Next Article in Journal
Global Geometry of Bayesian Statistics
Next Article in Special Issue
Entropy in Heart Rate Dynamics Reflects How HRV-Biofeedback Training Improves Neurovisceral Complexity during Stress-Cognition Interactions
Previous Article in Journal
Biometric Identification Method for Heart Sound Based on Multimodal Multiscale Dispersion Entropy
Previous Article in Special Issue
Complexity of Cardiotocographic Signals as A Predictor of Labor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Complexity Analysis of EEG, MEG, and fMRI in Mild Cognitive Impairment and Alzheimer’s Disease: A Review

College of Information and Computer, Taiyuan University of Technology, Taiyuan 030024, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Entropy 2020, 22(2), 239; https://doi.org/10.3390/e22020239
Submission received: 21 January 2020 / Revised: 15 February 2020 / Accepted: 17 February 2020 / Published: 20 February 2020

Abstract

:
Alzheimer’s disease (AD) is a degenerative brain disease with a high and irreversible incidence. In recent years, because brain signals have complex nonlinear dynamics, there has been growing interest in studying complex changes in the time series of brain signals in patients with AD. We reviewed studies of complexity analyses of single-channel time series from electroencephalogram (EEG), magnetoencephalogram (MEG), and functional magnetic resonance imaging (fMRI) in AD and determined future research directions. A systematic literature search for 2000–2019 was performed in the Web of Science and PubMed databases, resulting in 126 identified studies. Compared to healthy individuals, the signals from AD patients have less complexity and more predictable oscillations, which are found mainly in the left parietal, occipital, right frontal, and temporal regions. This complexity is considered a potential biomarker for accurately responding to the functional lesion in AD. The current review helps to reveal the patterns of dysfunction in the brains of patients with AD and to investigate whether signal complexity can be used as a biomarker to accurately respond to the functional lesion in AD. We proposed further studies in the signal complexities of AD patients, including investigating the reliability of complexity algorithms and the spatial patterns of signal complexity. In conclusion, the current review helps to better understand the complexity of abnormalities in the AD brain and provide useful information for AD diagnosis.

1. Introduction

Alzheimer’s disease (AD) is the most prevalent form of neurodegenerative dementia and includes a set of symptoms, such as memory loss and cognitive decline, that affect the ability to engage in daily activities and processes, including attention, thinking, orientation, or language [1,2]. In AD patients, proteins accumulate in the brain, forming amyloid plaques and neurofibrillary tangles, which have been shown to be associated with local synaptic disruptions [3,4]. Eventually, AD leads to the loss of connections between nerve cells, suggesting that AD is a disconnectivity disease. There are currently two recognized predementia stages: subjective cognitive impairment (SCI) and mild cognitive impairment (MCI) [5,6]. SCI refers to an individual’s main complaint of cognitive impairment with a lack of objective evidence of cognitive impairment or pathology. In recent years, SCI has become a hot topic in the research field of AD [5,7]. MCI increases the risk of and is an important risk factor for AD dementia, thus becoming an important target for early diagnosis of and intervention for AD [6]. Both SCI and MCI patients are at great risk of developing AD. Therefore, an in-depth understanding of the mechanisms involved in the early diagnosis and effective treatment of AD is crucial.
Brain imaging analyses have been widely used to explore the mechanisms of AD [8,9,10] and improve the accuracy of AD diagnosis [11,12]. Because the brain is a highly complex system and brain signals have complex nonlinear dynamics, there has been increasing interest in complexity analyses by using brain imaging data such as electroencephalograms (EEG), magnetoencephalogram (MEG), and functional magnetic resonance imaging (fMRI) [13,14,15]. Most studies have analyzed brain signals from a single channel, such as the signals from an electrode in EEG, a channel in MEG, or a voxel in fMRI. Recently, the complexity of brain signals has been widely used to better understand the complexity of abnormalities in the AD brain. Adequate study of brain imaging modalities provides an opportunity to outline the mechanisms underlying AD and useful information for its diagnosis [16,17,18]. More recently, some studies have proposed that the levels of complexity are potential biomarkers for identification in the early diagnosis of AD [19,20]. To date, there is no comprehensive review that summarizes the different imaging modalities and explains the complexity of abnormalities in the AD brain.
In the present review, we systematically examined 126 identified studies on the complexity of AD from 2000 to 2019. We aim to review the complexity indexes that can accurately represent the functional lesion in AD and outline the better complexity indicators. In addition, by analyzing changes in patients through general trends and comparative studies of brain regions, we identified our knowledge gaps as well as new issues for future research that can serve as a starting point for future applications of complexity analysis for AD patients.

2. Methods

2.1. The Analysis of Complexity

Entropy (En) is one of the most commonly used nonlinear concepts in evaluating the dynamic characteristics of signals [21]. This concept is an index of complexity analysis reflecting the degree of system confusion in a time series. These methods combine the complexity of the signal with its unpredictability: irregular signals are more complex than regular ones because they are more unpredictable. Some researchers believe that these techniques can be used to analyze time series in the time domain or frequency domain. In the time domain, entropy mainly reflects the changes in time, and these analyses are constantly improving. Approximate entropy (ApEn) is an indicator of the overall characteristics of the response signal from the point of view of the complexity of the signal. It is useful for small datasets and is effective for discriminating the signal from random signals [22,23]. Then, this index was replaced by sample entropy (SampEn), introduced by Richman and Moorman [24]. The sample entropy algorithm does not include a comparison to its own data; it is the exact value of the negative average natural logarithm of the conditional probability and has good consistency [25]. Fuzzy entropy (FuzzyEn) uses the exponential fuzzy similarity measure formula, which is more stable than the sample entropy algorithm [26]. Permutation entropy (PeEn) is a method for measuring nonstationary time series irregularities. PeEn considers only the grades of the samples but not their metrics [27]. PeEn has certain advantages over the other commonly used entropy metrics, including its simplicity, low computational complexity without further model assumptions, and robustness in the presence of observations and dynamic noise [27,28]. It has been successfully applied to EEG analyses and has been reported to be a good biomarker for distinguishing normal elderly individuals from patients with MCI and AD [29,30]. However, these methods mostly consider features at a single scale and can reflect only one aspect of the brain signal. Researchers have argued that multiscale entropy-based approaches better reflect the gradual transition process from coarse-grained entropy to fine-grained entropy, which can well reflect the complexity of biological signals on different time scales.
Although they continue to be rigorous and widespread methods used in the analysis of the frequency domain, linear decomposition methods, such as spectral analysis, have recently been suggested to lead to a loss of unique information that is orthogonal to average activity [31,32]. Renyi entropy (ReEn) is a generalization of Shannon entropy (ShEn), collision entropy, and minimum entropy, and it quantifies the diversity, uncertainty, or randomness of the system. Renyi entropy forms the basis of the concept of generalized dimensionality [33,34]. Tsallis entropy (TsEn) is nonexpansive [35]. For a composite system composed of two independent subsystems, it is not a simple sum of the entropy of two systems [36,37]. Spectral entropy (SpecEn) was developed to quantify the flatness of a spectrum [36,38]. SpecEn characterizes the distribution of power spectral density (PSD) by assessing disorder in the spectrum.
In addition to the entropy method, there are many other methods for assessing complexity, such as the Hurst exponent (HE), the Lempel-Ziv complexity (LZC), the correlation dimension (D2), and the fractal dimension (FD). The HE is mainly used to measure the long-term memory and fractal dimension of a time series [39]. The LZC reconstructs the original time series into a binary sequence [40]. The D2 and the largest Lyapunov exponent (LLE) were the first nonlinear techniques applied to EEG and MEG signals [41,42]. However, the calculation of D2 and LLE requires the signals to be stationary and long enough [43,44], which cannot be achieved for physiological data [45,46]. The FD has proven to be a reliable indicator for identifying healthy and pathological brains, and it can track changes in the complexity of neuronal dynamics, which might be related to cognitive or perceptual impairments [47]. Higuchi’s fractal dimension (HFD) is a fast computational method for obtaining the FD of a time series signal [48], even when very few data points are available. In addition, HFD provides a more accurate way to measure signal complexity [49,50], and it has proven to be an effective way to distinguish between AD patients and normal subjects.
Table 1 briefly introduces some widely used complexity methods. Although there are a large number of methods to assess complexity, entropy is the most popular. There are some problems with these methods, such as missing information, sensitivity to noise, and inaccurate results. The entropy method is advantageous in that it requires only a small amount of analysis data, possesses a strong anti-interference ability, and involves a simple algorithm. Different complexity analysis methods have their own advantages and disadvantages, and in this paper summarize their use in the analysis of brain signals acquired by different modalities in AD.

2.2. Literature Search

We examined the use of complexity techniques in the brain imaging of AD patients by performing an overview of these studies. Preferred Reporting Items for Systematic Reviews and Meta-Analyses (PRISMA) [62] was used to identify studies and narrow the collection for this review. We performed a search on Web of Science and PubMed using the following group of keywords: (“Complexity analysis” OR “Nonlinear dynamical analysis” OR “Lempel-Ziv complexity” OR “fractal dimension” OR “Hurst exponent” OR “entropy” OR “correlation dimension”) AND (“Alzheimer’s disease” OR “Mild Cognitive Impairment” OR “Subjective Cognitive Impairment”). References from 2000 until 2019 were used for further analysis. As shown in Figure 1, after excluding unqualified studies, this review narrowed the original count of 382 studies to the final count of 126 studies. Studies were divided into three categories: EEG (64%), MEG (28%), and fMRI and functional near-infrared spectroscopy (fNIRS) (7%) (Figure 2A). Various methods have been developed to examine the different types of brain imaging modalities, so the current status of these studies will also be described in the corresponding sections below. Unsurprisingly, EEG data are widely used in nonlinear analyses, accounting for 64% of all identified studies (Figure 2A). The four most commonly used analysis methods in the reviewed articles were time-domain entropy (TD-En), frequency-domain entropy (FD-En), LZC, and FD. The trends in the number of different techniques used in these brain imaging studies are shown in Figure 2B.

3. Results

3.1. Complexity Analysis of EEG Signals in AD

A large number of nonlinear methods have been applied to analyze the characteristics of brain activity in patients with AD, and numerous interesting results have been found. Since resting-state data are not influenced by task-related activation or differences in motivation or performance, these recordings provide more reliable estimates of brain adaptability [2,63]. Recordings of resting brain activity and task-related recordings exhibit similar network dynamics [64,65], and resting states often reflect the contribution of networks with the most metabolic activity [66]. The EEG signal has the advantage of high time resolution [67], and we found that the signals have been mainly analyzed in different frequency bands and from electrodes to reflect the variation in different signal values [68].

3.1.1. Complexity Analysis in Entropy

In this section, we review the signal complexity of the resting-state electroencephalogram (rsEEG) in SCI, MCI, and AD patients compared with normal controls (NCs). Several studies have shown that multiple complexity methods, such as LZC, entropy complexity, and other complexity features, differ among SCI, MCI, AD, and control subjects when applied to EEG signals. Hogan et al. [69] found that the entropy in MCI subjects was low. A recent study reported that in all channels, the complexity values of the EEG signals from AD patients were shown to be below those from SCI patients. It has been demonstrated that ApEn [70,71] and SampEn [72,73] in EEG signals are significantly reduced in MCI and AD patients compared to healthy individuals [74,75]; Garn et al. used different methods [76] to explore the complexity of EEG signals from AD patients and age-matched control subjects. In recent years, studies have included LZC, distance-based LZC [77], ApEn, SampEn, multiscale sample entropy (MSE), and FuzzyEn analyses [78]. Consistent results were found in the EEGs of patients with AD, including a significant reduction in complexity at electrodes P3, P4, O1, and O2 placed over the parietal, occipital, and temporal regions compared to healthy individuals. We found that at the MCI stage, the medial temporal lobe, associated with short-term memory, is affected, and the lateral temporal lobe and parietal lobe [79] are also affected. In the moderate stage of AD, the frontal lobe is affected. During the severe stage of AD, the occipital lobe is affected [18]. Multiple entropy methods have been used to study the brain states that develop in the transition from healthy conditions to AD. Most of the studies have focused on particular areas in the brain. Figure 3 presents comparative values of entropy shown over five regions in AD, MCI, and NC subjects. AD and MCI patients had lower En values in the five regions ( E n A D < E n M C I < E n C o n t r o l ), and significant differences were observed among the frontal, temporal, and central regions. These results suggest that the EEG signals in the brains of AD and MCI patients had significantly less complexity in the frontal, temporal, and central regions than those in the NC subjects. Furthermore, AD patients exhibit the lowest complexity and the greatest regularity. As expected, the complexity of the EEG signals gradually decreases with disease development, especially when comparing NC subjects with patients with AD.
We think that the reduction in the irregularity or complexity of brain signals can be described by a decrease in the dynamic complexity of the brain [80]. Our review demonstrated that aging and age-dependent diseases are frequently accompanied by losses in a broad range of physiological complexity or irregularity. A theory of discontinuous syndrome might explain the changes in AD: plaques and cell death can lead to the loss of connectivity between cortical neurons, which may lead to more regular brain signals (as recorded by cortical brain activity), thus destroying effective communication throughout the brain and producing the range of commonly seen AD symptoms.

3.1.2. Complexity Analysis in Multiscale Entropy

Entropy-based MSE analyses can measure the probability of sequences generating new information at different scales and have been applied to cognitive neuroscience. Deng et al. [82] studied changes on a 1–8 scale using multiscale weighted permutation entropy and found that the entropy in AD patients was decreased in the temporal, top, and right frontal occipital to the top and left occipital regions. Mizuno et al. [83] and Chai et al. [84] found that in large-scale entropy, AD and MCI patients had higher entropy than NCs. Studies [85] have shown that the variation in the complexity of EEG signals associated with cognitive impairment may be inconsistent on different time scales. We normalized the results of multiscale entropy and obtained the data presented in Figure 4. In the temporal, occipitoparietal, and right frontal regions, differences were statistically significant between groups. The entropy values on a 1–20 scale in each region in the AD, MCI, and NC groups are shown in Figure 4. On short scale factors, the entropy in the NCs was greater than that in the MCI and AD patients. On long scale factors, the entropy in the AD patients was greater than that in the MCI patients, and the entropy in the MCI patients was greater than that in NC subjects. A recent study also found that, on short time scales, compared to the NC group, the AD group and MCI group had lower values of entropy and showed relative preservation of coarse-grained entropy and selective loss in fine-grained entropy. This is consistent with studies that have found lower fine-grained entropy in AD patients than in healthy older adults [86]. Perhaps these changes accompany the development of the disease from its early stage to its relatively late stage. In this case, it may be a very useful quantitative biomarker of risk. These multiscale temporal features appear to arise from functional interactions of neural structural limitations.

3.1.3. Complexity Analysis in Frequency Entropy

Alqazzaz et al. [87] found that spectral results showed that EEG activity was slower in patients with AD and MCI. The SpecEn results showed that the frequency distribution of the power spectrum changed. These findings confirmed results from previous studies: the EEG signals of patients with AD and MCI gradually slowed down [76,88]. However, the physiological interpretation of all these changes is uncertain. A more scientific hypothesis is that significant brain cholinergic deficits are the basis of cognitive symptoms such as memory loss. The loss of neocortical cholinergic innervation in the modified cortex plays a key role in the EEG signal decreases associated with AD [89]. Similarly, because the cholinergic system regulates spontaneous cortical activity at low frequencies, this EEG signal decrease may also be due to the loss of the neurotransmitter acetylcholine, leading to a slowing of neural oscillations in AD. TsEn showed reductions in signal complexity in vascular dementia patients (AD) and MCI patients. In particular, the TsEn method has been shown to be a more promising complexity method for quantifying EEG changes [87,90]. Because of the speed of computation, it can serve as a theoretical basis for decision support tools in the expert diagnosis of AD [91]. Waser et al. [17] used the TsEn method to study differences between the EEGs of patients with AD and NCs and found significant differences in the t7 and t8 channels. There are also a large number of studies that have used multiple methods to explore complexity in AD. Al-Nuaimi et al. [78] found that for specific EEG frequency bands and channels, the HFD and LZC values of AD patients were significantly reduced compared to NCs. Coronel et al. [60] used automutual information (AMI), Shannon entropy, TsEn, MSE, and SpecEn to analyze the severity of AD, and the results showed that reduced complexity and AMI, SpecEn, and MSE values were associated with decreased Mini-Mental State Examination (MMSE) scores.
It is generally believed that AD leads to a decrease in high-frequency (alpha, beta, and gamma) power and an increase in low-frequency (delta and theta) power [92]. We averaged the values from five brain regions in each frequency band, resulting in the data presented in Figure 5. Figure 5 shows the differences in the frequency domain among the AD, MCI, and NC groups. On the one hand, the E n value in the delta ( δ ), theta ( θ ) and gamma ( γ ) bands ( E n A D > E n M C I >   E n C o n t r o l ) significantly increased. On the other hand, the En value in the alpha ( α ) band decreased ( α E n M C I   > α E n C o n t r o l > α E n A D ). Notably, α E n M C I was significantly higher than α E n c o n t r o l . This result may be related to a compensatory mechanism in patients with MCI during memory load and cognitive performance; for NCs, compensation is not required, and for AD patients, compensation is no longer possible [93]. The value of   β E n was lower in the AD and MCI patients than in the NCs ( β E n A D   < β E n M C I   < β E n C o n t r o l ).
It has been reported that different frequency bands reflect different brain dynamics [94]. We found that in applications for AD detection, the AD group showed lower complexity in different regions and sub-bands than the control group. This may be because high-frequency oscillations originate from short-range neural connections [95,96], while low-frequency oscillations include long-range neural connections [93,97]. Hence, the abnormal neural connectivity in patients with AD may be related to the abnormal complexity at different frequencies. Both the process of aging and the development of dementia has been associated with these low-frequency band increases [96]. This is partly due to the increasing local (rather than distributed) nature of the interactions between neuronal populations [98].

3.1.4. Complexity Analysis in Other Methods

Jeong et al. [99] found that in most EEG channels, AD patients had significantly lower FD values than NCs. In the detection of dementia, previous studies used the FD of the correlation dimension and HFD and found that the value of FD was lower in AD patients in the parietal and temporal regions compared to NCs [16,100]. Amezquita-Sanchez et al. [101] used box dimension (BD), HFD, Katz’s FD (KFD), and the integrated multiple signal classification and empirical wavelet transform (MUSIC-EWT) to diagnose MCI and AD patients with an accuracy of 90.3%. Al-Nuaimi et al. [102] studied HFD in EEGs for AD diagnosis, and they found that HFD is a promising EEG biomarker that can capture changes in the areas of the brain that are initially affected by AD. McBride et al. [103] researched complexity based on the LZC method to distinguish patients with early MCI (EMCI), AD, and NC, and they found that the EEG complexity features of specific bands with regional electrical activity provided promising results in distinguishing EMCI, AD, and NC. Liu et al. [77] used LZC and multiscale LZC methods for analysis and found significant differences between groups in the alpha-band in the parietal and occipital regions. Hornero et al. [74] used LZC to analyze EEGs and MEGs in patients with AD and found that LZC provides good insight into the characteristics of EEG background activity and the changes associated with AD. Through these studies, we found that the HFD and LZC of the EEG are potentially good biomarkers of AD diagnosis, as they are significantly lower in AD patients than in NCs.

3.1.5. Identification of AD

In this section, Table 2 shows the sensitivity, specificity, and accuracy in differentiating among AD, MCI, NC subjects were found with different nonlinear methods used in the EEG.

3.2. Complexity Analysis of MEG in AD and MCI

In this section, we review the signal complexity of the MEGs in MCI and AD patients compared with NC participants. The temporal resolution of MEG signals can reach the millisecond level, and the spatial resolution can be less than 2 mm. We found that the research could be generally divided based on the analysis of different brain regions to identify trends in these values. MEG is a noninvasive technique that allows recording of the magnetic fields generated by brain neuronal activity. MEG signals are independent of any reference point and are less affected by extracerebral tissues than EEG signals [106,107].

3.2.1. Complexity Analysis in Domain Entropy

Gómez et al. [108,109] analyzed MEG complexity based on cross-approximate entropy, which revealed decreases that indicated better synchronization in AD and MCI patients than in NC subjects. Using the ApEn, SampEn, and FuzzyEn methods to analyze MEG signals at 148 locations, it was found that the entropy in AD patients was lower than that in controls, suggesting that this neurological disorder may be accompanied by a regular increase in MEG activity. Hornero et al. [86] found that ApEn, SampEn, and MSE values in MEG data were lower in AD patients than in NCs. Juan P et al. [110] found that all PeEn values in the MCI group were larger than those in the normal group. Azami et al. [111] used the FuzzyEn, SampEn, and PeEn methods, and a 148-megabyte channel was analyzed to quantify the complexity of the signal. The FuzzyEn and SampEn values in AD patients were lower than those in the controls. AD patients showed significantly lower values than MCI subjects and NCs in almost all comparisons. Most studies have yielded information about the location of similar brain regions. Gómez et al. [112] reported MSE profiles that represented the SampEn values of each coarse-grained time series relative to the scale factor. Azami et al. [113] found that the values of multiscale dispersion entropy (MDE), multiscale permutation entropy (MPE), and MSE in AD patients were lower than those in NCs at short scale factors, while at long scale factors, the MDE and MSE values from AD subject signals had higher values [112]. In contrast, the MPE values at long scale factors were very similar for AD patients and NCs.
We found that most of the studies were divided based on the analysis of different brain regions and were analyzed on different scales. At low scale factors, the entropy value in AD patients was lower than that in NCs. For high scale factors, the values in AD patients were higher than those in controls. Figure 6 shows data for each region, and we report the average entropy values computed across the entire 1–20 scale factor range. In terms of the MEG signal, AD patients were reported to be more regular, less complex and more predictable than the controls, and these results were consistent with the EEG results [114].

3.2.2. Complexity Analysis in Frequency Entropy

Nonlinear analysis of frequency has also been reported in MEG-based studies by SpecEn and ratios. Poza et al. [115] studied the ratio of SpecEn (RSP). In the delta and theta bands, the RSP in AD patients was significantly higher than that in controls. However, in the beta and gamma bands, the RSP value was significantly lower in AD patients than in NCs. Regarding the spectral entropies, the results showed a statistically significant decrease in the value in the MEG signal in AD patients compared to NCs. Poza et al. [116] found that the spectral crest factor and both spectral turbulence and wavelet turbulence in AD patients were higher than those in NCs, which indicated that in AD patients, the oscillating signal was more regular. Bruner et al. [117] found that the SpecEn and TsEn values in patients with MCI were significantly lower than those in controls in the right lateral region, indicating a significant decrease in the irregularity of MEG signals in patients with MCI. All studies have shown that AD patients had slower brain activity than controls, which was reflected in a higher power in the lower frequency bands and lower power in the higher frequency bands.

3.2.3. Complexity Analysis in Other Methods

Gómez et al. [118] researched MEG background activity from AD and NC subjects using HFD and found that the value of HFD was less complex in AD patients, indicating an abnormal type of motility in AD. Shumbayawonda et al. [119] used LZC to research MEG signals in three groups: NCs, patients with subjective cognitive decline (SCD), and patients with MCI, and analyses were performed in theta, alpha, beta, and gamma bands. It was found that the LZC value in MCI patients was significantly lower than that in the control group and in SCD subjects, and the lower complexity was associated with smaller hippocampal volume. Another study, combining age with LZC, found that AD patients and controls showed a tendency of decreased LZC with age [120]. We found that both non-entropy and entropy methods for assessing complexity achieve the same results, but entropy methods were more widely used.

3.2.4. Identification of AD

In this section, Table 3 shows the sensitivity, specificity, and accuracy in differentiating among AD, MCI, NC subjects were found with different nonlinear methods used in the MEG.

3.3. Complexity Analysis of fMRI and fNIRS Signals in AD and MCI

The fMRI uses magnetic array imaging [127,128], while fNIRS uses hemoglobin in blood vessels to scatter near-infrared light [129,130]. In this section, we review signal complexity from fMRI and fNIRS in MCI and AD patients compared with NCs. A few studies have reported that biomarkers from fMRI and fNIRS signals, such as LZC, entropy, and other complexity characteristics, differ between MCI, AD, and NC subjects.
The fMRI signals have been used to detect functional abnormalities associated with neuropsychiatric and neurological disorders. Maxim et al. [131] applied the HE method to fMRI signals, and they found that the values of signals in the white matter were lower than those in the gray matter. Liu et al. reported that the complexity in certain brain regions (e.g., anterior cingulate gyrus and left cuneus) was reduced in a study of resting-state fMRI (rs-fMRI) signal complexity in familial AD patients [132]. Wang et al. [15] found significantly decreased PeEn values in the AD patient group compared with the MCI group and the normal group. Compared with the NC group, the complexity in the left wedge in the MCI group was also reduced. The complexity differences among the groups were mainly observed in the temporal, occipital, and frontal lobes. We found that AD patients had reduced mean whole-brain complexity in the gray matter and white matter compared to EMCI and NC subjects. At the regional level, five clusters showed significant differences in En, as illustrated in Figure 7. Niu et al. [133] extracted the average MSEs of the whole brain, gray matter, white matter, and cerebrospinal fluid using corresponding masks on all time scales. Only the gray matter showed a trend toward an entropy difference between the groups at scale factor six. Significant differences were found between the groups at scale factors two, four, five, and six, as shown in Figure 8. A significant difference was found in the right thalamus at scale factor two. A significant difference was found in the left superior frontal gyrus at scale factor four. Two significant differences were found in the right lingual gyrus and right insula at scale factor five. Five significant differences were found in the right superior temporal gyrus, left middle temporal gyrus, right olfactory cortex, left inferior occipital gyrus, and right supramarginal gyrus (SMG.R) at scale factor six. Grieder et al. [134] found that the AD group showed a lower global default mode network (DMN)-MSE than the NC group. A scientific explanation has been found for the reduced complexity of fMRI and fNIRS signals in AD. High regional functional homogeneity leads to lower complexity, so more differentially affected brain regions are found at high scale factors. Nerve cells are associated with complex dynamic processing in brain neural networks, and neuronal cell death leads to the loss of connectivity to local neural networks. It may be the death of neurons and the lack of neurotransmitters that lead to reduced irregularity in AD patients.

4. Discussion

Complexity methods applied to brain imaging data such as EEG, MEG and fMRI provide useful information for the diagnosis of AD using abnormal brain activity signals. This review combines previous findings with a larger overview and a further characterization of multiple modes for a better understanding of the functional lesion in AD. For the complexity of the single-channel time series, the development process of AD is clear and independent of the method used. The decline in AD may be due to plaques and cell death leading to loss of connectivity between cortical neurons, which may lead to more regular brain signals, thus destroying effective communication throughout the brain [135,136]. Furthermore, for each part of the brain, the trend is not consistent [137]. This may be related to the compensatory mechanisms that exist in the brain: when the synaptic structure slows down less, new synapses can be established to fill the gap, to change the connection pathway and establish connections with other regions or to increase the degree of added work, thus compensating for the altered brain function compensate, which indicates that the complexity of the AD brain changes [138,139]. The pattern of changes in this complexity is a good reflection of the pathological progression of AD and shows that complexity can be used as a biomarker to measure AD.
Complexity methods are suitable for the study of nonlinear brain changes and are sensitive to neurological changes associated with AD patients compared with normal subjects. The entropy method accounts for a large proportion of the complexity methods used. Better performance was exhibited at high scales, and when more brain regions were included in the analysis, the trends were more obvious. The exploration of test–retest reliability and improvements in entropy algorithms will provide great guidance for future applications. As the brain is a complex system in time and space, we can also study network entropy and spatio-temporal entropy in the future. While fMRI has a spatial resolution on the order of millimeters, only a small number of studies have applied complexity to fMRI data to date. Although the time resolution is not very high, it also reacts well and has been used to identify downward trends in different brain regions. It is also important to note that the potential utilization of the high spatial resolution in fMRI and fNIRS data can provide more in-depth information for AD brain dysfunction.
Complexity analysis of different types of brain imaging data in AD patients has yielded consistent results. The results showed consistent changes in that the signals in the brains of AD patients are slower, more regular, less complex, and less well organized than those of NCs. The reduction in the irregularity and complexity of brain signals in AD is the main finding obtained, and the occipital, frontal, parietal, and temporal areas are the most affected regions. We found that complexity can capture changes in areas of the brain that are initially affected by AD and accurately respond to its pathological mechanism. Complexity is a promising biomarker in reflecting the pathological mechanism of AD, and entropy is the more widely used of the numerous complexity indicators described in this review. For which entropy index is the best, more research is needed in the future to prove it. In general, different modalities for the same groups with large amounts of data were analyzed by choosing methods with high reliability and accuracy, the results of which will aid in truly understanding the functional lesion in AD. The results of the articles in this review can advance research on quantifying the complexity indexes of different subjects until clinical application is realized.

Author Contributions

Conceptualization: J.S., Y.N., Y.T. and C.F.; literature search: J.S., N.Z., J.X. (Jiayue Xue) and J.W.; validation: B.W., J.X. and J.S.; writing—original draft preparation: J.S.; writing—review and editing: B.W. and J.S.; supervision: J.X. (Jie Xiang); funding acquisition: J.X. (Jie Xiang). All authors have read and agreed to the published version of the manuscript.

Funding

This project was supported by the National Natural Science Foundation of China (61503272, 61305142, 61373101 and 61873178), the Natural Science Foundation of Shanxi (2015021090), the China Postdoctoral Science Foundation (2016M601287), the Shanxi Provincial Foundation for Returned Scholars, China (2016-037), and the Scientific Research Foundation for Returned Overseas Chinese Scholars.

Conflicts of Interest

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

References

  1. Stern, Y. Cognitive reserve in ageing and Alzheimer’s disease. Lancet Neurol. 2012, 11, 1006–1012. [Google Scholar] [CrossRef] [Green Version]
  2. Kumar, A.; Singh, A. A review on Alzheimer’s disease pathophysiology and its management: An update. Pharmacol. Rep. 2015, 67, 195–203. [Google Scholar] [CrossRef] [PubMed]
  3. Takahashi, R.H.; Nagao, T.; Gouras, G.K. Plaque formation and the intraneuronal accumulation of β-amyloid in Alzheimer’s disease. Pathol. Int. 2017, 67, 185–193. [Google Scholar] [CrossRef]
  4. Thies, W.; Bleiler, L. 2012 Alzheimer’s disease facts and figures. Alzheimer’s Dement. J. Alzheimer’s Assoc. 2012, 8, 131–168. [Google Scholar] [CrossRef]
  5. Buckley, R.F.; Villemagne, V.L.; Masters, C.L.; Ellis, K.A.; Rowe, C.C.; Johnson, K.; Sperling, R.; Amariglio, R. A Conceptualization of the Utility of Subjective Cognitive Decline in Clinical Trials of Preclinical Alzheimer’s Disease. J. Mol. Neurosci. 2016, 60, 354–361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Petersen, R.C. Mild cognitive impairment. Contin. Lifelong Learn. Neurol. 2016, 22, 404. [Google Scholar] [CrossRef]
  7. Visser, P.J.; Vos, S.; van Rossum, I.; Scheltens, P. Comparison of international working group criteria and national institute on Aging–Alzheimer’s association criteria for Alzheimer’s disease. Alzheimer’s Dement. 2012, 8, 560–563. [Google Scholar] [CrossRef]
  8. Blennow, K.; Mattsson, N.; Schöll, M.; Hansson, O.; Zetterberg, H. Amyloid biomarkers in Alzheimer’s disease. Trends Pharmacol. Sci. 2015, 36, 297–309. [Google Scholar] [CrossRef]
  9. Rosenberg, P.B.; Nowrangi, M.A.; Lyketsos, C.G. Neuropsychiatric symptoms in Alzheimer’s disease: What might be associated brain circuits? Mol. Asp. Med. 2015, 43, 25–37. [Google Scholar] [CrossRef] [Green Version]
  10. Ke, Q.; Zhang, J.; Wei, W.; Damaševičius, R.; Woźniak, M. Adaptive Independent Subspace Analysis of Brain Magnetic Resonance Imaging Data. IEEE Access 2019, 7, 12252–12261. [Google Scholar] [CrossRef]
  11. Acharya, U.R.; Fernandes, S.L.; WeiKoh, J.E.; Ciaccio, E.J.; Fabell, M.K.M.; Tanik, U.J.; Rajinikanth, V.; Yeong, C.H. Automated detection of Alzheimer’s disease using brain MRI images—A study with various feature extraction techniques. J. Med. Syst. 2019, 43, 302. [Google Scholar] [CrossRef]
  12. Bi, X.; Wang, H. Early Alzheimer’s disease diagnosis based on EEG spectral images using deep learning. Neural Netw. 2019, 114, 119–135. [Google Scholar] [CrossRef] [PubMed]
  13. Labate, D.; La Foresta, F.; Morabito, G.; Palamara, I.; Morabito, F.C. Entropic measures of EEG complexity in Alzheimer’s disease through a multivariate multiscale approach. IEEE Sens. J. 2013, 13, 3284–3292. [Google Scholar] [CrossRef]
  14. Luo, Q.; Xu, D.; Roskos, T.; Stout, J.; Kull, L.; Cheng, X.; Whitson, D.; Boomgarden, E.; Gfeller, J.; Bucholz, R.D. Complexity analysis of resting state magnetoencephalography activity in traumatic brain injury patients. J. Neurotrauma 2013, 30, 1702–1709. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Wang, B.; Niu, Y.; Miao, L.; Cao, R.; Yan, P.; Guo, H.; Li, D.; Guo, Y.; Yan, T.; Wu, J. Decreased complexity in Alzheimer’s disease: Resting-state fMRI evidence of brain entropy mapping. Front. Aging Neurosci. 2017, 9, 378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Staudinger, T.; Polikar, R. Analysis of complexity based EEG features for the diagnosis of Alzheimer’s disease. In Proceedings of the 2011 Annual International Conference of the IEEE Engineering in Medicine and Biology Society, Boston, MA, USA, 30 August–3 September 2011; pp. 2033–2036. [Google Scholar]
  17. Waser, M.; Deistler, M.; Garn, H.; Benke, T.; Dal-Bianco, P.; Ransmayr, G.; Grossegger, D.; Schmidt, R. EEG in the diagnostics of Alzheimer’s disease. Stat. Pap. 2013, 54, 1095–1107. [Google Scholar] [CrossRef]
  18. Sharma, A.; Rai, J.; Tewari, R. Relative Measures to Characterize EEG Signals for Early Detection of Alzheimer. In Proceedings of the 2018 5th International Conference on Signal Processing and Integrated Networks (SPIN), Noida, India, 22–23 February 2018; pp. 43–48. [Google Scholar]
  19. Atluri, G.; Padmanabhan, K.; Fang, G.; Steinbach, M.; Petrella, J.R.; Lim, K.; MacDonald, A., III; Samatova, N.F.; Doraiswamy, P.M.; Kumar, V. Complex biomarker discovery in neuroimaging data: Finding a needle in a haystack. Neuroimage Clin. 2013, 3, 123–131. [Google Scholar] [CrossRef] [Green Version]
  20. Polanco, J.C.; Li, C.; Bodea, L.-G.; Martinez-Marmol, R.; Meunier, F.A.; Götz, J. Amyloid-β and tau complexity—Towards improved biomarkers and targeted therapies. Nat. Rev. Neurol. 2018, 14, 22. [Google Scholar] [CrossRef]
  21. Shannon, C.E. A Mathematical Theory of Communication. Bell Syst. Tech. J. 1948, 27, 379–423. [Google Scholar] [CrossRef] [Green Version]
  22. Pincus, S. Approximate entropy (ApEn) as a complexity measure. Chaos Interdiscip. J. Nonlinear Sci. 1995, 5, 110–117. [Google Scholar] [CrossRef]
  23. Yentes, J.M.; Hunt, N.; Schmid, K.K.; Kaipust, J.P.; McGrath, D.; Stergiou, N. The appropriate use of approximate entropy and sample entropy with short data sets. Ann. Biomed. Eng. 2013, 41, 349–365. [Google Scholar] [CrossRef] [PubMed]
  24. Richman, J.S.; Lake, D.E.; Moorman, J.R. Sample Entropy. Methods Enzymol. 2004, 384, 172–184. [Google Scholar] [PubMed]
  25. Graff, B.; Graff, G.; Kaczkowska, A. Entropy measures of heart rate variability for short ECG datasets in patients with congestive heart failure. Acta Phys. Pol. B Proc. Suppl. 2012, 5, 153–158. [Google Scholar] [CrossRef]
  26. Kosko, B. Fuzzy entropy and conditioning. Inf. Sci. 1986, 40, 165–174. [Google Scholar] [CrossRef]
  27. Bandt, C.; Pompe, B. Permutation Entropy: A Natural Complexity Measure for Time Series. Phys. Rev. Lett. 2002, 88, 174102. [Google Scholar] [CrossRef]
  28. Zanin, M.; Zunino, L.; Rosso, O.A.; Papo, D. Permutation entropy and its main biomedical and econophysics applications: A review. Entropy 2012, 14, 1553–1577. [Google Scholar] [CrossRef]
  29. Kaufmann, A.; Kraft, B.; Michaleksauberer, A.; Weigl, L.G. Using Permutation Entropy to Measure the Electroencephalographic Effects of Sevoflurane. J. Am. Soc. Anesthesiol. 2008, 109, 448. [Google Scholar]
  30. Morison, G.; Tieges, Z.; Kilborn, K. Multiscale permutation entropy analysis of EEG in mild probable Alzheimer’s patients during an episodic memory paradigm. Alzheimer’s Dement. J. Alzheimer’s Assoc. 2012, 8, 522. [Google Scholar] [CrossRef]
  31. Faisal, A.A.; Selen, L.P.J.; Wolpert, D.M. Noise in the nervous system. Nat. Rev. Neurosci. 1966, 9, 292–303. [Google Scholar] [CrossRef]
  32. Marziani, E.; Pomati, S.; Ramolfo, P.; Cigada, M.; Giani, A.; Mariani, C.; Staurenghi, G. Evaluation of retinal nerve fiber layer and ganglion cell layer thickness in Alzheimer’s disease using spectral-domain optical coherence tomography. Investig. Ophthalmol. Vis. Sci. 2013, 54, 5953–5958. [Google Scholar] [CrossRef]
  33. Renyi, A. Probability Theory; North-Ho11 and Pub. Co.: Amsterdam, The Netherlands, 1970. [Google Scholar]
  34. Frank, R.L.; Lieb, E.H. Monotonicity of a relative Rényi entropy. J. Math. Phys. 2013, 54, 122201. [Google Scholar] [CrossRef] [Green Version]
  35. Tsallis, C. Possible generalization of Boltzmann-Gibbs statistics. J. Stat. Phys. 1988, 52, 479–487. [Google Scholar] [CrossRef]
  36. Vakkuri, A.; Yli-Hankala, A.; Talja, P.; Mustola, S.; Tolvanen-Laakso, H.; Sampson, T.; Viertiö-Oja, H. Time-frequency balanced spectral entropy as a measure of anesthetic drug effect in central nervous system during sevoflurane, propofol, and thiopental anesthesia & nbsp. Acta Anaesthesiol. Scand. 2004, 48, 145–153. [Google Scholar] [PubMed]
  37. Sarkar, S.; Das, S. Multilevel image thresholding based on 2D histogram and maximum Tsallis entropy—A differential evolution approach. IEEE Trans. Image Process. 2013, 22, 4788–4797. [Google Scholar] [CrossRef]
  38. Zaccarelli, N.; Li, B.-L.; Petrosillo, I.; Zurlini, G. Order and disorder in ecological time-series: Introducing normalized spectral entropy. Ecol. Indic. 2013, 28, 22–30. [Google Scholar] [CrossRef]
  39. Qian, B.; Rasheed, K. Hurst exponent and financial market predictability. In Proceedings of the IASTED Conference on Financial Engineering and Applications, Cambridge, MA, USA, 8–10 November 2004; pp. 203–209. [Google Scholar]
  40. Aboy, M.; Hornero, R.; Abásolo, D.; Álvarez, D. Interpretation of the Lempel-Ziv complexity measure in the context of biomedical signal analysis. IEEE Trans. Biomed. Eng. 2006, 53, 2282–2288. [Google Scholar] [CrossRef] [Green Version]
  41. Jeong, J.; Chae, J.H.; Kim, S.Y.; Han, S.H. Nonlinear dynamic analysis of the EEG in patients with Alzheimer’s disease and vascular dementia. J. Clin. Neurophysiol. 2001, 18, 58–67. [Google Scholar] [CrossRef] [Green Version]
  42. Sriraam, N. Correlation dimension based lossless compression of EEG signals. Biomed. Signal Process. Control 2012, 7, 379–388. [Google Scholar] [CrossRef]
  43. Grassberger, P.; Procaccia, I. Characterization of strange attractors. Phys. Rev. Lett. 1983, 50, 346. [Google Scholar] [CrossRef]
  44. Eckmann, J.-P.; Ruelle, D. Fundamental limitations for estimating dimensions and Lyapunov exponents in dynamical systems. Phys. D Nonlinear Phenom. 1992, 56, 185–187. [Google Scholar] [CrossRef]
  45. Jeong, J.; Kim, S.Y.; Han, S.-H. Non-linear dynamical analysis of the EEG in Alzheimer’s disease with optimal embedding dimension. Electroencephalogr. Clin. Neurophysiol. 1998, 106, 220–228. [Google Scholar] [CrossRef]
  46. Stam, K.J.; Tavy, D.L.; Jelles, B.; Achtereekte, H.A.; Slaets, J.P.; Keunen, R.W. Non-linear dynamical analysis of multichannel EEG: Clinical applications in dementia and Parkinson’s disease. Brain Topogr. 1994, 7, 141–150. [Google Scholar] [CrossRef] [PubMed]
  47. Smits, F.M.; Porcaro, C.; Cottone, C.; Cancelli, A.; Rossini, P.M.; Tecchio, F. Electroencephalographic fractal dimension in healthy ageing and Alzheimer’s disease. PLoS ONE 2016, 11, e0149587. [Google Scholar] [CrossRef] [PubMed]
  48. Bachmann, M.; Lass, J.; Suhhova, A.; Hinrikus, H. Spectral asymmetry and Higuchi’s fractal dimension measures of depression electroencephalogram. Comput. Math. Methods Med. 2013, 2013, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Kesić, S.; Spasić, S.Z. Application of Higuchi’s fractal dimension from basic to clinical neurophysiology: A review. Comput. Methods Programs Biomed. 2016, 133, 55–70. [Google Scholar] [CrossRef] [PubMed]
  50. Geng, S.J.; Zhou, W.D.; Yao, Q.M.; Ma, Z. Nonlinear analysis of EEG using fractal dimension and approximate entropy. In Advanced Materials Research; Trans Tech Publications Ltd.: Stafa-Zurich, Switzerland; Volume 532, pp. 988–992.
  51. Pincus, S.M. Approximate entropy as a measure of system complexity. Proc. Natl. Acad. Sci. USA 1991, 88, 2297–2301. [Google Scholar] [CrossRef] [Green Version]
  52. Richman, J.S.; Moorman, J.R. Physiological time-series analysis using approximate entropy and sample entropy. Am. J. Physiol. Heart Circ. Physiol. 2000, 278, H2039–H2049. [Google Scholar] [CrossRef] [Green Version]
  53. Costa, M.; Goldberger, A.L.; Peng, C.-K. Multiscale entropy analysis of complex physiologic time series. Phys. Rev. Lett. 2002, 89, 068102. [Google Scholar] [CrossRef] [Green Version]
  54. Chen, W.; Wang, Z.; Xie, H.; Yu, W. Characterization of surface EMG signal based on fuzzy entropy. IEEE Trans. Neural Syst. Rehabil. Eng. 2007, 15, 266–272. [Google Scholar] [CrossRef]
  55. Ross, S.B.; Renyi, A.L. Inhibition of the neuronal uptake of 5-hydroxytryptamine and noradrenaline in rat brain by (Z)- and (E)-3-(4-bromophenyl)-N,N-dimethyl-3-(3-pyridyl) allylamines and their secondary analogues. Neuropharmacology 1977, 16, 57–63. [Google Scholar] [CrossRef]
  56. Powell, G.; Percival, I. A spectral entropy method for distinguishing regular and irregular motion of Hamiltonian systems. J. Phys. A Math. Gen. 1979, 12, 2053. [Google Scholar] [CrossRef]
  57. Tsallis; Constantino. Generalized entropy-based criterion for consistent testing. Phys. Rev. E Stat. Phys. Plasmas Fluids Relat. Interdiscip. Top. 1998, 58, 1442–1445. [Google Scholar]
  58. Feller, W. The asymptotic distribution of the range of sums of independent random variables. Ann. Math. Stat. 1951, 22, 427–432. [Google Scholar] [CrossRef]
  59. Lempel, A.; Ziv, J. On the Complexity of Finite Sequences. IEEE Trans. Inf. Theory 1976, 22, 75–81. [Google Scholar] [CrossRef]
  60. Grassberger, P. Generalized dimensions of strange attractors. Phys. Lett. A 1983, 97, 227–230. [Google Scholar] [CrossRef]
  61. Higuchi, T. Approach to an irregular time series on the basis of the fractal theory. Phys. D Nonlinear Phenom. 1988, 31, 277–283. [Google Scholar] [CrossRef]
  62. Liberati, A.; Altman, D.G.; Tetzlaff, J.; Mulrow, C.; Gotzsche, P.C.; Ioannidis, J.P.A.; Clarke, M.; Devereaux, P.J.; Kleijnen, J.; Moher, D. The PRISMA statement for reporting systematic reviews and meta-analyses of studies that evaluate healthcare interventions: Explanation and elaboration. J. Clin. Epidemiol. 2009, 339, b2700. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Lehmann, M.; Madison, C.; Ghosh, P.M.; Seeley, W.W.; Greicius, M.D.; Gorno-Tempini, M.-L.; Kramer, J.H.; Miller, B.L.; Jagust, W.J.; Rabinovici, G.D. Loss of functional connectivity is greater outside the default mode network in non-familial early-onset ad variants. J. Alzheimer’s Assoc. 2014, 10, 105. [Google Scholar] [CrossRef]
  64. Sala-Llonch, R.; Pena-Gómez, C.; Arenaza-Urquijo, E.M.; Vidal-Pi?eiro, D.; Bargalló, N.; Junqué, C.; Bartrés-Faz, D. Brain connectivity during resting state and subsequent working memory task predicts behavioural performance. Cortex 2012, 48, 1187–1196. [Google Scholar] [CrossRef]
  65. Deco, G.; Jirsa, V.K.; McIntosh, A.R. Resting brains never rest: Computational insights into potential cognitive architectures. Trends Neurosci. 2013, 36, 268–274. [Google Scholar] [CrossRef] [Green Version]
  66. Ganzetti, M.; Mantini, D. Functional connectivity and oscillatory neuronal activity in the resting human brain. Neuroscience 2013, 240, 297–309. [Google Scholar] [CrossRef] [PubMed]
  67. Purdon, P.L.; Pierce, E.T.; Mukamel, E.A.; Prerau, M.J.; Walsh, J.L.; Wong, K.F.K.; Salazar-Gomez, A.F.; Harrell, P.G.; Sampson, A.L.; Cimenser, A. Electroencephalogram signatures of loss and recovery of consciousness from propofol. Proc. Natl. Acad. Sci. USA 2013, 110, E1142–E1151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Liechti, M.D.; Valko, L.; Müller, U.C.; Döhnert, M.; Drechsler, R.; Steinhausen, H.-C.; Brandeis, D. Diagnostic value of resting electroencephalogram in attention-deficit/hyperactivity disorder across the lifespan. Brain Topogr. 2013, 26, 135–151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Hogan, M.J.; Kilmartin, L.; Keane, M.; Collins, P.; Staff, R.T.; Kaiser, J.; Lai, R.; Upton, N. Electrophysiological entropy in younger adults, older controls and older cognitively declined adults. Brain Res. 2012, 1445, 1–10. [Google Scholar] [CrossRef] [Green Version]
  70. Abásolo, D.; Hornero, R.; Espino, P.; Poza, J.; Sánchez, C.I.; Rosa, R.D.L. Analysis of regularity in the EEG background activity of Alzheimer’s disease patients with Approximate Entropy. Clin. Neurophysiol. 2005, 116, 1826–1834. [Google Scholar] [CrossRef] [Green Version]
  71. Abásolo, D.; Escudero, J.; Hornero, R.; Gómez, C.; Espino, P. Approximate entropy and auto mutual information analysis of the electroencephalogram in Alzheimer’s disease patients. Med. Biol. Eng. Comput. 2008, 46, 1019–1028. [Google Scholar] [CrossRef] [Green Version]
  72. Abásolo, D.; Hornero, R.; Espino, P.; Álvarez, D.; Poza, J. Entropy analysis of the EEG background activity in Alzheimer’s disease patients. Physiol. Meas. 2006, 27, 241–253. [Google Scholar] [CrossRef] [Green Version]
  73. Woon, W.L.; Cichocki, A.; Vialatte, F.; Musha, T. Techniques for early detection of Alzheimer’s disease using spontaneous EEG recordings. Physiol. Meas. 2007, 28, 335–347. [Google Scholar] [CrossRef]
  74. Hornero, R.; Abásolo, D.; Escudero, J.; Gómez, C. Nonlinear analysis of electroencephalogram and magnetoencephalogram recordings in patients with Alzheimer’s disease. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2008, 367, 317–336. [Google Scholar] [CrossRef] [Green Version]
  75. Nesma, H.; François, V.; Esteve, G.-J.; Gérard, D.; Vi-Huong, N.-M.; Jean, M.; Kiyoka, K.; D., G.S. Diagnosis of Alzheimer’s disease with Electroencephalography in a differential framework. PLoS ONE 2018, 13, e0193607. [Google Scholar]
  76. Garn, H.; Waser, M.; Deistler, M.; Schmidt, R.; Dal-Bianco, P.; Ransmayr, G.; Zeitlhofer, J.; Schmidt, H.; Seiler, S.; Sanin, G. Quantitative EEG in Alzheimer’s disease: Cognitive state, resting state and association with disease severity. Int. J. Psychophysiol. 2014, 93, 390–397. [Google Scholar] [CrossRef] [PubMed]
  77. Liu, X.; Zhang, C.; Ji, Z.; Ma, Y.; Shang, X.; Zhang, Q.; Zheng, W.; Li, X.; Gao, J.; Wang, R. Multiple characteristics analysis of Alzheimer’s electroencephalogram by power spectral density and Lempel—Ziv complexity. Cogn. Neurodyn. 2016, 10, 121–133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Al-Nuaimi, A.H.H.; Jammeh, E.; Sun, L.; Ifeachor, E. Complexity measures for quantifying changes in electroencephalogram in Alzheimer’s disease. Complexity 2018, 2018, 33. [Google Scholar] [CrossRef]
  79. John, T.N.; Dharmapalan, P.S.; Menon, N.R. Exploration of time–frequency reassignment and homologous inter-hemispheric asymmetry analysis of MCI–AD brain activity. BMC Neurosci. 2019, 20, 38. [Google Scholar]
  80. Reyes-Coronel, C.; Waser, M.; Garn, H.; Deistler, M.; Dal-Bianco, P.; Benke, T.; Ransmayr, G.; Grossegger, D.; Schmidt, R. Predicting rapid cognitive decline in Alzheimer’s disease patients using quantitative EEG markers and neuropsychological test scores. In Proceedings of the 2016 38th Annual International Conference of the IEEE Engineering in Medicine and Biology Society (EMBC), Orlando, FL, USA, 16–20 August 2016; pp. 6078–6081. [Google Scholar]
  81. Al-Qazzaz, N.K.; Ali, S.H.B.M.; Ahmad, S.A.; Islam, M.S.; Escudero, J. Discrimination of stroke-related mild cognitive impairment and vascular dementia using EEG signal analysis. Med. Biol. Eng. Comput. 2018, 56, 137–157. [Google Scholar] [CrossRef]
  82. Deng, B.; Cai, L.; Li, S.; Wang, R.; Yu, H.; Chen, Y.; Wang, J. Multivariate multi-scale weighted permutation entropy analysis of EEG complexity for Alzheimer’s disease. Cogn. Neurodyn. 2017, 11, 217–231. [Google Scholar] [CrossRef]
  83. Mizuno, T.; Takahashi, T.; Cho, R.Y.; Kikuchi, M.; Murata, T.; Takahashi, K.; Wada, Y. Assessment of EEG dynamical complexity in Alzheimer’s disease using multiscale entropy. Clin. Neurophysiol. 2010, 121, 1438–1446. [Google Scholar] [CrossRef] [Green Version]
  84. Chai, X.; Weng, X.; Zhang, Z.; Lu, Y.; Niu, H. Quantitative EEG in Mild Cognitive Impairment and Alzheimer’s Disease by AR-Spectral and Multi-scale Entropy Analysis. In World Congress on Medical Physics and Biomedical Engineering; Springer: Singapore, 2019. [Google Scholar]
  85. Maturana-Candelas, A.; Gómez, C.; Poza, J.; Pinto, N.; Hornero, R. EEG characterization of the Alzheimer’s disease continuum by means of multiscale entropies. Entropy 2019, 21, 544. [Google Scholar] [CrossRef] [Green Version]
  86. Yang, A.C.; Huang, C.-C.; Yeh, H.-L.; Liu, M.-E.; Hong, C.-J.; Tu, P.-C.; Chen, J.-F.; Huang, N.E.; Peng, C.-K.; Lin, C.-P. Complexity of spontaneous BOLD activity in default mode network is correlated with cognitive function in normal male elderly: A multiscale entropy analysis. Neurobiol. Aging 2013, 34, 428–438. [Google Scholar] [CrossRef]
  87. Al-Qazzaz, N.K.; Ali, S.; Islam, M.S.; Ahmad, S.A.; Escudero, J. EEG markers for early detection and characterization of vascular dementia during working memory tasks. In Proceedings of the 2016 IEEE EMBS Conference on Biomedical Engineering and Sciences (IECBES), Kuala Lumpur, Malaysia, 4–8 December 2016; pp. 347–351. [Google Scholar]
  88. Sharma, N.; Kolekar, M.; Jha, K.; Kumar, Y. EEG and Cognitive Biomarkers Based Mild Cognitive Impairment Diagnosis. IRBM 2019, 40, 113–121. [Google Scholar] [CrossRef]
  89. Ruiz-Gómez, S.; Gómez, C.; Poza, J.; Gutiérrez-Tobal, G.; Tola-Arribas, M.; Cano, M.; Hornero, R. Automated multiclass classification of spontaneous EEG activity in Alzheimer’s disease and mild cognitive impairment. Entropy 2018, 20, 35. [Google Scholar] [CrossRef] [Green Version]
  90. De Bock, T.J.; Das, S.; Mohsin, M.; Munro, N.B.; Hively, L.M.; Jiang, Y.; Smith, C.D.; Wekstein, D.R.; Jicha, G.A.; Lawson, A. Early detection of Alzheimer’s disease using nonlinear analysis of EEG via Tsallis entropy. In Proceedings of the 2010 Biomedical Sciences and Engineering Conference, Oak Ridge, TN, USA, 5–6 May 2010; pp. 1–4. [Google Scholar]
  91. Coronel, C.; Garn, H.; Waser, M.; Deistler, M.; Benke, T.; Dal-Bianco, P.; Ransmayr, G.; Seiler, S.; Grossegger, D.; Schmidt, R. Quantitative EEG markers of entropy and auto mutual information in relation to MMSE scores of probable Alzheimer’s disease patients. Entropy 2017, 19, 130. [Google Scholar] [CrossRef] [Green Version]
  92. Dauwels, J.; Vialatte, F.; Cichocki, A. Diagnosis of Alzheimer’s disease from EEG signals: Where are we standing? Curr. Alzheimer Res. 2010, 7, 487–505. [Google Scholar] [CrossRef] [Green Version]
  93. Hirschmann, J.; Hartmann, C.J.; Butz, M.; Hoogenboom, N.; Özkurt, T.E.; Elben, S.; Vesper, J.; Wojtecki, L.; Schnitzler, A. A direct relationship between oscillatory subthalamic nucleus—Cortex coupling and rest tremor in Parkinson’s disease. Brain 2013, 136, 3659–3670. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Jelles, B.; Scheltens, P.; Van der Flier, W.; Jonkman, E.; da Silva, F.L.; Stam, C. Global dynamical analysis of the EEG in Alzheimer’s disease: Frequency-specific changes of functional interactions. Clin. Neurophysiol. 2008, 119, 837–841. [Google Scholar] [CrossRef] [PubMed]
  95. Hutcheon, B.; Yarom, Y. Resonance, oscillation and the intrinsic frequency preferences of neurons. Trends Neurosci. 2000, 23, 216–222. [Google Scholar] [CrossRef]
  96. Schnitzler, A.; Gross, J. Normal and pathological oscillatory communication in the brain. Nat. Rev. Neurosci. 2005, 6, 285. [Google Scholar] [CrossRef]
  97. Von Stein, A.; Sarnthein, J. Different frequencies for different scales of cortical integration: From local gamma to long range alpha/theta synchronization. Int. J. Psychophysiol. 2000, 38, 301–313. [Google Scholar] [CrossRef]
  98. Vakorin, V.A.; Lippé, S.; McIntosh, A.R. Variability of brain signals processed locally transforms into higher connectivity with brain development. J. Neurosci. 2011, 31, 6405–6413. [Google Scholar] [CrossRef] [Green Version]
  99. Jeong, J. EEG dynamics in patients with Alzheimer’s disease. Clin. Neurophysiol. 2004, 115, 1490–1505. [Google Scholar] [CrossRef]
  100. Henderson, G.; Ifeachor, E.; Wimalaratna, H.; Allen, E.; Hudson, N. Prospects for routine detection of dementia using the fractal dimension of the human electroencephalogram. IEE Proc. Sci. Meas. Technol. 2000, 147, 321–326. [Google Scholar] [CrossRef]
  101. Amezquita-Sanchez, J.P.; Mammone, N.; Morabito, F.C.; Marino, S.; Adeli, H. A novel methodology for automated differential diagnosis of mild cognitive impairment and the Alzheimer’s disease using EEG signals. J. Neurosci. Methods 2019, 322, 88–95. [Google Scholar] [CrossRef] [PubMed]
  102. Al-nuaimi, A.H.; Jammeh, E.; Sun, L.; Ifeachor, E. Higuchi fractal dimension of the electroencephalogram as a biomarker for early detection of Alzheimer’s disease. In Proceedings of the 2017 39th Annual International Conference of the IEEE Engineering in Medicine and Biology Society (EMBC), Seogwipo, Korea, 11–15 July 2017; pp. 2320–2324. [Google Scholar]
  103. Mcbride, J.C.; Zhao, X.; Munro, N.B.; Smith, C.D.; Jicha, G.A.; Hively, L.; Broster, L.S.; Schmitt, F.A.; Kryscio, R.J.; Jiang, Y. Spectral and complexity analysis of scalp EEG characteristics for mild cognitive impairment and early Alzheimer’s disease. Comput. Methods Programs Biomed. 2014, 114, 153–163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Fan, M.; Yang, A.C.; Fuh, J.-L.; Chou, C.-A. Topological Pattern Recognition of Severe Alzheimer’s Disease via Regularized Supervised Learning of EEG Complexity. Front. Neurosci. 2018, 12, 685. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Simons, S.; Espino, P.; Abásolo, D. Fuzzy entropy analysis of the electroencephalogram in patients with Alzheimer’s disease: Is the method superior to sample entropy? Entropy 2018, 20, 21. [Google Scholar] [CrossRef] [Green Version]
  106. Cuffin, B.N.; Cohen, D. Comparison of the magnetoencephalogram and electroencephalogram. Electroencephalogr. Clin. Neurophysiol. 1979, 47, 132–146. [Google Scholar] [CrossRef]
  107. Jirsa, V.K.; Friedrich, R.; Haken, H. Reconstruction of the spatio-temporal dynamics of a human magnetoencephalogram. Phys. D Nonlinear Phenom. 1995, 89, 100–122. [Google Scholar] [CrossRef]
  108. Gómez, C.; Martínez-Zarzuela, M.; Poza, J.; Díaz-Pernas, F.J.; Fernández, A.; Hornero, R. Synchrony analysis of spontaneous MEG activity in Alzheimer’s disease patients. In Proceedings of the 2012 Annual International Conference of the IEEE Engineering in Medicine and Biology Society, San Diego, CA, USA, 28 August–1 September 2012; pp. 6188–6191. [Google Scholar]
  109. Gómez, C.; Poza, J.; Monge, J.; Fernández, A.; Hornero, R. Analysis of magnetoencephalography recordings from Alzheimer’s disease patients using embedding entropies. In Proceedings of the 2014 36th Annual International Conference of the IEEE Engineering in Medicine and Biology Society, Chicago, IL, USA, 26–30 August 2014; pp. 702–705. [Google Scholar]
  110. Amezquita-Sanchez, J.P.; Adeli, A.; Adeli, H. A new methodology for automated diagnosis of mild cognitive impairment (MCI) using magnetoencephalography (MEG). Behav. Brain Res. 2016, 305, 174–180. [Google Scholar] [CrossRef]
  111. Azami, H.; Rostaghi, M.; Fernández, A.; Escudero, J. Dispersion entropy for the analysis of resting-state MEG regularity in Alzheimer’s disease. In Proceedings of the 2016 38th Annual International Conference of the IEEE Engineering in Medicine and Biology Society (EMBC), Orlando, FL, USA, 16–20 August 2016; pp. 6417–6420. [Google Scholar]
  112. Gómez, C.; Hornero, R.; Abásolo, D.; Fernandez, A.; Escudero, J. Analysis of MEG recordings from Alzheimer’s disease patients with sample and multiscale entropies. In Proceedings of the 2007 29th Annual International Conference of the IEEE Engineering in Medicine and Biology Society, Lyon, France, 22–26 August 2007; pp. 6183–6186. [Google Scholar]
  113. Azami, H.; Kinney-Lang, E.; Ebied, A.; Fernández, A.; Escudero, J. Multiscale dispersion entropy for the regional analysis of resting-state magnetoencephalogram complexity in Alzheimer’s disease. In Proceedings of the 2017 39th Annual International Conference of the IEEE Engineering in Medicine and Biology Society (EMBC), Seogwipo, Korea, 11–15 July 2017; pp. 3182–3185. [Google Scholar]
  114. Engels, M.; van der Flier, W.; Stam, C.; Hillebrand, A.; Scheltens, P.; van Straaten, E. Alzheimer’s disease: The state of the art in resting-state magnetoencephalography. Clin. Neurophysiol. 2017, 128, 1426–1437. [Google Scholar] [CrossRef]
  115. Poza, J.; Hornero, R.; Abásolo, D.; Fernandez, A.; Escudero, J. Analysis of spontaneous MEG activity in patients with Alzheimer’s disease using spectral entropies. In Proceedings of the 2007 29th Annual International Conference of the IEEE Engineering in Medicine and Biology Society, Lyon, France, 22–26 August 2007; pp. 6179–6182. [Google Scholar]
  116. Poza, J.; Hornero, R.; Abásolo, D.; Fernández, A.; Mayo, A. Evaluation of spectral ratio measures from spontaneous MEG recordings in patients with Alzheimer’s disease. Comput. Methods Programs Biomed. 2008, 90, 137–147. [Google Scholar] [CrossRef] [Green Version]
  117. BruñA, R.; Poza, J.; Gómez, C.; García, M.; Fernández, A.; Hornero, R. Analysis of spontaneous MEG activity in mild cognitive impairment and Alzheimer’s disease using spectral entropies and statistical complexity measures. J. Neural Eng. 2012, 9, 036007. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Gómez, C.; Mediavilla, Á.; Hornero, R.; Abásolo, D.; Fernández, A. Use of the Higuchi’s fractal dimension for the analysis of MEG recordings from Alzheimer’s disease patients. Med. Eng. Phys. 2009, 31, 306–313. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Shumbayawonda, E.; López-Sanz, D.; Bruña, R.; Serrano, N.; Fernández, A.; Maestú, F.; Abasolo, D. Complexity changes in preclinical Alzheimer’s disease: An MEG study of subjective cognitive decline and mild cognitive impairment. Clin. Neurophysiol. 2020, 131, 437–445. [Google Scholar] [CrossRef] [PubMed]
  120. Fernández, A.; Hornero, R.; Gómez, C.; Turrero, A.; Ortiz, T. Complexity Analysis of Spontaneous Brain Activity in Alzheimer Disease and Mild Cognitive Impairment an MEG Study. Alzheimer Dis. Assoc. Disord. 2010, 24, 182–189. [Google Scholar] [CrossRef] [PubMed]
  121. Azami, H.; Escudero, J.; Fernández, A. Refined composite multivariate multiscale entropy based on variance for analysis of resting-state magnetoencephalograms in Alzheimer’s disease. In Proceedings of the 2016 International Conference for Students on Applied Engineering (ICSAE), Newcastle upon Tyne, UK, 20–21 October 2016; pp. 413–418. [Google Scholar]
  122. Escudero, J.; Acar, E.; Fernández, A.; Bro, R. Multiscale entropy analysis of resting-state magnetoencephalogram with tensor factorisations in Alzheimer’s disease. Brain Res. Bull. 2015, 119, 136–144. [Google Scholar] [CrossRef]
  123. Poza, J.; Gómez, C.; Bachiller, A.; Hornero, R. Spectral and Non-Linear Analyses of Spontaneous Magnetoencephalographic Activity in Alzheimer’s Disease. J. Healthc. Eng. 2012, 3, 299–322. [Google Scholar] [CrossRef] [Green Version]
  124. Gómez, C.; Hornero, R. Entropy and complexity analyses in Alzheimer’s disease: An MEG study. Open Biomed. Eng. J. 2010, 4, 223. [Google Scholar] [CrossRef] [Green Version]
  125. Poza, J.; Hornero, R.; Escudero, J.; Fernández, A.; Sánchez, C.I. Regional analysis of spontaneous MEG rhythms in patients with Alzheimer’s disease using spectral entropies. Ann. Biomed. Eng. 2008, 36, 141–152. [Google Scholar] [CrossRef]
  126. Hornero, R.; Escudero, J.; Fernández, A.; Poza, J.; Gómez, C. Spectral and nonlinear analyses of MEG background activity in patients with Alzheimer’s disease. IEEE Trans. Biomed. Eng. 2008, 55, 1658–1665. [Google Scholar] [CrossRef] [Green Version]
  127. Filippi, M.; Rocca, M.A. Functional Magnetic Resonance Imaging; Sinauer Associates: Sunderland, MA, USA, 2002; Volume 1. [Google Scholar]
  128. Boynton, G.M.; Engel, S.A.; Glover, G.H.; Heeger, D.J. Linear Systems Analysis of Functional Magnetic Resonance Imaging in Human V1. J. Neurosci. 1996, 16, 4207–4221. [Google Scholar] [CrossRef]
  129. Bunce, S.C.; Izzetoglu, M.; Izzetoglu, K.; Onaral, B.; Pourrezaei, K. Functional near-infrared spectroscopy. Eng. Med. Biol. Mag. IEEE 2006, 25, 54–62. [Google Scholar] [CrossRef] [PubMed]
  130. Ferrari, M.; Quaresima, V. A brief review on the history of human functional near-infrared spectroscopy (fNIRS) development and fields of application. Neuroimage 2012, 63, 921–925. [Google Scholar] [CrossRef] [PubMed]
  131. Maxim, V.; Şendur, L.; Fadili, J.; Suckling, J.; Gould, R.; Howard, R.; Bullmore, E. Fractional Gaussian noise, functional MRI and Alzheimer’s disease. Neuroimage 2005, 25, 141–158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Liu, F.; Guo, W.; Liu, L.; Long, Z.; Ma, C.; Xue, Z.; Wang, Y.; Li, J.; Hu, M.; Zhang, J. Abnormal amplitude low-frequency oscillations in medication-naive, first-episode patients with major depressive disorder: A resting-state fMRI study. J. Affect. Disord. 2013, 146, 401–406. [Google Scholar] [CrossRef] [PubMed]
  133. Niu, Y.; Wang, B.; Zhou, M.; Xue, J.; Shapour, H.; Cao, R.; Cui, X.; Wu, J.; Xiang, J. Dynamic Complexity of Spontaneous Bold Activity in Alzheimer’s Disease and Mild Cognitive Impairment Using Multiscale Entropy analysis. Front. Neurosci. 2018, 12, 677. [Google Scholar] [CrossRef] [PubMed]
  134. Perpetuini, D.; Bucco, R.; Zito, M.; Merla, A. Study of memory deficit in Alzheimer’s disease by means of complexity analysis of fNIRS signal. Neurophotonics 2017, 5, 011010. [Google Scholar] [CrossRef]
  135. Coleman, P.; Federoff, H.; Kurlan, R. A focus on the synapse for neuroprotection in Alzheimer disease and other dementias. Neurology 2004, 63, 1155–1162. [Google Scholar] [CrossRef]
  136. Babiloni, C.; Del Percio, C.; Bordet, R.; Bourriez, J.-L.; Bentivoglio, M.; Payoux, P.; Derambure, P.; Dix, S.; Infarinato, F.; Lizio, R. Effects of acetylcholinesterase inhibitors and memantine on resting-state electroencephalographic rhythms in Alzheimer’s disease patients. Clin. Neurophysiol. 2013, 124, 837–850. [Google Scholar] [CrossRef]
  137. Hasegawa, M.; Morishima-Kawashima, M.; Takio, K.; Suzuki, M.; Titani, K.A.; Ihara, Y. Protein sequence and mass spectrometric analyses of tau in the Alzheimer’s disease brain. J. Biol. Chem. 1992, 267, 17047–17054. [Google Scholar]
  138. Bondi, M.W.; Houston, W.S.; Eyler, L.T.; Brown, G.G. fMRI evidence of compensatory mechanisms in older adults at genetic risk for Alzheimer disease. Neurology 2005, 64, 501–508. [Google Scholar] [CrossRef] [Green Version]
  139. Beaunieux, H.; Eustache, F.; Busson, P.; De La Sayette, V.; Viader, F.; Desgranges, B. Cognitive procedural learning in early Alzheimer’s disease: Impaired processes and compensatory mechanisms. J. Neuropsychol. 2012, 6, 31–42. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Selection diagram, including three stages: identification, screening, and inclusion. This process led from 382 initial studies to 126 final studies.
Figure 1. Selection diagram, including three stages: identification, screening, and inclusion. This process led from 382 initial studies to 126 final studies.
Entropy 22 00239 g001
Figure 2. (A) Three modes of data categorization reviewed in the study. The inner circle shows the different brain imaging modalities, while the outer circle shows specific complexity analysis methods. (B) Trends in the number of included studies using the different brain imaging techniques versus date.
Figure 2. (A) Three modes of data categorization reviewed in the study. The inner circle shows the different brain imaging modalities, while the outer circle shows specific complexity analysis methods. (B) Trends in the number of included studies using the different brain imaging techniques versus date.
Entropy 22 00239 g002
Figure 3. Comparative values of entropy from five regions across the brain in Alzheimer’s disease (AD), mild cognitive impairment (MCI), and control subjects [18,81].
Figure 3. Comparative values of entropy from five regions across the brain in Alzheimer’s disease (AD), mild cognitive impairment (MCI), and control subjects [18,81].
Entropy 22 00239 g003
Figure 4. Entropy at different scales in different regions of the brain [82,84].
Figure 4. Entropy at different scales in different regions of the brain [82,84].
Entropy 22 00239 g004
Figure 5. Entropy in different frequency bands across the brain in AD, MCI, and control subjects [81].
Figure 5. Entropy in different frequency bands across the brain in AD, MCI, and control subjects [81].
Entropy 22 00239 g005
Figure 6. Entropy from different brain regions [48,113].
Figure 6. Entropy from different brain regions [48,113].
Entropy 22 00239 g006
Figure 7. Brain regions with significant differences between groups [15].
Figure 7. Brain regions with significant differences between groups [15].
Entropy 22 00239 g007
Figure 8. Brain regions with significant differences between groups on scale factors two, four, five, and six [133].
Figure 8. Brain regions with significant differences between groups on scale factors two, four, five, and six [133].
Entropy 22 00239 g008
Table 1. Summary of widely used complexity analysis methods.
Table 1. Summary of widely used complexity analysis methods.
Complexity IndicesAbbreviations YearDescription
Time domain entropyApproximate entropyApEnPincus (1991) [51]Needs only a small dataset and is effective for discriminating the signal from random signals. A higher value indicates more irregularity.
Sample entropySampEnRichman (2000) [52]The exact value of the negative average natural logarithm of the conditional probability. A higher value indicates less predictable signals.
Permutation entropyPeEnBandt (2002) [27]Only considers the grades of the samples but not their metrics. A higher value indicates a more irregular signal.
Multiscale entropyMEnCosta (2005) [53]Can be observed at multiple different scales of signal change.
Fuzzy entropyFuzzyEnChen (2007) [54]Provides a mechanism for measuring the degree to which a pattern belongs to a given class.
Frequency domain entropyRenyi entropyReEnRenyi (1977) [55]Forms the basis of the concept of generalized dimensionality. If the Renyi entropy is high, the signal has high complexity.
Spectral entropySpecEnPowell (1979) [56]Predictability according to an analysis of the spectral content of a signal. A high value indicates a more irregular and less predictable signal.
Tsallis entropyTsEnTsallis (1998) [57]Explores the properties of a probability distribution from a new mathematical framework.
OthersHurst exponentHEHurst (1951) [58]Used to measure the long-term memory and fractal dimension of a time series.
Lempel-Ziv complexityLZCLempel (1976) [59]Reconstructs the original time series into a binary sequence. A high value indicates a high variation in the binary signal.
Correlation dimensionD2Grassberger (1983) [60]The number of independent variables needed to describe the time series dynamics after the time series is transferred to chaos theory-based state space.
Fractal dimension FDHiguchi (1988) [61]It complements the chaos theory of the dynamic system, showing the similarity with the whole.
Table 2. Sensitivity, specificity, and accuracy in differentiating among AD, MCI, and normal control (NC) subjects were found with different nonlinear methods used in the electroencephalogram (EEG) database (NR represents that the paper does not give this value accurately).
Table 2. Sensitivity, specificity, and accuracy in differentiating among AD, MCI, and normal control (NC) subjects were found with different nonlinear methods used in the electroencephalogram (EEG) database (NR represents that the paper does not give this value accurately).
Research MethodClassSensitivitySpecificity Accuracy AUC
Sharma et al. (2019) [88]SpecEn + FDNC vs. MCI86%81% 84.1%NR
MCI vs. AD83%63% 73.4% NR
NC vs. AD82% 82% 82% NR
Chai et al. (2019) [84]MSENC vs. MCINRNRNR73%
NC vs. ADNRNRNR81%
Fan et al. (2018) [104]MSENC vs. AD88.71% 69.09% 79.49%83%
Houmani et al. (2018) [105] EpEn (epoch-based entropy)SCI vs. AD87.8%100% 91.6%NR
Simons et al. (2018) [75]ApEnNC vs. AD90.91%63.64%77.27%NR
SampEn 90.91%63.64%77.27%NR
Al-Nuaimi (2018) [78]ApEnNC vs. AD72.73%81.82%77.27%85.95%
SampEn81.82%72.73%77.27%85.95%
LZC81.82%81.82%81.82%89.26%
FuzzyEn81.82%90.91%86.36%86.78%
MSE90.91%90.91%90.91%93.39%
AMI100%81.82%90.91%93.39%
HFD 66.67%100%80%NR
Al-Qazzaz (2016) [87]TsEnNC vs. AD85.71%84.62%85%NR
LZC100%92.31%95%NR
Liu et al. (2016) [77] LZCNC vs. AD80.0%78.1%78.5%89.21%
MS_LZC (multiscale_LZC)86.8%84.3%85.7%91.12%
Table 3. Sensitivity, specificity, and accuracy in differentiating among AD, MCI, and NC subjects were found with different nonlinear methods used in the magnetoencephalogram (MEG) database (NR represents that the paper does not give this value accurately).
Table 3. Sensitivity, specificity, and accuracy in differentiating among AD, MCI, and NC subjects were found with different nonlinear methods used in the magnetoencephalogram (MEG) database (NR represents that the paper does not give this value accurately).
Research MethodClassSensitivitySpecificity Accuracy AUC
Azami et al. (2016) [121]MFE (multiscale fuzzy entropy)NC vs. ADNRNR78.22%NR
Juan P. et al. (2016) [110] PeEn MCI vs. ADNRNR98.4%NR
Escuderoa et al. (2015) [122]MSENC vs. AD94.4%46.2%NR67%
Gómez et al. (2014) [109]SampEnNC vs. AD80.00% 61.90% 70.73%NR
LZC 80.00% 76.19% 78.05%NR
Bruña et al. (2012) [117]ShEnNC vs. ADNRNR69.4%79.0%
NC vs. MCINRNR65.9%64.1%
MCI vs. ADNRNR64.8%69.1%
TsEnNC vs. ADNRNR75.8%85.6%
NC vs. MCINRNR61.4%60.7%
MCI vs. ADNRNR66.7%75.6%
ReEnNC vs. ADNRNR83.9%89.0%
NC vs. MCINRNR63.6%65.2%
MCI vs. ADNRNR72.2%78.5%
Poza et al. (2012) [123]SampEnNC vs. AD88.9%57.7%75.8%80.6%
Gómez et al. (2010) [124]SampEn 77.78%50.00%66.13%71.26%
ApEn75.00%53.85%66.13%73.82%
HFD72.22%73.08%72.58%79.11%
LZC80.56%61.54%72.58%78.63%
ShEn91.67%57.69%77.42%79.27%
Hornero et al. [74]ApEnNC vs. AD75.0%66.7%70.7%NR
AMI75.0%90.5%82.9%NR
LZC85.0%85.7% 85.4%NR
Gómez et al. (2007) [112]SampEnNC vs. AD80%76.2%NR84%
MSE75%100%NR87.8%
Poza et al. (2008) [125]ShEnNC vs. AD85.0%81.0%82.9%NR
ReEn90.0%85.7%87.8%NR
Hornero et al. (2008) [126]ApEnNC vs. AD50.0%52.4%51.2%NR
LZC65.0%76.2%70.7%NR
SpecEn70.0%76.2%73.2%NR

Share and Cite

MDPI and ACS Style

Sun, J.; Wang, B.; Niu, Y.; Tan, Y.; Fan, C.; Zhang, N.; Xue, J.; Wei, J.; Xiang, J. Complexity Analysis of EEG, MEG, and fMRI in Mild Cognitive Impairment and Alzheimer’s Disease: A Review. Entropy 2020, 22, 239. https://doi.org/10.3390/e22020239

AMA Style

Sun J, Wang B, Niu Y, Tan Y, Fan C, Zhang N, Xue J, Wei J, Xiang J. Complexity Analysis of EEG, MEG, and fMRI in Mild Cognitive Impairment and Alzheimer’s Disease: A Review. Entropy. 2020; 22(2):239. https://doi.org/10.3390/e22020239

Chicago/Turabian Style

Sun, Jie, Bin Wang, Yan Niu, Yuan Tan, Chanjuan Fan, Nan Zhang, Jiayue Xue, Jing Wei, and Jie Xiang. 2020. "Complexity Analysis of EEG, MEG, and fMRI in Mild Cognitive Impairment and Alzheimer’s Disease: A Review" Entropy 22, no. 2: 239. https://doi.org/10.3390/e22020239

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