Next Article in Journal
Overt Word Reading and Visual Object Naming in Adults with Dyslexia: Electroencephalography Study in Transparent Orthography
Previous Article in Journal
Toe Box Shape of Running Shoes Affects In-Shoe Foot Displacement and Deformation: A Randomized Crossover Study
Previous Article in Special Issue
Thermal Cues Composed of Sequences of Pulses for Transferring Data via a Haptic Thermal Interface
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Intelligent Human–Computer Interaction: Combined Wrist and Forearm Myoelectric Signals for Handwriting Recognition

1
Department of Information Engineering, Università Politecnica delle Marche, 60131 Ancona, Italy
2
Deparment of Industrial, Electronics and Mechanical Engineering, Roma Tre University, 00146 Rome, Italy
*
Author to whom correspondence should be addressed.
Bioengineering 2024, 11(5), 458; https://doi.org/10.3390/bioengineering11050458
Submission received: 22 March 2024 / Revised: 19 April 2024 / Accepted: 2 May 2024 / Published: 4 May 2024

Abstract

:
Recent studies have highlighted the possibility of using surface electromyographic (EMG) signals to develop human–computer interfaces that are also able to recognize complex motor tasks involving the hand as the handwriting of digits. However, the automatic recognition of words from EMG information has not yet been studied. The aim of this study is to investigate the feasibility of using combined forearm and wrist EMG probes for solving the handwriting recognition problem of 30 words with consolidated machine-learning techniques and aggregating state-of-the-art features extracted in the time and frequency domains. Six healthy subjects, three females and three males aged between 25 and 40 years, were recruited for the study. Two tests in pattern recognition were conducted to assess the possibility of classifying fine hand movements through EMG signals. The first test was designed to assess the feasibility of using consolidated myoelectric control technology with shallow machine-learning methods in the field of handwriting detection. The second test was implemented to assess if specific feature extraction schemes can guarantee high performances with limited complexity of the processing pipeline. Among support vector machine, linear discriminant analysis, and K-nearest neighbours (KNN), the last one showed the best classification performances in the 30-word classification problem, with a mean accuracy of 95% and 85% when using all the features and a specific feature set known as TDAR, respectively. The obtained results confirmed the validity of using combined wrist and forearm EMG data for intelligent handwriting recognition through pattern recognition approaches in real scenarios.

1. Introduction

Nowadays, the role of surface electromyography (EMG) is central in the development of smart assistive technologies. This is certainly due to the large amount of tools available in the field of signal processing and machine learning. Indeed, the literature provides a variety of software packages that facilitate the training and validation of EMG-based human–machine interfaces able to decode the human intent of motion within a given set of movements [1,2]. Such interfaces find application in prosthetic control of upper and lower bionic limbs [3,4], but also for the realization of intelligent human–computer interactions in virtual and augmented reality, and for biometric identification [5,6,7,8]. Myoelectric interfaces can also find application in modern scenarios since they can be used to decode handwritten characters or digits [9,10], supporting the development of immersive rehabilitation protocols with a consistent involvement of the cognitive centres of the brain [11].
Many pieces of evidence in the literature suggest the use of strengthening the handwriting motor skills to mitigate the effects carried by the pathology [12,13,14]. In children affected by dysgraphia, for instance, handwriting exercises delivered through human–machine interfaces can be particularly beneficial as they may provide timely feedback to correct their style [13]. Another important rehabilitation scenario of handwriting involves Parkinson’s disease, where handwriting difficulties occur frequently and are generally known as micrographia, i.e., the reduction of writing amplitude, eventually resulting in a reduced legibility [15]. In this context, past studies have shown that handwriting exercises help improve the writing size in those subjects affected by Parkinson’s disease [16,17]. Another important application of handwriting rehabilitation is related to motor hand ability recovery in patients after a severe traumatic brain injury, or after coma. In both cases, the use of handwriting showed improvements in patient condition, showing progressively improved adherence to recovery of normal functionalities of the hands [18,19,20].
It should be noted that EMG-based interfaces for handwriting recognition software have been less investigated compared to other technologies that generally employ pattern recognition methods trained with 2D images recorded on touch screen tablets, which can also associate pressure information of the pen [21,22]. However, these data sources did not reflect the actual volition of human movement, and handwriting is recognized only after having complete 2D information [22]. On the other hand, EMG can directly mirror the motor control volition and data and can be used in a sequential manner to create architectures that recognize the handwriting online., i.e., with an update frequency that can be used to making the interaction more fluent [10]. Hence, although EMG has been considered as a potential option for handwriting recognition, only a few studies have explored its use in classifying digits or letters [9,10]. The recognition of handwritten characters from EMG data has been tackled using template matching, dynamic time warping, and deep-learning methods [9,23,24]. However, these approaches require large databases and impose a high computational burden during the training phase, making the practical applicability of such architectures limited. Indeed, in [9], each participant wrote 36 characters on a screen, and each character was repeated one thousand times to create a dataset useful to train deep-learning-based models. Confirmation regarding the large amount of data for delivering reliable EMG-based deep-learning models can also be found in the recent literature [8], especially when one would realize plug-and-play devices. Furthermore, similar to [24], the EMG data were first mapped to pen coordinates before the classification step, which was obtained from a tablet. Such aspects may challenge the applicability of EMG-based human–computer interfaces when they need to be tailored on specific subjects [10].
Hence, the development of reliable and fast-calibrating myoelectric interfaces based on shallow pattern recognition technique results are still appealing and deserve to be investigated. In passing, recent studies have reported advancements in the EMG electrode locations for the development of human–computer interfaces able to extend the capability of the human being in smartly interacting with software [8]. To do this, the recent literature has focused on showing the importance of the forearm and wrist as a good location to acquire EMG for the development of transferable architectures in real scenarios [8,10,25,26]. However, the full potentialities of the aforementioned architectures were not completely investigated in complex handwriting tasks as the recognition of a set of words. Hence, the aim of this study is to investigate the feasibility of using combined forearm and wrist EMG signals to develop a 30-word handwriting recognition architecture. For this reason, five consolidated feature sets in the field of myoelectric control were employed to extract signal characteristics in the time and frequency domains. Moreover, the possible advantage of opportunely smoothing the classifier output signal was also assessed by applying a majority voting approach [27].
The paper is thus organized as follows: Section 2 reports the experimental protocol, the methodologies for processing EMG data, the pattern recognition algorithms, and the metrics used for assessing classification performances. For results and discussion, respectively, Section 3 and Section 4 present the performances obtained for all the classifiers with respect to the five feature sets employed and the implications of using such models with and without correction strategies. Finally, Section 5 concludes the manuscript with the main findings of the present work and possible associated limitations.

2. Materials and Methods

2.1. Study Population

A total of six healthy subjects, three females and three males aged between 25 and 40 years, were recruited for the study considering an age range in accordance with [10,25,28]. Subjects did not manifest any impairment or cognitive disturbance, and they were not affected by neurological disorders. Before applying for the study, each subject was informed about the experimental procedure, which involved the use of commercial devices with conformity declaration approval. All of them provided written consent to participate in the experiment, which was conducted following the protocols of the Declaration of Helsinki.

2.2. Experimental Protocol

All participants were equipped with a total of six surface EMG electrodes with a sampling frequency of 1000 Hz that operated synchronously (FREEEMG system, BTS-Bioengineering, Milan, Italy). The electrodes were placed by the same expert operator over the forearm and wrist in order to reduce the inter-operator variability. More specifically, the first four electrodes were employed for covering the forearm: electrodes 1 and 4 were placed on the extensor digitorum and the flexor carpi ulnaris, whereas electrodes 2 and 3 recorded the myoelectric activity of the flexor carpi radialis and the brachioradialis [10]. The last two electrodes, i.e., electrodes 5 and 6, were placed respectively on the distal ending of the flexor carpi radialis in correspondence of the deep flexor pollicis longus, and in correspondence of the extensor digiti minimi (Figure 1). Such a configuration guaranteed a wide spatial coverage of the arm, facilitating the recording of EMG signals from the bellies of the deep layer of muscles involved in the control of fine finger and wrist movements [29,30]. All the experimental sessions were conducted within the Movement Analysis Laboratory of the Università Politecnica delle Marche.
After being instrumented, all subjects were asked to sit on a chair in a comfortable posture for writing with their dominant hand in their natural cursive style. Then, they were asked to take a pen and write on a sheet of paper a total of 30 words, i.e., respectively, 10 nouns, adjectives, and verbs chosen among the most commonly employed in English (http://oxforddictionaries.com (accessed on 22 of May 2023); see Table 1). Each word was written 10 times with a pause of 5 s between two consecutive written words. The number of repetitions was selected to obtain a consistent amount of data for creating balanced training–validation and testing folds. This mitigated possible biases in the computation of the performance metrics, and at the same time allowed us to keep the acquisition time within reasonable limits [4]. Moreover, after completion of a sequence, each subject was asked to take a break for at least two minutes to prevent fatigue. During each rest period, the EMG data were saved in an appropriate folder and the subject was ready to face another trial. Each trial contained information relative to a specific word, and subjects proceeded by writing nouns, adjectives, and verbs in the order indicated by the class reported in Table 1. Eventually, for each subject a total of 30 recordings, i.e., one per each word containing 6 EMG signals with 10 activation events, were available for pre-processing and feature extraction.

2.3. Data Pre-Processing and Feature Extraction

For each record, EMG signals were filtered with a fourth-order, zero-phase Butterworth band-pass filter between 35 Hz and 450 Hz. Then, muscle onset state for data segmentation was opportunely identified with the threshold method [31,32]. An example of the segmentation step for a data record is reported in Figure 2.
The six EMG signals associated with each word repetition were segmented to create a database, allowing for the extraction of features in both the time and frequency domains (TD and FD) [1,4]. A sliding window of 150 ms with an overlap of 75 ms was used to calculate the features listed in Table 2 [10].
For each subject, a matrix was created with a number of rows corresponding to the total number of segmented windows and a number of columns corresponding to all the features computed for each channel. Data at feature level were then aggregated according to the feature sets that are commonly used for myoelectric pattern recognition [1,10]. In this study, five sets of characteristics were considered, the first being the Hudgins set, which consists of four time-domain features (MAV, WL, SSC, ZC) [33]. The second set was the Du set, which includes six time-domain features i.e., IEMG, VAR, WAMP, WL, SSC, ZC [34]. Two other feature sets proposed in [1] were also taken into account: Phinyomark 1, which is composed of five time-domain and two frequency-domain features, i.e., MAV, WL, WAMP, ZC, AR, MNF and PSR, and Phinyomark 2, which is made up of four time-domain features, i.e., WPermEn, CC, RMS and WL. Finally, the last feature set considered was TDAR, which is composed of seven features, namely MAV, SSC, WL, VAR, WAMP, AR, and ZC [35].

2.4. Classification Algorithms

In this study, three state-of-the-art classification algorithms were used for myoelectric control: linear discriminant analysis (LDA), support vector machine (SVM), and K-nearest neighbours (KNN) [1,4,36]. Concerning the LDA, it is a statistical learning model that has demonstrated applicability in myoelectric hand gesture recognition [1,3,37]. Here, a principal diagonal covariance matrix model was assumed to model the data [37]. Regarding the SVM, a linear kernel was employed, and a one-versus-one approach was used to deal with the multi-class nature of the handwriting recognition problem through SVM [37,38]. Finally, KNN is a non-parametric machine-learning model that requires the definition of the number of neighbours (k) to consider, and a distance metric between data points [37,38]. In this case, k was set equal to 2 and Euclidean norm was used.

2.5. Pattern Recognition Tests

Two pattern recognition tests were conducted to assess the possibility of classifying fine hand movements, that is writing 30 words, through EMG-based pattern recognition. In the first one, all the features in TD and FD were combined and used to train and test intra-subject models. Thus LDA, SVM, and KNN were trained for each subject using a five-fold cross validation scheme on 70% of the subject’s data, while the remaining 30% was used for testing. As done in previous studies [4,10], attention was paid to balancing the data split in each stage, i.e., learning, validation, and testing, to reduce bias in the models, thus avoiding overoptimistic results.
The second test was conducted to evaluate the classification performance of the aggregated sets reported in Section 2.3. This experiment was implemented to determine if consolidated feature extraction schemes can be useful in a complex classification problem, such as the one faced in this study, which involves more precise control of the wrist and hand movement. To do this, LDA, SVM, and KNN were trained to model intra-subject EMG-based classification of the 30 written words using the aforementioned training–validation and classification scheme. Thus, the Hudgins, Du, Phinyomark 1, Phinyomark 2, and TDAR sets were split into 70% of training data, which were then used in a five-fold cross validation scheme, while the remaining 30% was used for testing. It is worth noting that, in both experiments, the training and testing sets were split randomly but balanced, as suggested in [10].
Moreover, the effect of smoothing the classification output using majority voting (MV) was investigated, as suggested when classifiers have to be used in real scenarios [29,39]. MV is a post-processing approach used to enhance the model’s performance, and it utilizes a stream of class decisions that resulted from a sliding window with an overlapping scheme to reduce any potential noisy decision [40]. More formally, given a streamed number of votes, M, from the classifier, the smoothed MV decision, d MV , can be expressed as
d MV = arg max c i = M 2 M 2 I ( d i = c ) ,
where c represent the class label and I ( d i = c ) is an indicator function that equals 1 if the ith vote, d i , is equal to class c, and 0 otherwise [27,29,39]. The number of votes, M, is determined by the processing time, the time consumed during feature extraction and classification, and the acceptable delay, which is the response time of the control system [27]. In this study, M was set to 4.

2.6. Performance Metrics and Statistical Analysis

To properly assess the results obtained in the pattern recognition tests described in Section 2.5, three metrics have been taken into account during the testing of the models, i.e., the accuracy, the F1 score, and the Matthews correlation coefficient (MCC). The accuracy can be computed as the ratio between the correct predictions over the total predictions made, more formally:
Acc = T P + T N T P + F P + T N + F N ,
where T P , T N , F P , and F N are, respectively, the number of true positives, true negatives, false positives, and false negatives.
The second metric used to assess the goodness of the performances is the F1 score, which is the harmonic mean between recall and precision. Such a metric combines a measure of the ability to correctly categorise the cases, with an index of robustness, given by the proportion of instances that are not missed. The F1 score determines how well the model performs and it is computed as
F 1 = Precision · Recall Precision + Recall ,
where
Precision = T P T P + F P
Recall = T P F N + T P .
The last metric employed was the MCC, which is more sensitive to unbalanced datasets. Thus, although in this study attention was paid to balancing the split between training and testing data, MCC can be employed to confirm or not the indication given by the accuracy metric. The MCC metric can range between −1 and 1, and positive values close to 1 indicate a good performances of the model. Such a metric can be computed as
MCC = T P · T N F P · F N ( T P + F P ) · ( T P + F N ) · ( T N + F P ) · ( T N + F N ) .
The statistical differences between different conditions were assessed using the Wilcoxon paired rank sum test, with statistical significance set at 0.05.

3. Results

Table 3 summarizes the results obtained testing for the KNN, SVM, and LDA models. The KNN showed good classification performances in the 30-word classification problem, with a mean accuracy of 85%. Such goodness is also confirmed by F1 and MCC, which yielded values of 84%. On the other hand, SVM and LDA showed poor performances, with a mean accuracy not superior to 65% and 37%, respectively. These results were confirmed by MCC and F1, which indicated that the use of all the features grouped together, if fed to the LDA and SVM models, did not lead to robust results among the subjects.
Regarding the second experiment, Figure 3 summarizes the accuracy obtained in testing for the five sets and for the three machine-learning models used. As confirmed in the first experiment, KNN outperformed both SVM and LDA, with the latter showing the worst performancs. These results are also confirmed when the MV scheme is applied, as reported in Figure 4.
The KNN model had an average accuracy higher than 70%, with the TDAR and Phinyomark 1 sets performing significantly better than the other sets (p < 0.05). However, when comparing the two sets, there was no significant difference in the mean accuracy (p = 0.5887). The mean accuracy for TDAR and Phinyomark 1 was 74.6% and 73.4%, respectively, with no relevant difference in the interquartile range (see Figure 3). This supports the lack of a significant difference between the two sets. The same can be said for the other two classification algorithms, with the TDAR set having the best performance metrics, even with poor performances, i.e., average accuracy of 39.3% for the SVM and of 29.0% for the LDA algorithm without the MV application.
Concerning the application of MV, it should be noted that it boosted the performances of all the classifiers employed. In the case of KNN (see Figure 4), the role of feature sets is completely lost, as no significant differences were found when comparing all five sets. This indicates that the applied post processor consistently improved the classification performance up to a mean accuracy that is higher than 95%. However, if one keeps the focus on the SVM model (Figure 4), TDAR and Phinyomark 1 confirmed their superior performances compared to the other sets (p < 0.05), with a non-significant difference between the two. In this case, LDA showed poor performance in terms of accuracy, with large variability among the subjects (Figure 4). Moreover, even if TDAR and Phinyomark 1 again seemed to be the best sets, i.e., mean accuracy, respectively, of 61.4% and 60.8%, the pair comparison between the sets did not reveal any significant difference (p > 0.05), showing a not preferred set when LDA was used as the classifier for this task.

4. Discussion

In this study, different intelligent models that combined EMG information extracted from the wrist and forearm were developed and evaluated for solving 30-word hand writing recognition. Although previous studies showed the possibility to transfer myoelectric control scheme in the field of handwriting recognition [10], the analysis was focused on a set of ten digits, whereas in the present work a large dataset containing EMG signals recorded for a large set of words was investigated to better unveil the nature of myoelectric control during handwriting. Indeed, the EMG signals from sensors positioned on the wrist resulted in rich information to characterise complex motor tasks, and they can be combined with standard electrode configurations that limit the acquisition only at the level of the forearm [10,25].
As shown in Figure 2, wrist EMG activity resulted in good quality signals with the amplitude significantly higher in the activation epochs with respect to rest phases. This was encountered in all the subjects analysed, supporting the availability of rich information from data recorded close to the wrist, even if muscle tendon terminations are present in a greater percentage close to the wrist with respect to active fibres [41]. However, the surface EMG probes may also have captured motor commands coming from the active fibres present in connection with the wrist joint [41], eventually providing signals with detectable information [10].
The first pattern recognition test confirmed the hypothesis about the feasibility of using combined forearm and wrist EMG probes for solving handwriting recognition problems with consolidated machine-learning techniques and aggregating state-of-the-art features extracted in TD and FD [1,27,42]. It should be noted that such an experiment confirmed previous results in the literature [9,43], but it also revealed the possibility of using consolidated myoelectric control technology with shallow machine-learning models in the field of handwriting recognition, thus supporting the transfer of methodologies available in the literature for developing human–machine interfaces able to recognize written words by the humans solely based on EMG signals [10,25]. Indeed, the aggregation of a large set of features in the TD and FD domains without a refinement in the processing pipeline (i.e., feature selection, feature reduction, postprocesssing in the classification output) was enough for reaching a classification accuracy greater than 80% for all the subjects in testing for the KNN model (Table 3). Moreover, the trends observed for F1 score and MCC metrics supported the goodness of the accuracy as the principal metric for comparing the results obtained in all the performed experiments since the MCC and F1 showed value that are close to 1, with a trend similar to the accuracy (Table 3).
Another point that should be highlighted in the first pattern recognition experiment is the superior performance of KNN as compared to LDA and SVM, and this appeared consistently among the subjects. This may be imputed to the non-parametric nature of KNN [37], which does not require a priori knowledge of the data distribution and it can fit irregular decision functions due to an increase of complex patterns present in the data. Conversely, both the LDA and linear SVM-based architectures may suffer in fine partitioning the feature space required in this study and are not encountered in typical hand gesture recognition problems where both linear SVM and LDA resulted to be efficient [25,44]. Hence, rather than employ linear approaches, the results suggest the use of nonlinear kernels to better capture the complex hidden pattern within the data. This was confirmed when features were aggregated in specific sets for training and testing the models, as indicated in the second experiment (see Section 2.5). Indeed, the KNN remains the best model with and without the use of the MV postprocessor (Figure 3 and Figure 4), supporting the use of KNN to solve the myoelectric control problem with a large number of classes. However, in the best scenario without MV, KNN reached a mean accuracy lower than 80% when TDAR or Phinyomark 1 was employed (Figure 3), and this may hamper the direct use of the model in a real context. In any case, the investigation of the performances given by specific feature sets was important to better frame the possible line of processing for extracting reliable information from EMG data. Thus, the present study supports TDAR as a viable set of EMG characteristics, even when sparse EMG probes are used instead of high-density channels [3]. It is also worth noting that TDAR is slightly different from Hudgins (see Section 2.3). The main difference lies in the inclusion of VAR WAMP and AR coefficients, which are not present in Hudgins. These additional features may be more sensitive to fine motor control commands, which could be more difficult to capture using Hudgins [3].
Comparing the KNN accuracy obtained in this work with respect to previous studies in which the letter recognition problem was faced, one can notice that, without the use of MV, the model fed with TDAR shows accuracy value greater than the one in [23], where a mean accuracy of 84.3% was obtained through dynamic time warping. However, the model shows lower performance with respect to other approaches proposed in [9,24,28,45], where a mean accuracy greater than 92% was obtained by using techniques based on template matching, modified time warping, and deep-learning approaches. These values of accuracy were reached only when KNN was combined with MV (see Figure 4), where a 99.0% detection accuracy was obtained, making the result comparable with the 98.6% obtained in [28], where auto-encoder for feature extraction and SVM for classification were used with 15 young adults. However, it is important to highlight that in the present study the problem faced has a higher level of complexity. Indeed, the EMG patterns belong to a word in which multiple letters are written consecutively and for which complex motor control planning is actuated by the central nervous system. Moreover, the present classification task took into account 30 classes, i.e., 30 words, which is a high number if compared with the majority of myoelectric pattern recognition studies focused on handwriting recognition, as confirmed in [9]. These considerations not only justify the lower performance obtained in testing (see Figure 3), i.e., not greater than 75%, but also motivates the use of MV for smoothing decision output [27], eventually boosting the performances of the KNN model toward values that can guarantee future validation steps in a real-time context. With all the feature sets, the KNN reached an accuracy value greater than 95%, suggesting the use of the model in practical applications. Furthermore, by comparing the accuracy of KNN in the first and second experiments without MV (respectively, Table 3 and Figure 3), one can recognize that the use of aggregated TD and FD features contributes to the development of classifiers with superior performance compared to the state-of-the-art feature sets employed in the second experiment. This suggests that the use of specific sets, as defined in Section 2.3, can lead to a loss of information that eventually limits the capability of training models to classify unseen data. However, enlarging the number of features increases the complexity of the KNN, indicating the potential use of feature reduction schemes, such as spectral regression methods, when considering a wider range of features, as done in the first experiment [3].
It should be mentioned that the MV technique introduces a time latency of M / 2 samples of the decision output since it provides the most probable class knowing M / 2 samples before and after the current decision output [27]. Considering the selection of M = 4 and considering a feature update rate of 1 / 0.075 = 13.33 Hz, this introduces a latency of 150 ms, and it defines a new decision output frequency of 1 / 0.15 = 6.66 Hz. Thus, performance improvements were paid at the cost of having a final model with an update frequency that is half of the original one. This aspect may represent a limitation, and it suggests a future research line that can be investigated. Indeed, modern myoelectric interfaces can require output update rates that guarantee efficiency up to 25 Hz. This finds confirmation in [27], where feature update and output decision rates were increased by selecting the TDAR feature set together with reduced window length and adopting the high-density EMG sensing technology. A further limitation of the study can be related to the number of subjects recruited to undergo the experiment. However, it should be noted that the number of subjects recruited is coherent with previous studies [9,10]. From this perspective, the enlargement of the dataset may represent a future step toward the generalization of such architecture in real practice, particularly if one is interested in investigating the use of deep-learning techniques [8], which have been shown to be efficient for handwriting recognition applications, even if they were not tested using small EMG window lengths [9]. Moreover, although in this study forearm and wrist EMG channels were used together, future research may focus on comparing the two mentioned locations for automatic handwriting recognition. Indeed, no consensus regarding the optimal location can be found in previous studies [10,25]. While forearm and wrist electrodes provided comparable results in hand gesture recognition problems [25], a drop in performance was observed when transitioning from forearm to wrist in the automatic recognition of handwritten digits [10].

5. Conclusions

The study investigated the combination of wrist and forearm EMG data for intelligent handwriting recognition architectures able to decode words from myolelectric signals in healthy young adults. Shallow pattern recognition models, i.e., SVM, LDA, and KNN, were investigated to evaluate the possibility of transfer methodologies typically employed in myeoelctric control in the field of handwriting recognition. Among the aforementioned classifiers, KNN showed the best performances in the pattern recognition experiment proposed, i.e, when a large feature set made of 26 different type of features where considered, and when state-of-the-art feature sets were used with an MV post processor. This research suggests the possibility to develop myoelectric handwriting recognition systems with a consistent number of classes, i.e., 30 words using myoelectric information with approaches that can be used in real scenario.

Author Contributions

Conceptualization, A.T., F.V. and A.M.; methodology, A.T., S.R., F.V. and A.M.; software, A.T., S.R. and A.M.; validation, A.T., F.V. and A.M.; formal analysis, A.T. and A.M.; investigation, A.T., F.V., R.M., M.S. (Mara Scattolini) and A.M.; resources, L.B. and S.F.; data curation, A.T. and F.V.; writing—original draft preparation, A.T., S.R. and A.M.; writing—review and editing, A.T., S.R., F.V., S.C., M.S. (Maurizio Schmid), L.B., E.G., S.F. and A.M.; visualization, A.T., F.V., R.M., M.S. (Mara Scattolini) and A.M.; supervision, F.V., S.C., M.S. (Maurizio Schmid), L.B., E.G., S.F. and A.M.; project administration, A.T., F.V. and A.M.; funding acquisition, L.B., E.G. and S.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research has received funding from the financial program DM MiSE 5 Marzo 2018 project ChAALenge—F/180016/01-05/X43 and from the project Vitality—Project Code ECS00000041, CUP I33C22001330007—funded under the National Recovery and Resilience Plan (NRRP), Mission 4 Component 2 Investment 1.5—‘Creation and strengthening of innovation ecosystems’, construction of ‘territorial leaders in R&D’—Innovation Ecosystems—Project ‘Innovation, digitalization and sustainability for the diffused economy in Central Italy—VITALITY’ Call for tender No. 3277 of 30/12/2021, and Concession Decree No. 0001057.23-06-2022 of Italian Ministry of University funded by the European Union—NextGenerationEU.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

Data are made available upon reasonable request to the authors.

Conflicts 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. Phinyomark, A.; Quaine, F.; Charbonnier, S.; Serviere, C.; Tarpin-Bernard, F.; Laurillau, Y. EMG feature evaluation for improving myoelectric pattern recognition robustness. Expert Syst. Appl. 2013, 40, 4832–4840. [Google Scholar] [CrossRef]
  2. Khushaba, R.N. Correlation analysis of electromyogram signals for multiuser myoelectric interfaces. IEEE Trans. Neural Syst. Rehabil. Eng. 2014, 22, 745–755. [Google Scholar] [CrossRef] [PubMed]
  3. Khushaba, R.N.; Nazarpour, K. Decoding hd-emg signals for myoelectric control-how small can the analysis window size be? IEEE Robot. Autom. Lett. 2021, 6, 8569–8574. [Google Scholar] [CrossRef]
  4. Tigrini, A.; Verdini, F.; Fioretti, S.; Mengarelli, A. On the Decoding of Shoulder Joint Intent of Motion from Transient EMG: Feature Evaluation and Classification. IEEE Trans. Med. Robot. Bionics 2023, 5, 1037–1044. [Google Scholar] [CrossRef]
  5. Kwon, Y.; Dwivedi, A.; McDaid, A.J.; Liarokapis, M. Electromyography-based decoding of dexterous, in-hand manipulation of objects: Comparing task execution in real world and virtual reality. IEEE Access 2021, 9, 37297–37310. [Google Scholar] [CrossRef]
  6. Sugiarto, T.; Hsu, C.L.; Sun, C.T.; Hsu, W.C.; Ye, S.H.; Lu, K.T. Surface EMG vs. high-density EMG: Tradeoff between performance and usability for head orientation prediction in VR application. IEEE Access 2021, 9, 45418–45427. [Google Scholar] [CrossRef]
  7. Jiang, X.; Liu, X.; Fan, J.; Ye, X.; Dai, C.; Clancy, E.A.; Chen, W. Measuring neuromuscular electrophysiological activities to decode HD-sEMG biometrics for cross-application discrepant personal identification with unknown identities. IEEE Trans. Instrum. Meas. 2022, 71, 4009515. [Google Scholar] [CrossRef]
  8. Ctrl-labs at Reality Labs; Sussillo, D.; Kaifosh, P.; Reardon, T. A generic noninvasive neuromotor interface for human-computer interaction. bioRxiv 2024. [CrossRef]
  9. Beltrán Hernández, J.G.; Ruiz Pinales, J.; López Rodríguez, P.; López Ramírez, J.L.; Aviña Cervantes, J.G. Multi-Stroke handwriting character recognition based on sEMG using convolutional-recurrent neural networks. Math. Biosci. Eng. 2020, 17, 5435–5448. [Google Scholar]
  10. Tigrini, A.; Verdini, F.; Scattolini, M.; Barbarossa, F.; Burattini, L.; Morettini, M.; Fioretti, S.; Mengarelli, A. Handwritten Digits Recognition from sEMG: Electrodes Location and Feature Selection. IEEE Access 2023, 11, 58006–58015. [Google Scholar] [CrossRef]
  11. Liao, Y.Y.; Tseng, H.Y.; Lin, Y.J.; Wang, C.J.; Hsu, W.C. Using virtual reality-based training to improve cognitive function, instrumental activities of daily living and neural efficiency in older adults with mild cognitive impairment. Eur. J. Phys. Rehabil. Med. 2019, 56, 47–57. [Google Scholar] [CrossRef] [PubMed]
  12. Van Drempt, N.; McCluskey, A.; Lannin, N.A. A review of factors that influence adult handwriting performance. Aust. Occup. Ther. J. 2011, 58, 321–328. [Google Scholar] [CrossRef] [PubMed]
  13. Danna, J.; Velay, J.L.; Vietminh, V.; Capel, A.; Petroz, C.; Gondre, C.; Thoret, E.; Aramaki, M.; Ystad, S.; Kronland-Martinet, R.; et al. Handwriting movement sonification for the rehabilitation of dysgraphia. In Proceedings of the 10th International Symposium on Computer Music Multidisciplinary Research (CMMR)-Sound, Music & Motion, Marseille, France, 15–18 October 2013; pp. 200–208. [Google Scholar]
  14. Danna, J.; Velay, J.L. Basic and supplementary sensory feedback in handwriting. Front. Psychol. 2015, 6, 169. [Google Scholar] [CrossRef] [PubMed]
  15. Shukla, A.W.; Ounpraseuth, S.; Okun, M.S.; Gray, V.; Schwankhaus, J.; Metzer, W.S. Micrographia and related deficits in Parkinson’s disease: A cross-sectional study. BMJ Open 2012, 2, e000628. [Google Scholar] [CrossRef] [PubMed]
  16. Ziliotto, A.; Cersosimo, M.G.; Micheli, F.E. Handwriting rehabilitation in Parkinson disease: A pilot study. Ann. Rehabil. Med. 2015, 39, 586–591. [Google Scholar] [CrossRef] [PubMed]
  17. Nackaerts, E.; Broeder, S.; Pereira, M.P.; Swinnen, S.P.; Vandenberghe, W.; Nieuwboer, A.; Heremans, E. Handwriting training in Parkinson’s disease: A trade-off between size, speed and fluency. PLoS ONE 2017, 12, e0190223. [Google Scholar] [CrossRef] [PubMed]
  18. Kao, H.S. Calligraphy therapy: A complementary approach to psychotherapy. Asia Pac. J. Couns. Psychother. 2010, 1, 55–66. [Google Scholar] [CrossRef]
  19. Kao, H.S.; Lam, S.P.; Kao, T.T. Chinese calligraphy handwriting (CCH): A case of rehabilitative awakening of a coma patient after stroke. Neuropsychiatr. Dis. Treat. 2018, 14, 407–417. [Google Scholar] [CrossRef] [PubMed]
  20. Pąchalska, M.; Moskała, M.; MacQueen, B.D.; Polak, J.; Wilk-Frańczuk, M. Early neurorehabilitation in a patient with severe traumatic brain injury to the frontal lobes. Med. Sci. Monit. 2010, 16, CS157–CS167. [Google Scholar] [PubMed]
  21. Impedovo, D.; Pirlo, G. Dynamic handwriting analysis for the assessment of neurodegenerative diseases: A pattern recognition perspective. IEEE Rev. Biomed. Eng. 2019, 12, 209–220. [Google Scholar] [CrossRef] [PubMed]
  22. Drotár, P.; Mekyska, J.; Rektorová, I.; Masarová, L.; Smékal, Z.; Faundez-Zanuy, M. Evaluation of handwriting kinematics and pressure for differential diagnosis of Parkinson’s disease. Artif. Intell. Med. 2016, 67, 39–46. [Google Scholar] [CrossRef] [PubMed]
  23. Huang, G.; Zhang, D.; Zheng, X.; Zhu, X. An EMG-based handwriting recognition through dynamic time warping. In Proceedings of the 2010 Annual International Conference of the IEEE Engineering in Medicine and Biology, Buenos Aires, Argentina, 31 August–4 September 2010; pp. 4902–4905. [Google Scholar]
  24. Linderman, M.; Lebedev, M.A.; Erlichman, J.S. Recognition of handwriting from electromyography. PLoS ONE 2009, 4, e6791. [Google Scholar] [CrossRef] [PubMed]
  25. Botros, F.S.; Phinyomark, A.; Scheme, E.J. Electromyography-based gesture recognition: Is it time to change focus from the forearm to the wrist? IEEE Trans. Ind. Inform. 2020, 18, 174–184. [Google Scholar] [CrossRef]
  26. Botros, F.S.; Phinyomark, A.; Scheme, E.J. Day-to-Day Stability of Wrist EMG for Wearable-Based Hand Gesture Recognition. IEEE Access 2022, 10, 125942–125954. [Google Scholar] [CrossRef]
  27. Khushaba, R.N.; Kodagoda, S.; Takruri, M.; Dissanayake, G. Toward improved control of prosthetic fingers using surface electromyogram (EMG) signals. Expert Syst. Appl. 2012, 39, 10731–10738. [Google Scholar] [CrossRef]
  28. Singh, S.K.; Chaturvedi, A. Leveraging deep feature learning for wearable sensors based handwritten character recognition. Biomed. Signal Process. Control 2023, 80, 104198. [Google Scholar] [CrossRef]
  29. Englehart, K.; Hudgins, B. A robust, real-time control scheme for multifunction myoelectric control. IEEE Trans. Biomed. Eng. 2003, 50, 848–854. [Google Scholar] [CrossRef] [PubMed]
  30. Van Oudenaarde, E.; Oostendorp, R. Functional relationship between the abductor pollicis longus and abductor pollicis brevis muscles: An EMG analysis. J. Anat. 1995, 186, 509–515. [Google Scholar] [PubMed]
  31. Tigrini, A.; Mengarelli, A.; Cardarelli, S.; Fioretti, S.; Verdini, F. Improving EMG signal change point detection for low SNR by using extended Teager-Kaiser energy operator. IEEE Trans. Med. Robot. Bionics 2020, 2, 661–669. [Google Scholar] [CrossRef]
  32. Lorrain, T.; Jiang, N.; Farina, D. Influence of the training set on the accuracy of surface EMG classification in dynamic contractions for the control of multifunction prostheses. J. Neuroeng. Rehabil. 2011, 8, 25. [Google Scholar] [CrossRef] [PubMed]
  33. Hudgins, B.; Parker, P.; Scott, R.N. A new strategy for multifunction myoelectric control. IEEE Trans. Biomed. Eng. 1993, 40, 82–94. [Google Scholar] [CrossRef] [PubMed]
  34. Du, Y.C.; Lin, C.H.; Shyu, L.Y.; Chen, T. Portable hand motion classifier for multi-channel surface electromyography recognition using grey relational analysis. Expert Syst. Appl. 2010, 37, 4283–4291. [Google Scholar] [CrossRef]
  35. Huang, H.; Xie, H.B.; Guo, J.Y.; Chen, H.J. Ant colony optimization-based feature selection method for surface electromyography signals classification. Comput. Biol. Med. 2012, 42, 30–38. [Google Scholar] [CrossRef] [PubMed]
  36. Mengarelli, A.; Tigrini, A.; Fioretti, S.; Verdini, F. Identification of neurodegenerative diseases from gait rhythm through time domain and time-dependent spectral descriptors. IEEE J. Biomed. Health Inform. 2022, 26, 5974–5982. [Google Scholar] [CrossRef] [PubMed]
  37. Bishop, C.M.; Nasrabadi, N.M. Pattern Recognition and Machine Learning; Springer: New York, NY, USA, 2006; Volume 4. [Google Scholar]
  38. Taborri, J.; Palermo, E.; Rossi, S. Automatic detection of faults in race walking: A comparative analysis of machine-learning algorithms fed with inertial sensor data. Sensors 2019, 19, 1461. [Google Scholar] [CrossRef] [PubMed]
  39. Wahid, M.F.; Tafreshi, R.; Langari, R. A multi-window majority voting strategy to improve hand gesture recognition accuracies using electromyography signal. IEEE Trans. Neural Syst. Rehabil. Eng. 2019, 28, 427–436. [Google Scholar] [CrossRef] [PubMed]
  40. Farrell, T.R. Determining delay created by multifunctional prosthesis controllers. J. Rehabil. Res. Dev. 2011, 48, xxi–xxxviii. [Google Scholar] [CrossRef] [PubMed]
  41. Wang, S.; Hayashi, S.; Jin, Z.W.; Kim, J.H.; Yamamoto, M.; Murakami, G.; Abe, S. Distally-extending muscle fibers across involved joints: Study of long muscles and tendons of wrist and ankle in late-term fetuses and adult cadavers. Anat. Cell Biol. 2023, 56, 46–53. [Google Scholar] [CrossRef] [PubMed]
  42. Micera, S.; Carpaneto, J.; Raspopovic, S. Control of hand prostheses using peripheral information. IEEE Rev. Biomed. Eng. 2010, 3, 48–68. [Google Scholar] [CrossRef] [PubMed]
  43. Okorokova, E.; Lebedev, M.; Linderman, M.; Ossadtchi, A. A dynamical model improves reconstruction of handwriting from multichannel electromyographic recordings. Front. Neurosci. 2015, 9, 389. [Google Scholar] [CrossRef]
  44. Al-Timemy, A.H.; Serrestou, Y.; Khushaba, R.N.; Yacoub, S.; Raoof, K. Hand gesture recognition with acoustic myography and wavelet scattering transform. IEEE Access 2022, 10, 107526–107535. [Google Scholar] [CrossRef]
  45. Li, C.; Ma, Z.; Yao, L.; Zhang, D. Improvements on EMG-based handwriting recognition with DTW algorithm. In Proceedings of the 2013 35th Annual International Conference of the IEEE Engineering in Medicine and Biology Society (EMBC), Osaka, Japan, 3–7 July 2013; pp. 2144–2147. [Google Scholar]
Figure 1. Electromyographic (EMG) probes location in the dorsal (panel (a)) and medial (panel (b)) view of the arm. This setup was used to record the electrical activity of the wrist and forearm muscle for each subject. All probes was placed by the same expert in order to minimize effects due to inter-operator variability.
Figure 1. Electromyographic (EMG) probes location in the dorsal (panel (a)) and medial (panel (b)) view of the arm. This setup was used to record the electrical activity of the wrist and forearm muscle for each subject. All probes was placed by the same expert in order to minimize effects due to inter-operator variability.
Bioengineering 11 00458 g001
Figure 2. Segmentation of forearm (blue traces) and wrist (orange traces) EMGs performed to identify the handwriting 10-fold repetition of the word “great” by subject 2.
Figure 2. Segmentation of forearm (blue traces) and wrist (orange traces) EMGs performed to identify the handwriting 10-fold repetition of the word “great” by subject 2.
Bioengineering 11 00458 g002
Figure 3. Mean accuracy (ACC) among the subjects obtained with respect to the specific feature sets considered (see Section 2.5), reported in the x-axis of the figure.
Figure 3. Mean accuracy (ACC) among the subjects obtained with respect to the specific feature sets considered (see Section 2.5), reported in the x-axis of the figure.
Bioengineering 11 00458 g003
Figure 4. Mean accuracy (ACC) among the subjects obtained with respect to the specific feature sets considered (see Section 2.5), reported in the x-axis of the figure. In this case, the effect of using the majority voting (MV) approach is highlighted with respect to Figure 3.
Figure 4. Mean accuracy (ACC) among the subjects obtained with respect to the specific feature sets considered (see Section 2.5), reported in the x-axis of the figure. In this case, the effect of using the majority voting (MV) approach is highlighted with respect to Figure 3.
Bioengineering 11 00458 g004
Table 1. Each subject was asked to handwrite nouns (10), adjectives (10), and verbs (10) according to the experimental protocol described in Section 2.2. To each word corresponds a label number that associates the word with the specific class used in the myoelectric pattern recognition problem.
Table 1. Each subject was asked to handwrite nouns (10), adjectives (10), and verbs (10) according to the experimental protocol described in Section 2.2. To each word corresponds a label number that associates the word with the specific class used in the myoelectric pattern recognition problem.
NounsAdjectivesVerbs
Word Label Word Label Word Label
time1good11be21
person2new12have22
year3first13do23
way4last14say24
day5log15get25
thing6great16make26
man7little17go27
world8own18know28
life9other19take29
hand10old20see30
Table 2. EMG features extracted in the time (16) and frequency (10) domains (TD and FD, respectively), with their abbreviations. More information regarding their computation can be found in [1,10].
Table 2. EMG features extracted in the time (16) and frequency (10) domains (TD and FD, respectively), with their abbreviations. More information regarding their computation can be found in [1,10].
Domain Feature   Name Abbreviation
TDIntegrated EMGIEMG
Mean Absolute ValueMAV
Variance of sEMGVAR
Root Mean SquareRMS
Waveform LengthWL
Difference Absolute Mean ValueDAMV
Difference Absolute Standard Deviation ValueDASDV
Zero CrossingZC
Myopulse Percentage RateMYOP
Willison AmplitudeWAMP
Slope Sign ChangeSSC
Fuzzy EntropyFuzEN
Weighted Permutation EntropyWPermEN
Histogram of EMG, 10-binsHIST
Auto-Regressive Coefficients, 4th OrderAR
Cepstrum coefficients of the 4th Order AR processCC
FDMean FrequencyMNF
Median FrequencyMDF
Peak FrequencyPKF
Total PowerTTP
1st Spectral MomentSM1
2nd Spectral MomentSM2
3rd Spectral MomentSM3
Frequency RatioFR
Power Spectrum RatioPSR
Variance of Central FrequencyVCF
Table 3. Acc, F1, and MCC metrics for the first pattern recognition test for all the subjects with KNN, SVM, and LDA classifiers.
Table 3. Acc, F1, and MCC metrics for the first pattern recognition test for all the subjects with KNN, SVM, and LDA classifiers.
SubjectKNNSVMLDA
Acc F1 MCC Acc F1 MCC Acc F1 MCC
10.820.820.820.640.650.620.370.380.32
20.870.820.820.660.680.640.330.330.26
30.850.850.840.600.610.580.310.310.23
40.860.860.850.630.630.610.390.410.35
50.830.830.830.630.640.620.380.420.38
60.880.880.880.680.680.660.420.420.38
Average0.85 0.84 0.84 0.640.650.620.370.37 0.32
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Tigrini, A.; Ranaldi, S.; Verdini, F.; Mobarak, R.; Scattolini, M.; Conforto, S.; Schmid, M.; Burattini, L.; Gambi, E.; Fioretti, S.; et al. Intelligent Human–Computer Interaction: Combined Wrist and Forearm Myoelectric Signals for Handwriting Recognition. Bioengineering 2024, 11, 458. https://doi.org/10.3390/bioengineering11050458

AMA Style

Tigrini A, Ranaldi S, Verdini F, Mobarak R, Scattolini M, Conforto S, Schmid M, Burattini L, Gambi E, Fioretti S, et al. Intelligent Human–Computer Interaction: Combined Wrist and Forearm Myoelectric Signals for Handwriting Recognition. Bioengineering. 2024; 11(5):458. https://doi.org/10.3390/bioengineering11050458

Chicago/Turabian Style

Tigrini, Andrea, Simone Ranaldi, Federica Verdini, Rami Mobarak, Mara Scattolini, Silvia Conforto, Maurizio Schmid, Laura Burattini, Ennio Gambi, Sandro Fioretti, and et al. 2024. "Intelligent Human–Computer Interaction: Combined Wrist and Forearm Myoelectric Signals for Handwriting Recognition" Bioengineering 11, no. 5: 458. https://doi.org/10.3390/bioengineering11050458

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