Skip to main content

Multiclass EEG motor-imagery classification with sub-band common spatial patterns

Abstract

Electroencephalogram (EEG) signal classification plays an important role to facilitate physically impaired patients by providing brain-computer interface (BCI)-controlled devices. However, practical applications of BCI make it difficult to decode motor imagery-based brain signals for multiclass classification due to their non-stationary nature. In this study, we aim to improve multiclass classification accuracy for motor imagery movement using sub-band common spatial patterns with sequential feature selection (SBCSP-SBFS) method. Filter bank having bandpass filters of different overlapped frequency cutoffs is applied to suppress the noise signals from raw EEG signals. The output of these sub-band filters is sent for feature extraction by applying common spatial pattern (CSP) and linear discriminant analysis (LDA). As all of the extracted features are not necessary for classification therefore, selection of optimal features is done by passing the extracted features to sequential backward floating selection (SBFS) technique. Three different classifiers were then trained on these optimal features, i.e., support vector machine (SVM), Naïve-Bayesian Parzen-Window (NBPW), and k-Nearest Neighbor (KNN). Results are evaluated on two datasets, i.e., Emotiv Epoc and wet gel electrodes for three classes, i.e., right-hand motor imagery, left hand motor imagery, and rest state. The proposed model yields a maximum accuracy of 60.61% in case of Emotiv Epoc headset and 86.50% for wet gel electrodes. The computed accuracy shows an increase of 7% as compared to previously implemented multiclass EEG classification.

1 Introduction

Brain signals, produced as a result of interneuronal brain activity, can be measured using neuroimaging technique known as an electroencephalogram (EEG) [1]. Brain-computer interface (BCI) translates these electrical signals into control commands thus providing a communication pathway between the brain and outside world devices such as BCI wheelchair, neuro-gaming, and prosthetics. BCI facilitates physically impaired patients by capturing their brain signals which after preprocessing, major features are extracted and then on the basis of classified results, outside world devices are controlled [2]. There are two most common approaches to BCI based on measuring methods of EEG signals, i.e., invasive and noninvasive BCIs. The invasive technique is more complex and dangerous as electrodes are directly implanted into the brain by neurosurgery, as a result, firing of neurons can be read without any external interference which detects high-quality signal [3]. Conversely, in non-invasive BCI, the electrodes are placed on the human scalp. Although, these signals have reduced spatial resolution and have more noise contents but due to its easiness, safety, and cost effectiveness, it is used widely [4]. For non-invasive BCI, EEG data can be acquired in two ways, i.e., by using wet gel electrodes and dry electrodes. Contact impedance measured between the skin and an electrode represents the quality of EEG signal. High contact impedance can result in increased noise, as a result, gives poor signal quality. This impedance can be reduced by applying conductive gel paste but require extensive skin preparation which involves extensive skin abrasion and dead cells removal from the skin surface [5].

Acquisition of EEG signals by using non-invasive techniques has an influence of external noise; therefore, the acquired signal is contaminated with artifacts such as signals produced as a result of muscle movement, cable noise, and environment noise [6]. These artifacts make the process of BCI signal translation very complex because pure EEG signal has low amplitude and frequency components as compared to those of artifacts. So, after EEG signal acquisition, first of all, the noise is suppressed from the signal using different spatial filters [7]; this process of noise suppression is known as preprocessing. This signal can directly be fed to the classifier but this will make the process of classifier training very time-consuming. Moreover, EEG signal has a low spatial resolution, low signal-to-noise ratio (SNR) and its measurement mainly attribute to the volume conduction, which signifies the electrical field of the brain that is conducted from the source to the scalp [8]. To measure this low SNR and spatial resolution, common spatial pattern (CSP) was proposed to efficiently extract spatial features for motor imagery brain signals. This method computes the maximum ration of variances between different classes of data by applying filters on it. Research shows that common spatial patterns (CSP) is proved an efficient and widely used technique to extract feature from EEG signal [9]. As CSP is mainly carried out on the complete data while whole EEG data does not contain useful information; therefore it ignores the importance of features. So, to overcome this problem, EEG signal is divided into different sub-bands to extract information from different portions of the signal and then the selection of features from extracted features on the basis of different information produces better results [10,11,12,13,14].

The paper is organized as follows: the “Related work” section presents the detailed literature review, the “Methodology” section provides the details about proposed methodology while the “Results and discussion” section describes the findings of this paper. Finally, the last section presents the conclusion.

1.1 Related work

Recently, various machine learning (ML) approaches for BCI have been developed to circumvent the complexity of brain signals. These different methods have shown remarkable results for binary class EEG classification [10,11,12,13,14] while for better and more control commands of real-time applications such as BCI controlled wheelchair, neuro-gaming, and prosthetics, binary class is not sufficient; therefore, a requirement for multiclass classification arises. Nicolas et al. implemented adaptive generalization method for classification of multiclass motor imagery (MI)-based EEG signals, by the selection of optimal extracted subband CSP features using mutual-information best individual features (MIBIF) [15]. Classification is done by using stacked regularized linear discriminant analysis (SRLDA) [16]. The resulted accuracy has shown better results for binary class rather than for multiclass, i.e., 85% and 74% respectively. Shiratori et al. [17] included “rest state” as a third class along with other two classes for MI movement of the left and right hand. They extracted CSP features from the EEG signal is divided into subbands. Support vector machine (SVM) [18] was used as a classifier to train the system on selected features based on mutual information. Although the implemented methodology has shown 88.7% results for classification of EEG signals produced as a result of finger tapping, it did not show good results for multiclass MI data, i.e., 56.7%.

A lot of research has been conducted for the classification of multiclass EEG signals of same limb movements. Yong and Menon [19] applied a bandpass filter to preprocess raw data while both CSP and Filter bank common spatial pattern (FBCSP) [11] were used to extract features from multiclass EEG signal acquired from the same limb. This approach showed better results for data acquired from the same limb than data from different limbs but the overall accuracy of the system was reported 60.7%. Shiman et al. [20] investigated that MI-based brain signals produced as a result of the imagination of movement from the same limb produce more results than the signals acquired from different limbs. They applied FBCSP having unique frequency cutoffs, i.e., (7 to 15 Hz, 15 to 25 Hz, 25 to 30 Hz) on the acquired same limb EEG data for feature extraction. Then, linear discriminant analysis (LDA) [21] is trained for classification on the basis of extracted features. The results show an accuracy of 69.1% and 62.75% for 3 and 4 class MI-based brain signal respectively.

Apart from feature extraction, classifier improvements have also been implemented to enhance the overall accuracy of the system. She et al. [22] implemented a multiclass posterior probability classification technique for twin SVM through ranking continuous output and pairwise coupling. Platt’s estimating technique and the ranking continuous output techniques were used for two class posterior probability approximation. Then, each pair of class probabilities were combined by using pairwise coupling for multiclass probabilistic outputs. This technique has not proven efficient for classification of multiclass MI-based EEG signal. Further, Gao et al. [23] tested the effectiveness of Adaboost extreme learning machine (AdaboostELM) for classification of features extracted by using Kolmogorov complexity (Kc). They followed the approach of Adaboost model, which states that by calling a base lerning algorithm rapidly, the performance of that algorithm can be enhanced. So, the features were extracted from muthe lticlass MI-based brain signal by using Kc after removing EOG artifacts from the signal using a single wide band filter, then classified by using AdaboostELM. Although, the results showed improved results for adboost approach on the classification accuracy, but these results were calculated by using one versus rest approach which is considered as a binary class classification technique [24]. Meisheri et al. [25] produced better CSP features after identification and removal of artifacts using joint approximate diagonalization (JAD) [26] in preprocessing of the MI-based EEG data. Fast Frobenius Diagonalization (FFDIAG) [27] was applied on the EEG signal for obtaining spatial filters by JAD then CSP is applied on the resulted signal for feature extraction. Self-regulated interval type 2 neuro fuzzy inference system (SRIT2NFIS) is used for classification of these extrated features. The implemented technique have shown good results on the produced algorithm, i.e., maximum of 74.65% accuracy for a single subject. Table 1 shows a brief comparison of previous researches for classification of multiclass MI based brain signals.

Table 1 Comparison of related work in literature

The related work for multiclass EEG classification shows that most of the results were evaluated by using one versus rest approach which is considered as a binary class classification. Therefore, for run-time control of BCI-controlled devices, multiclass classification is required [24]. This paper aims to implement multiclass classification to improve the results of motor imagery brain signals by using sub-band common spatial patterns with sequential backward floating selection (SBCSP-SBFS). Feature extraction and then selection of optimal features on the basis of the final performance of the classifier is the principle of this approach. The motor imagery-based EEG signal is decomposed into different subbands using a filter bank which contains filters of different frequency cutoffs. Once the signal is decomposed, then CSP is applied on it to extract sub-band CSP features, these features are fed into LDA for eigenvector computation. All of the features are not optimal for getting better accuracy from classifier; therefore, sequential backward floating selection (SBFS) technique is used to select ‘k’ optimal features from ‘n’ extracted features. Different classifiers SVM, k-nearest neighbor (KNN) [29] and naïve Bayesian Parzen window (NBPW), are trained and tested on these ‘k’ optimal features for evaluation of the proposed model.

2 Methodology

In this section, the proposed methodology implemented to classify multiclass motor imagery EEG signal is presented. The complete block diagram is shown in Fig. 1. In the first stage, preprocessing is done using a filter bank by decomposing EEG signal into sub-bands. In the second stage, CSP is applied to every bandpass filtered signal. Then LDA is applied to these features to acquire scores which represent the classification capability of every frequency band. The third stage involves feature extraction in which SBFS is implemented to pick out the most effective features from all of the sub-band features resulted via LDA. This feature selection is carried out on the basis of best accuracy of the classifier for specific combinations from extracted sub-band features. Finally, three classifiers, i.e, SVM, KNN, and NMPW are applied for training and testing of the proposed system. EEG processing blocks of the proposed model are explained in the following section.

Fig. 1
figure 1

Block diagram of SBCSP-SBFS. Step by step processing of EEG signal by applying different machine learning algorithms is shown in this picture. Each main block represents a specific step, i.e., preprocessing, feature extraction, and classification

2.1 Bandpass filtering by overlapping filter bank

For preprocessing, filter bank having each bandpass filter is implemented instead of applying one wide bandpass filter of 8–30 Hz on raw EEG signal. The proposed filter bank includes 12 filters of two distinct categories having unique frequency edges between 6 and 32 Hz. Every category in filter bank has 6 filters, and cutoffs of the filters in every category are different in a manner that there is an overlapping of 2 Hz between frequencies of both categories. One filter bank category has a gap of 4 Hz in a range of 8 to 32 Hz, at the same time another category has also a gap of 4 Hz ranging from 6 to 30 Hz. The overlapping of 2 Hz helps to avoid the possible information loss between two consecutive bands.

2.2 Feature extraction by common spatial pattern and linear discriminant analysis

Common spatial pattern (CSP) is applied after filter bank in order to extract the features which are utilized for classification of multiclass data. It yields excellent outcomes by means of extracting considerable EEG features for MI-based signal. The goal of CSP is to grasp the spatial filters for maximizing the converted data variance ration among distinctive classes in EEG data. Basically, CSP is used for a binary class problem, but in the proposed approach, it is used for the multiclass problem by using one-vs-rest (OVR) algorithm.

Multiclass EEG data is generally assumed to be centered without generality loss, i.e., Ej, 1, Ej, 2, and Ej, n. Shape of the acquired data matrix represents “channels by number of samples,” where j = unique trail in data. For ‘n’ classes, Eq. (1) can be used to compute the composite spatial covariance matrix E.

$$ E=\sum \limits_{j=0}^{M_n}{E}_{j,n}{E}_{j,n}^T $$
(1)

Where, Mn = All trials in ‘n’ classes

$$ n=1,2\dots \dots n $$

Here, the classification of three classes is done; therefore, n = 3 results three covariance matrices, i.e., E1, E2 and E3 which results in the composite covariance matrix, i.e., E = E1, E2 and E3. This matrix is further factored by using (2).

$$ E={P}_0\partial\ {P}_0^T $$
(2)

Where, P0 = L × L unitary matrix of principal components

= L × L diagonal matrix of eigenvalues

The major goal of CSP is to find transformed data variance between distinct multiclass data which is computed by using (3)

$$ {\max}_sW(s)=\frac{s^T{E}_{c1}s}{s^T{E}_{c2}s}s.t.{\left\Vert s\right\Vert}_2=1 $$
(3)

Where, W(s) = Rayleigh quotient maximization

s = Spatial filter

s2 = the n2 normal

Ec1 = Covariance matrix of class 1

Ec2 = Covariance matrix of class 2

The solution of the generalized eigenvalue Eq. (4) results in W(s).

$$ {\sum}_1s=\lambda {\sum}_2s $$
(4)

Finally, from the learned spatial filter matrix ‘S’, the eigenvectors are calculated to compute the projection of an EEG signal using Eq. (5).

$$ {P}_n={S}_n^T\ G $$
(5)

Where

P = EEG signal’s CSP projections matrix

S = Learned spatial matrix

G = EEG signal

After CSD, sub-band scores are generated by applying LDA which are then used as features. These features are computed by using the projection matrix that guarantees maximum separability by maximizing the ratio of the variance between and within different classes [21]. The cost function for the projection matrix is generated by LDA using Eq. (6).

$$ C=\frac{S_{\mathrm{lda}}^T\ {V}_B{S}_{\mathrm{lda}}\ }{S_{\mathrm{lda}}^T\ {V}_W{S}_{\mathrm{lda}}} $$
(6)

Where, C = cost function for LDA projection matrix of EEG signal

Slda = Projection matrix generated by LDA

VB = Variance among MI-based right and left hand classes

VW = Variance within MI-based right or left hand classes

VB and VW are calculated by (7) and (8) respectively.

$$ {V}_B=\left({f}_2-{f}_1\right){\left({f}_2-{f}_1\right)}^T $$
(7)
$$ {V}_W=\sum \limits_{Z_p\in {n}_1}{\left({Z}_p-{f}_1\right)}^2+\sum \limits_{Z_p\in {n}_2}{\left({Z}_p-{f}_2\right)}^2 $$
(8)

Where f1 and f2 represent means of CSP computed features for empirical classes. Finally, the one-dimensional score by LDA can be calculated by (9). These scores are then sent to the feature selector.

$$ O={S}_{\mathrm{lda}}^T{z}_p $$
(9)

2.3 Feature SBFS

Optimal features are selected using selection using sequential backward floating selection (SBFS). It is a feature selection technique which selects among all the provided features in such a way that the overall performance of the system could be improved. For example, ‘k’ optimal features are selected from ‘n’ total features in a way that the performance of the classifier should be maximized due to selected features. Let us assume F is a set of all features,

$$ F=\left\{{f}_1,{f}_2,\dots \dots, {f}_n\right\} $$

If ‘F’ set is given to the SBFS technique as input then the selection of set ‘P’ (set of output features) is done in such a way that, first of all, the classifier will be trained on all of the provided features, i.e., set ‘F.’ In the next iterations, those features are removed from the input set which causes the low average accuracy of the system. These iterations for feature elimination are carried out until the ‘k’ number of features is not obtained. Once the iterations are finished, then the set ‘P’ (a subset of ‘F’) containing ‘k’ selected features is returned.

$$ P=\left\{{p}_j|j=1,2,\dots \dots, k;{p}_j\in B\right\}\mathrm{where}\;k=\left(0,1,2,\dots \dots, n-1\right) $$

2.4 Classification

After feature selection, the classifier is trained which estimates the accuracy of the system by predicting each class label from the testing model. The classifiers used in the proposed methodology are explained in this section.

2.5 SVM

Support vector machine (SVM) is called the supervised technique of machine learning because it requires labeled training dataset. It acts as a linear classifier because it draws a separating line known as hyperplane between data samples of different classes in a dataset. For margin maximization of training data, measurement of an optimal hyperplane is the goal of SVM. Hyperplane can be calculated by equation (10),

$$ {v}^Ty=0 $$
(10)

Where v and y = vectors containing EEG samples

As the hyperplane is a line which separates different classes in data, the equation for hyperplane is the same as the equation of a line which can be calculated by the dot product of v and y where

$$ v=\left[\begin{array}{c}-d\\ {}-c\\ {}1\end{array}\right]\kern2.25em \&\kern2em y=\left[\begin{array}{c}1\\ {}s\\ {}t\end{array}\right] $$
(11)

Once the hyperplane is computed then the distance among hyperplane and the closest data sample is calculated, which is known as margin value. Double of the calculated margin value is known as margin, no data sample occurs in this region. After hyperplane calculation, the classifier is considered ready for classification purpose. Now, when a test data is passed through the trained classifier it calculates the distance of the data sample from the margin. Then that class label is assigned to the data from which the computed distance is nearest.

2.6 KNN

The k-nearest neighbor (KNN) algorithm works on the principle of forming a majority vote between the ‘k’ most similar instances and a given test data sample (an unseen data). While ‘k’ represents a positive integer, which should be small. The performance of this algorithm depends on two factors, i.e., a suitable similarity function and an appropriate value for k. This similarity is found according to a distance matrix between two data samples. Euclidean distance is used as the most popular way to find the distance between the data samples. The equation for this method is

$$ d\left(p,{p}^{\prime}\right)=\sqrt{\left({p}_1+{p}_1^{\prime}\right)+\left({p}_2+{p}_2^{\prime}\right)+\dots +\left({p}_n+{p}_n^{\prime}\right)}\ \Big) $$
(12)

The algorithm is also known as lazy learning algorithm because it refers to the decision to generalize the training data samples until a new query is faced. The major assumption of this algorithm is that the class probabilities are locally constant approximately, so it is one of the simplest machine learning algorithms.

2.7 NBPW

Naïve Bayesian Parzen window (NBPW) is a fast classification algorithm as compared to other classification algorithms which follow the principle of Bayes theorem of probability. A major assumption in this algorithm is that one feature in the data does not relate to any other feature. As NBPW is based on Bayes theorem which provides the methodology to calculate posterior probability P (a | y) from P (a), P (y) and P (y | a) using Naive Bayesian equation

$$ P\ \left(a\ |\ y\right)=\frac{P\ \left(y\ |\ a\right)\ P\ (a)}{P\ (y)} $$
(13)

Where

a = target class

y = predictor

P(a | y) = posterior probability

P(a) = prior probability of class

P(y) = prior probability of predictor

P(y | a) = probability of predictor given class

The working principle of NBPW is that, firstly the dataset is converted into a frequency table then probabilities are computed to create a likelihood table. In the end, the posterior probability is computed by using the naïve Bayesian equation.

3 Results and discussion

For the evaluation of the proposed system, results are validated on two different datasets i.e., dry and wet gel electrodes. Dry electrodes consist of Emotiv Epoc which is a 14-channel device with 128 Hz sampling frequency, whereas, wet gel electrodes consist of 8-channel EEG device and data was sampled at 256 Hz. Figure 2 shows the 10–20 electrode system used for placement of wet gel electrodes. EEG signals were acquired for three distinct MI tasks (neutral, the imagination of left and right-hand movements). The subjects were instructed to stay calm and imagine their hand movements according to different stimulus shown on the screen. For each trial of 8 s, “blank screen” for 4 s, “+sign” for 2 s, and a “specific stimulus” for 2 s were shown on the screen. In each session, there were 240 trails, i.e., 80 trials for each task. Figure 3 shows the scheme used for data acquisition used for acquisition of both datasets. EEG data of those electrodes which add eye artifact data in the whole dataset (i.e., electrodes placed at frontal lobe) are not included so that noise could be minimized.

Fig. 2
figure 2

Electrode placement according to 10–20 electrode system. Different electrode placement positions where the electrodes are placed for EEG data acquisition are shown in this picture. The shaded ones represent the electrodes that are used to acquire EEG data for this study

Fig. 3
figure 3

Data acquisition scheme. This picture represents the whole scheme used for EEG data acquisition. It is followed to show various stimuli at different time. By using this scheme, a specific time stamp for a specific task is also placed in the dataset file which further helps to interpret that which specific data samples are associated with a specific task

After preprocessing by filter bank, 12 features are extracted from CSP and LDA. Optimal features are then selected from these extracted features. Then the accuracy of the system is compared for different classifiers (i.e., SVM, KNN & NBPW) by varying number of selected features. Figures 4 and 5 show a graphical comparison for the accuracy results of cross-validations performed on proposed model with various classification algorithms. For dry electrodes, the proposed approach shows a maximum accuracy of 60.61% for 10 selected features with NBPW as classifier as shown in Figure 4. In the case of SVM and KNN, maximum system output is 55.31% and 50.76% which are resulted for 6 and 3 selected features respectively. Figure 5 shows that for MI-based EEG signals acquired using wet gel electrode, a maximum of 86.50% accuracy is yielded. This accuracy resulted for four selected features with KNN as a classifier. For the same dataset, the maximum system output for SVM and NBPW is 76.59% with 6 selected features and 75.11% with 7 selected features respectively. Table 2 provides a tabular comparison of the accuracy of the proposed model for different selected features on SVM, KNN, and NBWP classifiers for both dry and wet gel electrodes.

Fig. 4
figure 4

Graphical comparison of accuracies by different classifiers for Emotiv Epoc dataset. This figure represents the graphical representation of calculated results for EEG dataset acquired by using Emotiv Epoc headset. These results are computed by varying number of selected features for three different classifiers. The gradual increase in the accuracies shows the importance of classifier

Fig. 5
figure 5

Graphical comparison of accuracies by different classifiers for wet gel electrodes headset dataset. This figure represents the graphical representation of calculated results for EEG dataset acquired by using wet gel electrodes. These results are computed by varying number of selected features for three different classifiers. The gradual increase in the accuracies shows the importance of classifier

Table 2 Comparison of results for different selected features, classifiers, and datasets

The proposed system yields 7% increase in accuracy as compared to literature which shows that the use of filter bank, optimal feature selection, and classifier in the proposed methodology has a remarkable influence on the classification accuracy of the system. The results for Emotiv Epoc dataset shows that NBPW produces better results than SVM and KNN with the proposed model but for wet gel electrode dataset, KNN outperforms in comparison to SVM and KNN.

4 Conclusion

This paper presents a framework for the classification of multiclass MI-based EEG signals. In the proposed model, results were evaluated on two different headsets. The proposed methodology shows a maximum of 60.61% and 86.50% accuracy for classification of motor imagery signals by using Emotiv Epoc and wet gel electrode headset respectively. The overall performance of the system shows a 7% increase in accuracy than previously used techniques for multiclass EEG classification. The results show that the use of overlapped filter bank, optimal feature selection, and classifier for multiclass classification have the capability to control BCI applications and can be tested for controlling BCI applications in the real world. Currently, further research is being done on the improvement of results by optimal channel selection.

Availability of data and materials

The data and materials used in this study are not publicly available. EEG data acquisition requires a lot of resources and work, i.e., EEG device and trained staff for headset placement. As the authors have acquired EEG data on their own by bearing all costs therefore, they have copyrights. But data can be provided on request with proper reason on email and with permission of Dr. Usman Ghani Khan.

Abbreviations

Adaboost-ELM:

Adaboost extreme learning machine

BCI:

Brain-computer interface

CSP:

Common spatial pattern

EEG:

Electroencephalogram

FBCSP:

Filter bank common spatial pattern

FFDIAG:

Fast Frobenius Diagonalization

JAD:

Joint approximate diagonalization

K c :

Kolmogorov complexity

KNN:

k-nearest neighbor

LDA:

Linear discriminant analysis

MI:

Motor imagery

MIBIF:

Mutual-information best individual features

NBPW:

Naïve Bayesian Parzen window

RF:

Random forest

SBCSP-SBFS:

Sub-band common spatial patterns with sequential feature selection

SBFS:

Sequential backward floating selection

SRIT2NFIS:

Self-regulated interval type2 neuro fuzzy inference system

SRLDA:

Stacked regularized linear discriminant analysis

STFT:

Short-time Fourier transform

SVM:

Support vector machine

References

  1. M.A. Jatoi, N. Kamel, A.S. Malik, I. Faye, T. Begum, Biomed. Signal Process. Control (2014)

  2. L. Bi, X.-A. Fan, Y. Liu, IEEE Trans. Human-Machine Syst. 43, 161 (2013)

    Article  Google Scholar 

  3. T. Ball, M. Kern, I. Mutschler, A. Aertsen, A. Schulze-Bonhage, Neuroimage 46, 708 (2009)

    Article  Google Scholar 

  4. B. Zhang, J. Wang, and T. Fuhlbrigge, in 2010 IEEE Int. Conf. Autom. Logist. (IEEE, 2010), pp. 379–384.

  5. T. Alexandra-Maria, V. Mihajlovic, Y.-H. Chen, B. Grundlehner, J. Penders, S. Wouter, Biodevices (2014), pp. 12–22

    Google Scholar 

  6. A.B. Usakli, Comput. Intell. Neurosci. 2010 (2010)

  7. P. Sarma, P. Tripathi, M.P. Sarma, K.K. Sarma, ADBU-Journal Eng. Technol. Sarma 5, 2348 (2016)

    Google Scholar 

  8. I. Xygonakis, A. Athanasiou, N. Pandria, D. Kugiumtzis, P.D. Bamidis, Comput. Intell. Neurosci. 2018, 1 (2018)

    Article  Google Scholar 

  9. F. Lotte, C. Guan, IEEE Trans. Biomed. Eng. 58, 355 (2011)

    Article  Google Scholar 

  10. Q. Novi, C. Guan, T.H. Dat, P. Xue, Proc. 3rd Int. IEEE EMBS Conf. Neural Eng. 204 (2007)

  11. K.K. Ang, Z.Y. Chin, H. Zhang, C. Guan, 2008 IEEE Int. Jt. Conf. Neural Networks (IEEE World Congr. Comput. Intell.) 2390 (2008)

  12. H. Higashi, T. Tanaka, IEEE Trans. Biomed. Eng. 60, 1100 (2013)

    Article  Google Scholar 

  13. Y. Zhang, G. Zhou, J. Jin, X. Wang, A. Cichocki, J. Neurosci. Methods 255, 85 (2015)

    Article  Google Scholar 

  14. S. Kumar, A. Sharma, T. Tsunoda, BMC Bioinformatics 18, 545 (2017)

    Article  Google Scholar 

  15. K.K. Ang, Z.Y. Chin, C. Wang, C. Guan, H. Zhang, Front. Neurosci. 6(1) (2012)

  16. L.F. Nicolas-Alonso, R. Corralejo, J. Gomez-Pilar, D. Álvarez, R. Hornero, IEEE Trans. Neural Syst. Rehabil. Eng. 23, 702 (2015)

    Article  Google Scholar 

  17. T. Shiratori, H. Tsubakida, A. Ishiyama, and Y. Ono, 3rd. Int. Winter Conf. Brain-Computer Interface, BCI 2015 3 (2015).

  18. A. Mammone, M. Turchi, N. Cristianini, Adv. Rev. 1, 283 (2009)

    Google Scholar 

  19. X. Yong, C. Menon, PLoS One 10(1) (2015)

    Article  Google Scholar 

  20. F. Shiman, E. López-Larraz, A. Sarasola-Sanz, N. Irastorza-Landa, M. Spüler, N. Birbaumer, A. Ramos-Murguialday, J. Neural Eng. 14 (2017)

  21. J. Pylkkönen, P.A. Conf, Int. Speech Commun. Assoc. INTERSPEECH 1, 389 (2006)

  22. Q. She, Y. Ma, M. Meng, Z. Luo, Comput. Intell. Neurosci. 2015 (2015)

  23. L. Gao, W. Cheng, J. Zhang, J. Wang, Rev. Sci. Instrum. 87, 1 (2016)

    Google Scholar 

  24. W. Chmielnicki, K. Stąpor, Int. J. Appl. Math. Comput. Sci. 26, 191 (2016)

    Article  MathSciNet  Google Scholar 

  25. H. Meisheri, N. Ramrao, and S. Mitra, (2018).

  26. X. Shi, Blind Signal Process (Springer, Berlin Heidelberg, Berlin, Heidelberg, 2011), pp. 175–204

    Book  Google Scholar 

  27. S.R. Liyanage, J.X. Xu, C.T. Guan, K.K. Ang, T.H. Lee, Proc. Int. Jt. Conf. Neural Networks (2010)

  28. S. Ge, R. Wang, D. Yu, PLoS One 9, e98019 (2014)

    Article  Google Scholar 

  29. A. Bablani, D.R. Edla, S. Dodia, Procedia Comput. Sci. 143, 242 (2018)

    Article  Google Scholar 

Download references

Acknowledgements

We would like to acknowledge our subjects who helped us with EEG data acquisition.

Funding

Not applicable.

Author information

Authors and Affiliations

Authors

Contributions

JK participated in the development of the proposed design and its implementation, coordinated in the study, and helped out for document drafting. MH participated in the EEG data acquisition, design and implementation of the proposed study, and drafted the manuscript. UGK supervised, coordinated, and helped out for study and manuscript draft completion. RI supervised the study and helped out for manuscript draft completion. All authors read and approved the final manuscript.

Authors’ information

Not applicable.

Corresponding author

Correspondence to Razi Iqbal.

Ethics declarations

Competing interest

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Khan, J., Bhatti, M.H., Khan, U.G. et al. Multiclass EEG motor-imagery classification with sub-band common spatial patterns. J Wireless Com Network 2019, 174 (2019). https://doi.org/10.1186/s13638-019-1497-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13638-019-1497-y

Keywords