phase advancing the human circadian system with short pulses (30 min) of blue light exposure. Abstracts of the XII Congress of the European Biological Rhythms Society 2011:136. [41] Beck AT, Steer RA, Ball R, Ranieri W. Comparison of Beck Depression

Inventories -IA and -II in psychiatric outpatients. J Pers Assess 1996;67:588– 97. doi:10.1207/s15327752jpa6703_13.

[42] Roenneberg T, Wirz-Justice A, Merrow M. Life between Clocks: Daily Temporal Patterns of Human Chronotypes. J Biol Rhythms 2003;18:80–90. doi:10.1177/0748730402239679.

[43] Box GEP, Jenkins GM, Reinsel GC. Time Series Analysis: Forecasting and Control. 4th ed. Wiley; 2008.

(22)

59

2

[45] Ganesh U, Krishna M, Hariprasad S, Rau S. Review on models for generalized

predictive controller, Chennai, India: 2011, p. 418–24.

[46] Perrott MH, Cohen RJ. An efficient approach to ARMA modeling of biological systems with multiple inputs and delays. IEEE Trans Biomed Eng 1996;43:1– 14. doi:10.1109/10.477696.

[47] Hannan EJ. The Identification and Parameterization of Armax and State Space Forms. Econometrica 1976;44:713. doi:10.2307/1913438.

[48] Wills A, Ninness B. On Gradient-Based Search for Multivariable System Estimates. IEEE Trans Autom Control 2008;53:298–306. doi:10.1109/TAC.2007.914953.

[49] Stone M. Cross-Validatory Choice and Assessment of Statistical Predictions. J Roy Statist Soc Ser B 1974;36:111–47. doi:10.2307/2984809.

[50] Dumont M, Beaulieu C. Light exposure in the natural environment: relevance to mood and sleep disorders. Sleep Med 2007;8:557–65. doi:10.1016/j.sleep.2006.11.008.

[51] Altman D, Bland J. Measurement in Medicine: The Analysis of Method Comparison Studies. J Roy Statist Soc Ser D 1983;32:307–17. doi:10.2307/2987937.

[52] Bland J, Altman D. Statistical Methods for assessing agreement between two methods of clinical measurement. The Lancet 1986;327:307–10. doi:10.1016/S0140-6736(86)90837-8.

[53] Bland J, Altman D. Measuring agreement in method comparison studies. Stat Methods Med Res 1999;8:135–60. doi:10.1177/096228029900800204. [54] Kräuchi K, Wirz-Justice A. Circadian rhythm of heat production, heart rate, and

skin and core temperature under unmasking conditions in men. American Journal of Physiology - Regulatory, Integrative and Comparative Physiology 1994;267:819–29.

[55] Kräuchi K, Cajochen C, Werth E, Wirz-Justice A. Functional link between distal vasodilation and sleep-onset latency? American Journal of Physiology - Regulatory, Integrative and Comparative Physiology 2000;278:741–8.

(23)

Referenties

GERELATEERDE DOCUMENTEN

Heart rate variability analyses of 24-hour electrocardiograms from sleep onset insomnia patients show an altered SDNN profile, with its maximum occurring shortly before

This study, however, serves as an initial investigation to pilot the assessment of circadian phase estimation models based on RR intervals, activity levels, and

Models have been presented based on heart rate and heart rate variability features which provide accurate circadian phase estimates in both healthy and sleep onset

Ambulatory assessment of human circadian phase and related sleep disorders from heart rate variability and other non-invasive physiological measurements.. Gil

The solid line labeled “Mean Diff” shows the mean difference between the predicted DLMO and the measured DLMO (bias), the dashed lines labeled “Mean Diff ± SD” show the

Ambulatory assessment of human circadian phase and related sleep disorders from heart rate variability and other non-invasive physiological measurements.. University

Ambulatory assessment of human circadian phase and related sleep disorders from heart rate variability and other non-invasive physiological measurements.. Gil

Estas limitaciones han sido la motivación para el trabajo en esta tesis de desarrollar un método para estimar con exactitud la fase circadiana que sea no-invasivo, se puede