Next Article in Journal
A Machine Learning Strategy for the Quantitative Analysis of the Global Warming Impact on Marine Ecosystems
Next Article in Special Issue
Ignoring Internal Utilities in High-Utility Itemset Mining
Previous Article in Journal
Influence of Centrifugal Buoyancy in Thermal Convection within a Rotating Spherical Shell
Previous Article in Special Issue
A Continuous Region-Based Skyline Computation for a Group of Mobile Users
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Improving the Performance and Explainability of Indoor Human Activity Recognition in the Internet of Things Environment

1
Graduate School of Natural and Applied Sciences, Dokuz Eylul University, Izmir 35390, Turkey
2
Department of Computer Engineering, Dokuz Eylul University, Izmir 35390, Turkey
3
School of Computing, Newcastle University, Newcastle upon Tyne NE1 7RU, UK
*
Author to whom correspondence should be addressed.
Symmetry 2022, 14(10), 2022; https://doi.org/10.3390/sym14102022
Submission received: 21 August 2022 / Revised: 3 September 2022 / Accepted: 23 September 2022 / Published: 26 September 2022
(This article belongs to the Special Issue Information Technology and Its Applications 2021)

Abstract

:
Traditional indoor human activity recognition (HAR) has been defined as a time-series data classification problem and requires feature extraction. The current indoor HAR systems still lack transparent, interpretable, and explainable approaches that can generate human-understandable information. This paper proposes a new approach, called Human Activity Recognition on Signal Images (HARSI), which defines the HAR problem as an image classification problem to improve both explainability and recognition accuracy. The proposed HARSI method collects sensor data from the Internet of Things (IoT) environment and transforms the raw signal data into some visual understandable images to take advantage of the strengths of convolutional neural networks (CNNs) in handling image data. This study focuses on the recognition of symmetric human activities, including walking, jogging, moving downstairs, moving upstairs, standing, and sitting. The experimental results carried out on a real-world dataset showed that a significant improvement (13.72%) was achieved by the proposed HARSI model compared to the traditional machine learning models. The results also showed that our method (98%) outperformed the state-of-the-art methods (90.94%) in terms of classification accuracy.

1. Introduction

Human activity recognition (HAR) is the task of correctly identifying human activities (i.e., walking, eating, standing, and working) by analyzing sensor data collected by Internet of Things (IoT) devices. It is useful for understanding the behavioral human patterns in an IoT system. Our work focuses on human activity recognition in indoor environments.
The indoor HAR systems are important in many domains, such as assisted living and healthcare [1,2], biometric user identification for security [3], wellbeing in smart homes [4], evaluating employee performances in smart factories for Industry 4.0 [5], body motion analysis in sports, and monitoring safety (falls, injuries, and collisions) [6,7] in an IoT environment. Activity recognition is a significant indicator of participation, quality of life, and lifestyle. Human activities carry a lot of information about the context (i.e., a person’s identity, personality, and mental state) and help systems to achieve context-awareness. For example, patient activity recognition is critical in analyzing treatment progress and can provide context information for decision-making for better treatment and care. Similarly, rehabilitation specialists and therapists can remotely benefit from information on patient activities outside of a health center. Having detected the activities, a broad range of analyses (i.e., activities by age group, by gender, by location, by day of the week) can be performed to answer the questions of where and when users perform which types of activities. It can help detect the abnormalities in surveillance systems, therefore preventing any unfavorable consequences. Using wearable sensors, HAR applications detect the actions of the users to provide them with intelligent personal assistance and recommendation. In the military, it is important to recognize the activities of the soldiers to provide feedback to their managers that assist them in real time. Consequently, there are numerous potential computing systems where recognizing human activities plays an important role.
One of the main problems associated with the current indoor HAR systems is that they have been considered as black-box systems, using nonunderstandable sensor data and providing predictions without being able to explain them. Without explainability and transparency, an HAR model is not trustworthy for making real-world decisions, especially the high-risk ones in the IoT systems. The main aim of this study is to provide an interpretable, basic, and reliable approach to indoor HAR problems.
The main contributions of this study can be summarized as sixfold. (i) It proposes a new approach, called Human Activity Recognition on Signal Images (HARSI), which converts the time-series data to signal images and feeds them into a CNN for the image classification task. (ii) It is the first attempt to combine four methodologies: signal image-based indoor HAR, IoT, explainable artificial intelligence (XAI), symmetry, and deep learning (DL). (iii) It provides an important contribution by improving human-level explainability for smart sensor data by using signal images in the field of indoor HAR. (iv) It takes into consideration symmetric human activities. (v) This study is also original in that it compares the performances of different nine CNN architectures on signal image data in terms of accuracy to determine the best one for indoor HAR. (vi) The proposed method outperformed both the classical machine learning methods (13.72% improvement) and the state-of-the-art methods (7.06% improvement) on the same dataset.
With quantitative evaluation in experiments, we demonstrated the effectiveness of the proposed HARSI approach on a real-world dataset. The experimental results showed that HARSI accurately recognized symmetric human activities, including walking, jogging, standing, sitting, moving downstairs, and moving upstairs. The results also showed that a significant improvement was achieved by the proposed HARSI method (98%) compared to the traditional machine learning methods (84.28%) and the state-of-the-art methods (90.94%) in terms of recognition accuracy.
The organization of the paper is as follows. Section 2 explains the recent previous studies on HAR that use a deep learning technique. Section 3 describes the proposed HARSI method in detail. Section 4 provides a brief description of the data and presents the experimental results. This section also gives debates on the subject and explains our solutions. Section 5 presents concluding remarks with the main findings and opportunities for further research.

2. Related Work

Research in the field of HAR is becoming increasingly important with the rapid development of smart sensor systems [8]. Especially, HAR is of great significance in the Internet of Things (IoT) applications, which include sensor and communication technologies. Different sensor types have been utilized in the HAR systems, including wearable sensors [9,10,11,12,13], vision-based sensors/cameras [14,15,16], health sensors [17], and environmental sensors [6,18,19]. The ambient sensor-based HAR applications detect activities from the sensors that are installed at fixed locations (i.e., home, factory) or placed on a fixed object (i.e., fridge, door, toilet flush). In this study, we focused on wearable sensors since they provide many advantages such as privacy protection, wide coverage area, and high robustness when modeling an activity classifier.
Wearable sensors are lightweight (few grams), small in size (few mm), easy to program, and low cost. Using either a strap or an adhesive, they can be easily attached to many different body parts (i.e., arm, waist, shoulder, wrist, or leg) depending on the human activities being studied [11,20]. A pair of sensors can be symmetrically located on a human body to collect synchronized measurements of them, allowing for the assessment of symmetric human behaviors. The accelerometer (A) [21,22,23,24] is one of the widely used sensors to collect acceleration data related to human activity. Besides the accelerometer, gyroscope (G) and magnetometer (M) sensors are attached to the body in various ways for monitoring actions at a particular point in time. The combination of the sensors (A, G, M) can also provide useful information when analyzed by machine learning methods to recognize human activities [25]. Most HAR systems [26,27] have been currently developed by smartphones; even other types of smart devices such as smartwatches, wristbands, and smart glasses have also been successfully employed.
Deep learning (DL) is one of the most exciting technologies that implements symmetry in computer science. In the literature, deep learning techniques have been proven to be powerful in classification. The recent studies [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32] on human activity recognition that use deep learning technology are given in Table 1. A variety of DL methods have been successfully used for HAR, such as recurrent neural networks (RNN) [24,26,28,29], long short-term memory (LSTM) [22,23,30], autoencoder (AE) [4,20], deep neural network (DNN) [1,9,13], and convolutional neural network (CNN) [31,32].
In the literature, most HAR systems [33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77] applied a supervised learning method; on the other hand, other types of machine learning, such as unsupervised or semisupervised learning, have also been investigated. The most widely used classification methods in HAR systems are decision trees [33,34,35,36,37,39,45,48,49], multilayer perceptron [33,34,41,46,48,49], support vector machines [12,35,37,38,39,40,42,44,45], naive Bayes [37,45,46], logistic regression [33,34,39,48,49], k-nearest neighbors [35,36,37,42,45], AdaBoost [47], and random forest [12,35,36,37,38,39,43,50].
In the literature, most HAR studies [12,13,21] focused on the identification of daily living activities such as standing, sitting, and walking. However, some previous studies tried to detect more specific types of activities such as cheating activity in an exam [14], rope jumping [25], shopping [30], housekeeping [18], hand-oriented activities (i.e., eating, clapping, writing) [78], virtual reality (VR) users’ activities (slash, thrust, guard) [79], and sports activities (i.e., basketball, bowling, boxing, and tennis) [16]. Besides these high-level activities, some works [2,22] focused on the transitions between the activities, such as sit-to-stand or stand-to-sit. In addition, recognizing group activities such as punching, kicking, and pushing were also investigated in previous studies [80]. In this study, we built a CNN model to recognize six different activities, including walking, jogging, standing, sitting, moving downstairs, and moving upstairs.
Typically, an HAR framework contains the following stages: data collection, data preprocessing, feature extraction, feature selection, training, performance evaluation, classification, and decision-making. Using raw sensor data directly in machine learning is usually not practical since it does not carry sufficient information to distinguish different human activities. In other words, only one particular value at a specific time instant of an action does not carry enough information to describe an activity itself. For this reason, the standard HAR studies involve the feature extraction phase to transform time-series data into samples that summarize the data over a particular time period. In general, frequency-domain features and time-domain features are extracted from the raw sensor data, such as max, min, median, mean, peak-to-peak value, standard deviation, the number of zero crossings, skewness, kurtosis, and signal entropy [10]. Among them, the skewness informs about the symmetry of the signal. In the literature, some studies [17,21,27] mainly concentrated on the feature extraction issue since it plays an important role in the final system performance.
HAR models can be considered in two categories: per-subject (personalized) and cross-subject (generalized) models. In per-subject models, both the training and testing data are all from the same individual, since each person has unique characteristics of movements, such as speed corresponding to its physical properties (i.e., age, gender, height, and weight) and habits. On the other hand, in cross-subject models, the training and testing data come from all the persons. They provide many advantages, such as working on a large amount of data to build a robust classifier and dealing with a single classifier instead of multiple ones.
Our study differs from the studies aforementioned here in several aspects. In our study, we do not use the features that were extracted from sensor data as most HAR systems do; rather, we transfer time-series data into signal images that reflect the properties of activities. Here, we present a detailed analysis of the performances of different CNN architectures on human activity signal image data for the first time. Our aim is to provide an explainable artificial intelligence (XAI) approach that can give human-understandable information and prediction in an IoT system. In other words, in this study, we provide human-level explainability for smart sensor data in the field of indoor HAR. To the best of our knowledge, five concepts together (signal image-based indoor HAR, XAI, IoT, symmetry, and DL) have not been studied so far.

3. Proposed Method

3.1. Problem Definition

Human activity recognition is a problem of classifying data obtained from sensors into well-defined actions performed by humans, e.g., walking, jogging, and sitting. Recognition of activities is a challenging time-dependent task since there is no single and precise way or formula to define specific movements. Machine learning methods have played a major role in the analysis of sensor data for providing real-time feedback in HAR applications.
In an HAR system, an accelerometer sensor embedded in an IoT device is placed at a specific location on the human body and is synchronized to emit data in an IoT environment. The accelerometer measures three-dimensional acceleration referenced with the Earth’s gravity during dynamic states. The sensor generates a signal along the x-axis, y-axis, and z-axis at a time step t ϵ {1,2,…,T}. Accelerometer measurements are collected in a time interval for a person. For this sensor, the sensing data along time can be represented by a multidimensional time series S = [s1, s2, …, st, …] where st is the sensing signal of the sensor placed at a body location at time t. Each signal record has six attributes {datetime, sensorID, x-axis, y-axis, z-axis, activity}.
A single measurement is not enough to classify an activity because of the time-dependent nature of activities. To deal with this problem, S is segmented into multiple frames, called windows, and then, each frame is mapped into a predefined activity label. In our study, temporal segmentation through the sliding window technique is necessary to define the boundaries of signal images. Each image is annotated by an activity from a label set, which is denoted by ai ϵ {A1, A2, …., Am}, where m is the number of potential activities to be recognized. For instance, the class labels can be as follows: A1 = walking, A2 = sitting, and A3 = stairs.
The activity recognition problem can be described as follows. Given a recorded signal time series S, the task is to detect an activity (e.g., walking) that infers human behavior in a time period. In the proposed approach, time-series data are converted to images by drawing three lines according to the recorded x–y–z values, and then, the images are fed into a CNN for the image classification task.
The concept of symmetry has been considered in many topics; similarly, it can be also discussed related to indoor human activity recognition. Figure 1 shows the example of symmetric and asymmetric activities with respect to the y-axis or arms/legs positions. Human activity such as walking can be considered as a symmetric movement depending on the biped’s parameters such as slope angles. Similarly, a jogging activity is also symmetric since the arm and legs are coordinated and moving together at the same frequency; i.e., the phase-plane cycles of the two legs are identical. Some group activities are also defined as symmetric, such as shaking hands and hugging. Similarly, the activity “WalkTogether” is symmetric because “I am walking together with you” is the same as “you are walking together with me”. On the other hand, some activities, such as kicking, falling down, pushing, picking up, and punching, can be categorized as asymmetric activities since two legs or two arms do not move simultaneously at the same angle or at the regions on symmetric sides.

3.2. Proposed Approach

In a traditional indoor HAR system, features are commonly extracted from time-series sensor data by using statistical methods such as max, min, mean, peak-to-peak value, standard deviation, the number of zero crossings, skewness, kurtosis, and entropy. However, this input data cannot be interpretable by humans, as can be seen in Figure 2. For humans, the numerical values such as in Figure 2 cannot be matched with the activities. For instance, when the numeric feature vector [339, 27, 0.3, 0.4, 0.08, 0.06, 0.05, 0.07, …] is seen by a human, it cannot be directly associated with the “walking” activity since it lacks visual representation.
To provide human-level explainability for smart sensor data, we propose an approach that visualizes the data with charts, as can be seen in Figure 3. Generating signal images makes data understandable for humans. Each chart can be easily associated with activities by humans. For example, Figure 3b corresponds to the “jogging” activity since both the amplitude and frequency of the signal are high as a result of the high velocity and displacement of the person on the ground.
Figure 3 shows sample signal images that include the x, y, and z axes values of an accelerometer sensor (Ax (red), Ay (green), and Az (blue)) over time for each activity. The activity images can be easily understandable, interpretable, and explainable by humans since each one has different characteristics. For example, walking is more periodic than standing. Jogging requires higher effort and power than walking since it requires more intense muscle contractions. It can also be seen that very low x–y–z-values are observed for sitting. Moreover, a human may require different relaxation times when moving upstairs, and sometimes, he/she stops to have a rest, moves slowly, and spends more time because he/she feels tired. The acceleration curve of moving downstairs is similar to moving upstairs, but the cycle of motion is shorter. These observations are also consistent with the energy harvesting from human activities. Sitting and standing activities are nonperiodic since they are stable and straightforward postures, while other activities are mainly repetitive and quasi-periodic. The walking activity is symmetric since the human legs are coordinated and moving together at the same frequency and the phase-plane cycles of the two legs are the same. Signal data show a symmetric pattern for some human activities. For jogging activity, the signal amplitude is symmetric with the zero axis. By using the differences in the figures, a CNN algorithm can successfully distinguish the activities.
In this study, we do not extract features from raw sensor data as the traditional HAR studies do. Instead, we assemble x-axis, y-axis, and z-axis accelerometer signal sequences into an image to enable CNNs to learn the optimal features automatically from the signal image for the human activity classification task. In other words, we propose an approach, called HARSI, which transforms numerical sensor data into image format data and builds a CNN model that enables human activity recognition on these signal images. The main purpose of our study is to provide an interpretable and robust approach to indoor HAR problems.
Figure 4 shows a general overview of the proposed HARSI approach. The approach mainly includes the following stages: data collection, data transformation, training, testing, and classification. (i) The data collection stage comprises obtaining raw signal values via an accelerometer sensor available in an IoT environment when performing human actions. After that, the collected raw data are transferred to a server through WiFi communication technology. (ii) In the data transformation stage, the time-series signal data are divided into fixed-size segments, called windows, by using a sliding-window method. After that, an image is generated for each window by drawing lines on sample values. Here, each signal image corresponds to a single activity such as standing, sitting, or walking. (iii) In the training stage, each signal image is fed to CNN as an input vector, and then CNN learns different features of the image through different layers. (iv) In the classification stage, the CNN model gives insight according to the features of the signal image. In other words, the CNN model makes a prediction for a given input image according to the class probabilities of activities, such as standing, sitting, jogging, walking, moving downstairs, and moving upstairs. (v) In the testing stage, the performance of the CNN model is assessed by using a test set to evaluate how well it recognizes human activities. If the prediction accuracy of the CNN model is at an acceptable level (i.e., >80%), it can be further used to recognize real-time human activities. Afterward, the final prediction can be considered in a decision support stage to provide guidance to the decision-maker. Since the indoor environment of HAR systems is dynamic and ever-evolving, it is required to update the model periodically by following the same stages to achieve high accuracy consistently.
As seen in Figure 4, the CNN contains an input layer, multiple hidden layers, and an output layer. Signal image data are processed layer-by-layer, where the output of each layer becomes the input for the next layer. Each layer contains multiple units, which are denoted by U i l to indicate the ith unit in layer l. The hidden layers are composed of convolutional, pooling, and fully connected layers.
Convolutional layer (CL): These are used as feature extractors to automatically obtain high-level representations of input images. Formally, a feature map is extracted using a convolution procedure, as follows:
F j l + 1 = α i = 1 F l K j , i l F j l + b j l
where F j l denotes the jth feature map in layer 𝑙, | F l | is the number of feature maps in layer 𝑙, α() is an activation function, b j l is a bias vector, and K j , i l represents the kernel applied on feature map i in layer 𝑙 to obtain jth feature map in layer (𝑙 + 1).
Pooling layer (PL): Pooling layers are used to reduce dimensionality, as well as the number of parameters. A PL is usually inserted between successive CLs in a CNN architecture. Formally, max pooling is given by:
v i l + 1 = max 1 k r v i + k l
where 𝑟 is the pooling size and v i l refers to the value of the 𝑖th unit in layer 𝑙.
Fully connected layer (FCL): After multiple CLs and PLs, the classification process is handled in a fully connected layer, which produces an output vector, as given in Equation (3).
z l + 1 = v i l w
where w represents a weight vector and vector z includes nonnormalized log probabilities. The output of the FCL is fed into a softmax classifier, which predicts the activity label as follows:
S o f t m a x z i = P o = a   |   z i = e z i j = 1 m e z j
Where a is an activity label, o denotes the output of the classification model, z j represents the jth element of log probability vector 𝘇, and m is the number of class labels. The predicted activity label (al) for a given image is assigned to the one with the highest probability, as given in Equation (5).
a l argmax a = 1 m P ( o = a   |   i m a g e )

3.3. Formal Definition

Let the raw dataset D be a set of instances collected by an accelerometer sensor. Each instance in dataset D includes a set of pairs of x–y–z axis values and the corresponding activity label, which is denoted by D = {(x1, y1, z1, a1), (x2, y2, z2, a2),…., (xn, yn, zn, an)}, where n is the number of instances. In other words, ai is the activity (class label) belonging to the axes values of the sensor (xi, yi, zi). The output attribute O = {a1, a2, …, an} has m different human activities, which is denoted by ai ϵ {A1, A2, …., Am} for i = 1,2,…,n. For example, in a four-activity classification (sitting, standing, stairs, walking), the class labels of the instances are A1 = sitting, A2 = standing, A3 = stairs, and A4 = walking.
In the proposed HARSI approach, the raw sensor data D are transformed into signal images using a sliding window method. In this process, a large time-series dataset is split into fixed-sized chunks, referred to as windows, denoted as W = (w1, w2, …, wn/q), where q is the window size.
Definition 1 (window).
A window is defined as a set of consecutive sensor measurements obtained within a time interval such that w = {sr, sr+1, …, sr+q-1}, where q refers to the window size and r corresponds to an arbitrary position, such as 1 ≤ r ≤ n-q+1, where n is the data size.
After generating windows, a single activity label ai ϵ {A1, A2, …., Am} is assigned to each window {(w1, ai), (w2, ai),…, (wn/q, ai)} such that all the samples within the window belong to the respective class. After that, an image is generated for each window by drawing lines on sample values. Each signal image is labeled with the corresponding activity a.
Definition 2 (activity).
An activity is a human movement characterized by a body action or posture, e.g., walking. An activity label aiϵ {A1, A2,…., Am} is associated with an image that is generated from a window with fixed length (q) by segmenting the raw sensor data D, where m is the number of potential activities to be recognized.
The problem studied in this work is to detect a corresponding activity implicated in a certain temporal sequence based on the classification. In other words, the aim of HAR is to build a model M(image, •) to infer the correct activity label for a given image, where • denotes all the parameters to be learned during the training process.
Definition 3 (activity recognition task).
Given a set of training images with their corresponding activity labels and a query image, the aim is to find a mapping function𝑓: image → activity that correctly infers the human behavior for the query image. The predicted activity label should be as similar as possible to the actual class label. Therefore, the task is to build a classification model M by minimizing total loss L(M).
It should be noted that a sliding-window method can be performed in either an overlapping or nonoverlapping way. A nonoverlapping method indicates that the values in one image do not intersect with the values of the other successive image, i.e., w1w2 = ∅. On the other hand, an overlapping method is defined by a particular percentage, which indicates how many samples from the previous image are repeated in the current image, i.e., w1w2 ≠ ∅. In this study, we prefer to use the nonoverlapping image technique to prevent information duplication.
A crucial factor in a sliding window method is to select a suitable window size to achieve high recognition accuracy since the ideal window size varies in accordance with the characteristics of signals being processed. In general, a small window size can be useful to detect faster-changing activities better; however, using short windows may lead to misclassification because some vital information about a complex activity may not be captured by multiple windows. On the other hand, large windows can detect complex activities and semi-complex activities. However, a large window generates a signal image that belongs to more than one human activity, and this leads to a decrease in recognition accuracy. Considering this tradeoff, researchers usually determine the optimal window size by trying empirical values and assessing classification accuracy. In this study, the size of each window was set to 100 samples since the dataset was collected at a rate of 20 samples per second, and it is a sufficient sampling value to make a reasonable prediction for human activity.
Algorithm 1 shows the pseudocode of the proposed HARSI method. In the algorithm, first, a sliding window technique is used to split accelerometer sensor data (D) into windows with size q. In other words, it segments data streams into windows of equal length. For a dataset with n samples, the algorithm generates n/q windows, where each one ranges between I × q and I × q + q − 1 for i = 0, 1, …, n/q. After that, an image is generated for each window by drawing lines on sample values. Here, a small window is shifted along the continuous data stream, converting contiguous portions of sensor readings into images. Each image is labeled with the corresponding activity. A CNN classifier M is then trained on the signal image dataset. In the final step, the activity (class label) of each unseen image in the test set T is predicted by using the classifier.
Algorithm 1. Human Activity Recognition on Signal Images (HARSI)
Inputs:D = {(x1, y1, z1, a1), (x2, y2, z2, a2),…., (xn, yn, zn, an)}
     q: window size
     T: Test set
Output: O = {o1, o2, …, ot} a set of outputs for test images
Begin:
     for i = 0 to n/q do
      W = Ø
      for j = i *q to i * q + q − 1 do
       W = W U (xj, yj, zj,)
       activity = aj
      end for
      image = ConvertToImage(W)
      I = I U <image, activity>
     end for
     M = CNN(I)
     foreach image i in T do
      o = Classify(M, i)
      O = O U o 
     end foreach
     Return O
End

4. Experimental Studies

This section presents a detailed study that was carried out to evaluate the performance of the proposed HARSI method. The effectiveness of the method was demonstrated on a real-world dataset by using different CNN architectures, including AlexNet, ResNet, SqueezeNet, DenseNet, and VGG. The parameter settings and the number of parameters for each CNN architecture are given in Table 2. The structures of CNNs are different from each other in several aspects, such as the number of parameters and the number of layers. For instance, ResNet34 consists of a residual network with 34 layers. Compared to ResNet50, the DenseNet121 has more layers, whilst VGG19 has fewer layers. Some parameter settings are common in all models, i.e., rectified linear unit (ReLU) was used as activation function and the output probability was calculated by softmax. As the nature of the models used, batch layers predate each ReLU. In all models, Adam and cross entropy were used as the optimizer and loss function, respectively. These techniques have lately gained popularity and performed promising outcomes for deep learning applications.
In this study, optimal learning rate parameters were determined for each model separately to speed up the training process, adapt itself to the problem, and strengthen the generalization ability of the classifiers. While a low learning rate slows the convergence of the training process, a high learning rate can cause an unpleasant divergence in performance. Therefore, a suitable learning rate is vital for obtaining a satisfactory performance; however, finding an appropriate learning rate is both laborious and hard to decide. To solve this problem, we used the lr_find() method in Fast.AI, which is a deep learning library built on top of PyTorch. This method works on the principle of using a very low learning rate initially to train a minibatch and calculate the loss. In the next step, the method trains the next minibatch with a small-scale higher learning rate than the previous one until it finds a learning rate where the model diverges. The optimal learning rate values determined for each model are listed in Table 2.
One of the main differences between the CNN models is the number of parameters, which can reflect the computational complexity of the model. It may be noted that as the number of total parameters increases, the time required for training usually increases. There are two categories of parameters: one is trainable parameters (i.e., weights of connections between layers) that are continuously updated to reduce the loss, and the other one is nontrainable parameters (i.e., biases) that are not optimized during the training process. For instance, in VGG19 architecture, the number of trainable parameters is 541,440, while the number of nontrainable parameters is approximately 20 million. As seen in Table 2, SqueezeNet has the smallest total number of parameters, whilst ResNet101 has the largest. The architectures have approximately 2, 15, and 25 million parameters for AlexNet, VGG16, and ResNet50, respectively. These values may be associated with computational complexity, where the higher the number of parameters, the greater the computational load during the training process.
The method was implemented in Python by using the PyTorch framework and various libraries such as Fastai, NumPy, Pandas, Scikit-Learn, Matplotlib, and Seaborn. In this study, the CNN models were trained on a computer equipped with an Nvidia GTX 1060 graphics card using the Cuda toolkit in order to make use of the GPU computational capability and reduce implementation time through a rapid and simple design.
To demonstrate the superiority of the proposed approach (HARSI) over the previous approaches, we compared it with the classical machine learning methods (SVM, DT, NB, KNN, MLP, AdaBoost, and RF) [12,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50] and also compared it with the state-of-the-art methods [5,12,13,24,26,34,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77] on the same dataset.
In this study, we split the dataset into two subsets: 80% of the data were used for training and the remaining 20% were used for testing. This standard split approach was chosen since it is common in the previous studies [5,71,72,74] that used the same dataset. In order to provide comparability with the literature, the same split approach was preferred. In addition, the training part of the data was divided into training and validation sets as 80% and 20%, respectively. The test set contains 100 images from each category; therefore, it includes 600 images in total. Four different metrics were used to evaluate the performance of each CNN architecture: accuracy, recall, precision, and f-measure. Accuracy is the fraction of correct predictions of the model to total prediction. Equation (6) shows how the accuracy rate is calculated.
A c c u r a c y = T P + T N T P + F N + F P + T N
where TP is true positive, FP is false positive, TN is true negative, and FN is false negative. Precision describes how precise the model is out of the samples predicted positive, and how many of them are actually positive. Recall indicates how many of the actual positives the model captures through classifying them as positive. F-measure offers a single score that balances both the concerns of recall and precision values. Equations (7)–(9) show the calculations of precision, recall, and f-measure values, respectively.
P r e c i s i o n =     T P T P + F P
R e c a l l   =     T P T P + F N
F m e a s u r e = 2 × p r e c i s i o n × r e c a l l p r e c i s i o n + r e c a l l

4.1. Dataset Description

In order to show the effectiveness of the proposed approach, a publicly available dataset, named WISDM (Wireless Sensor Data Mining) dataset [33], was used in the experiments. The dataset is available at the website https://www.cis.fordham.edu/wisdm/dataset.php (accessed on 25 September 2022). It was released by the Laboratory at Fordham University in the United States. This dataset is one of the important and popular large-scale benchmark datasets in the field of HAR. It has been used in many studies [5,12,13,24,26,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77], so it is suitable for making comparisons with previous works. The dataset is appropriate for detecting symmetric activities. Before collecting the data, the researchers obtained approval from the University Board since it involved research on human subjects and involved some risks, i.e., the subject could fall down while jogging. The data collection process was fully monitored and guided by researchers in the laboratory environment to ensure the quality of the data. The dataset contains routine motion patterns with a significant number of processable movement samples. The dataset has 1,098,207 samples that were collected from 36 different participants while performing six activities. Therefore, it could possibly be used to analyze the movement behaviors of different persons. The percentages of each activity in the dataset are as follows: jogging 31.2%, moving downstairs 9.1%, walking 38.6%, moving upstairs 11.2%, standing 4.4%, and sitting 5.5%. There is no missing value in the dataset. While collecting data, the participants are requested to carry an accelerometer sensor in their front pockets. They were asked to jog, walk, descend stairs, ascend stairs, stand, and sit for specific periods of time. With this experiment setup, accelerometer data were retrieved at every 50 ms, which means 20 samples per second, while participants were performing activities. The raw dataset consists of x-, y-, and z-axis values obtained by an accelerometer sensor embedded in a smart device.
In this study, image representations were first generated from the raw x, y, and z values in the accelerometer sensor data. In other words, we converted the time-series data into signal images by drawing three lines on sample x, y, and z values. Here, a small window was shifted along the continuous data stream converting contiguous portions of sensor readings into images. In the generated images, x, y, and z drawing lines are represented with different colors; red, green, and blue, respectively. While generating charts from sensor data values, attention was paid to choosing a fixed range for all the graphs. Accordingly, the max and min values were found by searching the x, y, and z axes values in the dataset. The vertical range of the graph in each image approximately lies between [–20, 20]. The horizontal range of the graph was set to 100 samples since the dataset was collected at a rate of 20 samples per second, and therefore it is a sufficient range to make a reasonable prediction for activity. Each image is labeled with the corresponding activity, such as standing, sitting, or walking. To create a balanced dataset, 400 images were generated for each activity; therefore, in total, 2400 images were generated. Figure 3 shows sample signal images for each activity.
By transforming numerical sensor data into image data, we aim to improve both explainability and recognition accuracy. The generated images provide human-level explainability for smart sensor data. Since each image reflects the properties of activities, they can be easily interpretable by humans. Rather than solving a time-series data classification problem, we define the HAR problem as an image classification problem. In this way, we provide an interpretable and robust approach to HAR problems.

4.2. Comparison of Different CNN Architectures

On the dataset described in the previous section, the effectiveness of the proposed approach (HARSI) was demonstrated by using different CNN architectures, including Alex Network (AlexNet), Residual Network (ResNet), Visual Geometry Group (VGG) Network, SqueezeNet, and Dense Convolutional Network (DenseNet). These CNN architectures were selected because of their popularity, high robustness, proven efficiency, and ability in image classification. They automatically extract features of images that are useful in the identification of human activities. They use a gradient descent algorithm to optimize the CNN parameters.
Table 3 shows the performance of the proposed HARSI method on different CNN architectures for the same dataset. Based on the accuracy rates, it is possible to say that all the CNN models have good classification ability. However, VGG19 is the most successful model among them with a 98% of success rate. Following this, ResNet34 has also a high accuracy rate (97.33%) in distinguishing human activities. This success is the result of the strengths of CNNs in classifying images. CNNs are capable of extracting key features directly and effectively from images, learning useful information layer-by-layer, and successfully classifying them into different classes.
In addition to accuracy, we also evaluated the performance of the proposed HARSI approach on different CNN architectures in terms of recall, precision, and f-measure metrics. The values of these metrics range between 0 and 1, where 1 is the best value. As can be seen in Table 3, the recall value obtained by the VGG19 model is closer to 1 than the others. This means that the VGG19 model often tends to give better predictions than the rest. As can be observed, the VGG16 model also outperformed the others in terms of precision and f-measure.
In Figure 5, the loss values in both the training and validation processes are shown. While the vertical axis indicates the loss value, the horizontal axis represents the number of batches processed. In initial batches, the training loss is higher than the validation loss. As can be seen, both the training and validation losses reduce with the increase of the batches. The training loss and validation loss converged after approximately 200 batches were processed. When the minimum validation loss was obtained, the training process was stopped to avoid overfitting.
Figure 6 presents the confusion matrix to show the predictive performance of the proposed HARSI method on each human activity separately. The rows in the confusion matrix represent the predicted activity labels, whereas the columns represent the actual activity labels. Each cell in the matrix is a percentage value, indicating what percent of the data belongs to the column class but is incorrectly classified as the row class. All correctly classified samples are positioned on the diagonal of a confusion matrix, so, its diagonal should contain the highest values possible, and all the other elements should be close to zero. According to the matrix given in Figure 6, it is possible to say that the model usually had no difficulty in distinguishing human activities. For example, 98 out of 100 walking activities were predicted correctly; however, only two of the walking activities were misclassified by the classifier. Although each activity was recognized with a high accuracy rate, downstairs and upstairs activities were slightly confused with each other since they are similar activities. The algorithm produced an equal accuracy value (95%) for moving upstairs and moving downstairs activities since they have similar characteristics to others. It can be concluded from the confusion matrix that the best accuracies were achieved in the sitting, standing, and jogging activities.
Figure 7 shows the execution times of the proposed HARSI method on different CNN architectures in minutes. Although all the times are close to each other, AlexNet and SqueezeNet are the fastest ones among their counterparts. They are followed by the ResNet34 model (1.11 min). The VGG models are also efficient in terms of training time (1.19 and 1.24 min). The DenseNet model may take a longer time (1.28 min), especially handling large image datasets. This is probably because of the higher number of layers in its architecture. Similarly, the required time for training ResNet101 is higher than others since it has a higher number of parameters to be assessed. The size and resolution of the images are also factors that affect computation time. When the sizes of the images are reduced by the resizing process, the time required for analyzing them decreases, and therefore, the performance of the HAR system is positively affected.

4.3. Comparison with the Classical Machine Learning Methods

In order to show the superiority of our method, we compared it with the classical machine learning methods such as multilayer perceptron (MLP), support vector machines (SVM), decision tree (DT), naive Bayes (NB), logistic regression (LR), k-nearest neighbors (KNN), AdaBoost, and random forest (RF). In order to make a plausible comparison, the most important factor is to use the same data. Therefore, the results obtained by the classical machine learning methods on the same dataset [33] were used in the comparison. Table 4 lists the related studies with their methods and the corresponding accuracy rates. It can be seen from the table that the proposed HARSI method outperformed the other methods [12,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50] with a 13.72% improvement on average. Employing HARSI achieved higher accuracy (98%) than the traditional machine learning models on the same dataset.

4.4. Comparison with the State-of-the-Art Methods

This section presents comparative results which highlight the performance of the proposed method over the state-of-the-art methods in the literature. Table 5 shows the performance improvement of our method over the state-of-the-art methods [5,12,13,24,26,34,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77]. The results were taken directly from the referenced studies since the researchers used the same dataset [33] as our study. It can be seen from the table that the proposed HARSI method outperformed the other methods with a 7.06% improvement on average.

4.5. Discussion

The main debates in the field of HAR and our solutions can be summarized as follows.
  • In HAR, the ideal input data format is still a subject of much debate and there are various ongoing works for improving the accuracy of the models. Traditional HAR has been defined as a time-series data classification problem and requires feature extraction. In contrast, we transfer time-series data into signal images that reflect the properties of human activities. It avoids the need to perform an explicit feature generation and selection stage. We improved accuracy by working on signal image data, instead of numerical time-series data.
  • Many applications in HAR [33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50] have used classical machine learning methods such as DT, SVM, MLP, NB, LR, KNN, and RF. However, the performance of these methods is still highly debated. In this study, we take advantage of the strengths of deep learning approaches.
  • Another debate is how to design CNN architecture to be able to obtain good performance. For example, the number of layers and parameter settings are still subjects of much debate. In this study, we compared nine different CNN architectures to determine the best suitable one.
  • In the activity recognition community, there is an open debate on providing explainability in the HAR systems. The main problem is how to increase the transparency and interpretability of the models. In this study, to increase human-level explainability, we visualize the data with charts since generating signal images makes data understandable for humans.
  • Another ongoing debate is which activities can be predicted more precisely. This study showed that the best accuracies were achieved in the sitting, standing, and jogging activities due to their diverse natures.
  • The proposed HAR model can be connected to many different fields of study such as health monitoring, fitness tracking, home and work automation, and self-managing system. With the rapid technological developments in smartphones, the model can enable new opportunities for developing informative systems on a large scale to perceive and act on what users (i.e., your children, elderly mother, or sick family member) are doing. Recognizing human activities is important for the treatment of patients and can provide useful feedback to the clinicians since the activity is associated with health. For example, it can be used to monitor patients in rehabilitation since the functional status of a person is an important parameter in this area. In addition, it could be used to offer activity-aware services to smartphone users, such as movement recommendations. A number of lifestyle diseases and movement disorders are associated with inactivity; therefore, our model can be used to give information to prevent diseases. The users can participate in the tracking of their activities for the sake of health, fitness, or other purposes due to its strength in providing personalized support.

5. Conclusions and Future Works

Classical HAR has been defined as a standard data classification problem and extracts statistical features (i.e., min, max, skewness, kurtosis) from data, which cannot be readable and interpretable by humans. Transparent and explainable indoor HAR systems are required to generate human-understandable information. For this purpose, an approach, called Human Activity Recognition on Signal Images (HARSI), is proposed in this study. The proposed approach creates image representations of the time-series sensor data to improve both explainability and recognition accuracy. This is the first attempt to combine five methodologies: signal image-based indoor HAR, XAI, IoT, symmetry, and DL. It takes advantage of the strengths of CNNs in handling signal image data. In the experimental studies, we demonstrated the effectiveness of the proposed HARSI approach compared to the previous studies on a real-world dataset.
The main findings of the study can be concluded as follows:
  • The proposed approach improves human-level explainability for smart sensor data by using signal images in the field of HAR.
  • The proposed HARSI approach improves the recognition accuracy in the HAR problems by converting time-series data to image data.
  • The experimental results showed that HARSI successfully (98%) recognized six symmetric human activities, including walking, jogging, standing, sitting, moving downstairs, and moving upstairs.
  • According to the experimental results, it can be concluded that the best suitable and consistent CNN model for the WISDM dataset is VGG19. It achieved the best results on all the metrics (accuracy, precision, recall, and f-measure). Therefore, this model can be successfully used to identify human activities.
  • The prediction accuracy changes according to human activities. Among the activities, sitting, standing, and jogging were correctly predicted by the proposed method. On the other hand, the model had a little difficulty in classifying downstairs and upstairs activities with an accuracy of 95% for the WISDM dataset.
  • The number of layers and number of parameters of a CNN model may be associated with computational complexity, where the higher the number of layers and parameters, the greater the computational load during the training process.
  • A significant improvement (13.72% on average) was achieved by the proposed HARSI model compared to the classical machine learning methods such as KNN, DT, SVM, NB, LR, MLP, AdaBoost, and RF.
  • Our approach achieved higher classification accuracy than the state-of-the-art approaches. It outperformed them by 7.06% on average on the same dataset.
  • The proposed HARSI approach has the potential to expand the application of machine learning in many different sectors, thanks to its advantages.
One limitation of this study is related to sensors such as signal delays, noises, damages, battery capacity, and shelf-life. However, this limitation is also valid for all other wearable sensor-based HAR applications. It can be overcome in the future with developments in sensor technology. Another limitation is that it focuses on single-person activity detection. In the future, we plan to adapt it for recognizing group activities such as handshaking and hugging.

Author Contributions

Conceptualization, K.U.B. and M.C.; methodology, K.B.; software, A.B.C.; validation, A.B.C.; formal analysis, D.B.; investigation, K.U.B., M.C. and K.B.; data curation, A.B.C.; writing—original draft preparation, D.B.; writing—review and editing, A.B.C.; visualization, A.B.C.; supervision, D.B.; funding acquisition, K.U.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The “WISDM (Wireless Sensor Data Mining)” dataset [33] is publicly available at the following website: https://www.cis.fordham.edu/wisdm/dataset.php (accessed on 20 August 2022).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hassan, M.M.; Ullah, S.; Hossain, M.S.; Alelaiwi, A. An end-to-end deep learning model for human activity recognition from highly sparse body sensor data in internet of medical things environment. J. Supercomput. 2021, 77, 2237–2250. [Google Scholar] [CrossRef]
  2. Kanjilal, R.; Uysal, I. The future of human activity recognition: Deep learning or feature engineering? Neural Process. Lett. 2021, 53, 561–579. [Google Scholar] [CrossRef]
  3. Mekruksavanich, S.; Jitpattanakul, A. Biometric user identification based on human activity recognition using wearable sensors: An experiment using deep learning models. Electronics 2021, 10, 308. [Google Scholar] [CrossRef]
  4. Mihoub, A. A deep learning-based framework for human activity recognition in smart homes. Mob. Inf. Syst. 2021, 11, 6961343. [Google Scholar] [CrossRef]
  5. Mohsen, S.; Elkaseer, A.; Scholz, S.G. Industry 4.0-oriented deep learning models for human activity recognition. IEEE Access 2021, 9, 150508–150521. [Google Scholar] [CrossRef]
  6. Madokoro, H.; Nix, S.; Woo, H.; Sato, K. A mini-survey and feasibility study of deep-learning-based human activity recognition from slight feature signals obtained using privacy-aware environmental sensors. Appl. Sci. 2021, 11, 11807. [Google Scholar] [CrossRef]
  7. Casilari, E.; Álvarez-Marco, M.; García-Lagos, F. A study of the use of gyroscope measurements in wearable fall detection systems. Symmetry 2020, 12, 649. [Google Scholar] [CrossRef]
  8. Shalaby, E.; ElShennawy, N.; Sarhan, A. Utilizing deep learning models in CSI-based human activity recognition. Neural Comput. Appl. 2022, 34, 5993–6010. [Google Scholar] [CrossRef]
  9. Bijalwan, V.; Semwal, V.B.; Gupta, V. Wearable sensor-based pattern mining for human activity recognition: Deep learning approach. Ind. Robot-Int. J. Robot Res. Appl. 2021, 49, 21–33. [Google Scholar] [CrossRef]
  10. Ferrari, A.; Micucci, D.; Mobilio, M.; Napoletano, P. Deep learning and model personalization in sensor-based human activity recognition. J. Reliab. Intell. Environ. 2022, 2022, 1–13. [Google Scholar] [CrossRef]
  11. Manjarrés, J.; Lan, G.; Gorlatova, M.; Hassan, M.; Pardo, M. Deep learning for detecting human activities from piezoelectric-based kinetic energy signals. IEEE Internet Things J. 2022, 9, 7545–7558. [Google Scholar] [CrossRef]
  12. Bhat, O.; Khan, D.A. Evaluation of deep learning model for human activity recognition. Evol. Syst. 2022, 13, 159–168. [Google Scholar] [CrossRef]
  13. Bozkurt, F. A comparative study on classifying human activities using classical machine and deep learning methods. Arab. J. Sci. Eng. 2022, 47, 1507–1521. [Google Scholar] [CrossRef]
  14. Khan, A.R.; Saba, T.; Khan, M.Z.; Fati, S.M.; Khan, M.U.G. Classification of human’s activities from gesture recognition in live videos using deep learning. Concurr. Computat. Pract. Exper. 2022, 34, e6825. [Google Scholar] [CrossRef]
  15. Khan, I.U.; Afzal, S.; Lee, J.W. Human activity recognition via hybrid deep learning based model. Sensors 2022, 22, 323. [Google Scholar] [CrossRef]
  16. Tasnim, N.; Islam, M.K.; Baek, J.-H. Deep learning based human activity recognition using spatio-temporal image formation of skeleton joints. Appl. Sci. 2021, 11, 2675. [Google Scholar] [CrossRef]
  17. Maitre, J.; Bouchard, K.; Gaboury, S. Alternative deep learning architectures for feature-level fusion in human activity recognition. Mob. Netw. Appl. 2021, 26, 2076–2086. [Google Scholar] [CrossRef]
  18. Hwang, Y.M.; Park, S.; Lee, H.O.; Ko, S.-K.; Lee, B.-T. Deep Learning for human activity recognition based on causality feature extraction. IEEE Access 2021, 9, 112257–112275. [Google Scholar] [CrossRef]
  19. Ronald, M.; Poulose, A.; Han, D.S. iSPLInception: An inception-ResNet deep learning architecture for human activity recognition. IEEE Access 2021, 9, 68985–69001. [Google Scholar] [CrossRef]
  20. Pei, L.; Xia, S.; Chu, L.; Xiao, F.; Wu, Q.; Yu, W.; Qiu, R. MARS: Mixed virtual and real wearable sensors for human activity recognition with multidomain deep learning model. IEEE Internet Things J. 2021, 8, 9383–9396. [Google Scholar] [CrossRef]
  21. Yen, C.-T.; Liao, J.-X.; Huang, Y.-K. Feature fusion of a deep-learning algorithm into wearable sensor devices for human activity recognition. Sensors 2021, 21, 8294. [Google Scholar] [CrossRef]
  22. Irfan, S.; Anjum, N.; Masood, N.; Khattak, A.S.; Ramzan, N. A novel hybrid deep learning model for human activity recognition based on transitional activities. Sensors 2021, 21, 8227. [Google Scholar] [CrossRef]
  23. Al-Wesabi, F.N.; Albraikan, A.A.; Hilal, A.M.; Al-Shargabi, A.A.; Alhazbi, S.; Duhayyim, M.A.; Rizwanullah, M.; Hamza, M.A. Design of optimal deep learning based human activity recognition on sensor enabled internet of things environment. IEEE Access 2021, 9, 143988–143996. [Google Scholar] [CrossRef]
  24. Alawneh, L.; Alsarhan, T.; Al-Zinati, M.; Al-Ayyoub, M.; Jararweh, Y.; Lu, H. Enhancing human activity recognition using deep learning and time series augmented data. J. Ambient Intell. Humaniz. Comput. 2021, 12, 10565–10580. [Google Scholar] [CrossRef]
  25. Elsts, A.; McConvill, R. Are microcontrollers ready for deep learning-based human activity recognition? Electronics 2021, 10, 2640. [Google Scholar] [CrossRef]
  26. Ghate, V.; Hemalatha, S.C. Hybrid deep learning approaches for smartphone sensor-based human activity recognition. Multimed. Tools Appl. 2021, 80, 35585–35604. [Google Scholar] [CrossRef]
  27. Thakur, D.; Biswas, S. Feature fusion using deep learning for smartphone based human activity recognition. Int. J. Inf. Tecnol. 2021, 13, 1615–1624. [Google Scholar] [CrossRef]
  28. Buffelli, D.; Vandin, F. Attention-based deep learning framework for human activity recognition with user adaptation. IEEE Sens. J. 2021, 21, 13474–13483. [Google Scholar] [CrossRef]
  29. Alhersh, T.; Stuckenschmidt, H.; Rehman, A.U.; Belhaouari, S.B. Learning human activity from visual data using deep learning. IEEE Access 2021, 9, 106245–106253. [Google Scholar] [CrossRef]
  30. Chen, L.; Liu, X.; Peng, L.; Wu, M. Deep learning based multimodal complex human activity recognition using wearable devices. Appl. Intell. 2021, 51, 4029–4042. [Google Scholar] [CrossRef]
  31. Thu, N.H.T.; Han, D.S. HiHAR: A hierarchical hybrid deep learning architecture for wearable sensor-based human activity recognition. IEEE Access 2021, 9, 145271–145281. [Google Scholar] [CrossRef]
  32. Stuart, M.; Manic, M. Deep learning shared bandpass filters for resource-constrained human activity recognition. IEEE Access 2021, 9, 39089–39097. [Google Scholar] [CrossRef]
  33. Kwapisz, J.R.; Weiss, G.M.; Moore, S.A. Activity recognition using cell phone accelerometers. ACM SIGKDD Explor. Newsl. 2010, 12, 74–82. [Google Scholar] [CrossRef]
  34. Suwannarat, K.; Kurdthongmee, W. Optimization of deep neural network-based human activity recognition for a wearable device. Heliyon 2021, 7, e07797. [Google Scholar] [CrossRef]
  35. Vijayvargiya, A.; Kumari, N.; Gupta, P.; Kumar, R. Implementation of machine learning algorithms for human activity recognition. In Proceedings of the 3rd International Conference on Signal Processing and Communication (ICPSC), Coimbatore, India, 13–14 May 2021; pp. 440–444. [Google Scholar]
  36. Semwal, V.B.; Lalwani, P.; Mishra, M.K.; Bijalwan, V.; Chadha, J.S. An optimized feature selection using bio-geography optimization technique for human walking activities recognition. Computing 2021, 103, 2893–2914. [Google Scholar] [CrossRef]
  37. Kee, Y.J.; Zainudin, M.S.; Idris, M.I.; Ramlee, R.H.; Kamarudin, M.R. Activity recognition on subject independent using machine learning. Cybern. Inf. Technol. 2020, 20, 64–74. [Google Scholar] [CrossRef]
  38. Jalal, A.; Quaid, M.A.K.; Kim, K. A study of accelerometer and gyroscope measurements in physical life-log activities detection systems. Sensors 2020, 20, 6670. [Google Scholar] [CrossRef]
  39. Khare, S.; Sarkar, S.; Totaro, M. Comparison of sensor-based datasets for human activity recognition in wearable IoT. In Proceedings of the IEEE 6th World Forum on Internet of Things (WF-IoT), New Orleans, LA, USA, 2–16 June 2020; pp. 1–6. [Google Scholar]
  40. Arigbabu, O.A. Entropy decision fusion for smartphone sensor based human activity recognition. arXiv 2021, arXiv:2006.00367v1. [Google Scholar]
  41. Xu, H.; Li, J.; Yuan, H.; Liu, Q.; Fan, S.; Li, T.; Sun, X. Human activity recognition based on gramian angular field and deep convolutional neural network. IEEE Access 2022, 8, 199393–199405. [Google Scholar] [CrossRef]
  42. Lu, W.; Fan, F.; Chu, J.; Jing, P.; Yuting, S. Wearable computing for internet of things: A discriminant approach for human activity recognition. IEEE Internet Things J. 2019, 6, 2749–2759. [Google Scholar] [CrossRef]
  43. Ignatov, A. Real-time human activity recognition from accelerometer data using convolutional neural networks. Appl. Soft. Comput. 2018, 62, 915–922. [Google Scholar] [CrossRef]
  44. Xu, W.; Pang, Y.; Yang, Y.; Liu, Y. Human activity recognition based on convolutional neural network. In Proceedings of the 24th International Conference on Pattern Recognition, Beijing, China, 20–24 August 2018; pp. 165–170. [Google Scholar]
  45. Quispe, K.G.M.; Lima, W.S.; Batista, D.M.; Souto, E. MBOSS: A symbolic representation of human activity recognition using mobile sensors. Sensors 2018, 18, 4354. [Google Scholar] [CrossRef] [PubMed]
  46. Azmi, M.S.M.; Sulaiman, M.N. Accelerator-Based human activity recognition using voting technique with NBTREE and MLP classifiers. Int. J. Adv. Sci. Eng. Inf. Technol. 2017, 7, 146–152. [Google Scholar] [CrossRef]
  47. Walse, K.H.; Dharaskar, R.V.; Thakare, V.M. A study of human activity recognition using adaboost classifiers on WISDM dataset. Inst. Integr. Omics Appl. Biotechnol. J. 2016, 7, 68–76. [Google Scholar]
  48. Catal, C.; Tufekci, S.; Pirmit, E.; Kocabag, G. On the use of ensemble of classifiers for accelerometer-based activity recognition. Appl. Soft. Comput. 2015, 37, 1018–1022. [Google Scholar] [CrossRef]
  49. Zainudin, M.S.; Sulaiman, M.N.; Mustapha, N.; Perumal, T. Activity recognition based on accelerometer sensor using combinational classifiers. In Proceedings of the IEEE Conference on Open Systems (ICOS), Melaka, Malaysia, 24–26 August 2015; pp. 68–73. [Google Scholar]
  50. Kolosnjaji, B.; Eckert, C. Neural network-based user-independent physical activity recognition for mobile devices. In Lecture Notes in Computer Science; Jackowski, K., Burduk, R., Walkowiak, K., Wozniak, M., Yin, H., Eds.; Springer: Cham, Switzerland, 2015; pp. 378–386. [Google Scholar]
  51. Zhang, J.; Qiao, S.; Lin, Z.; Zhou, Y. Human activity recognition based on residual network. In Proceedings of the 8th Annual International Conference on Geo-Spatial Knowledge and Intelligence, Xian, China, 18–19 December 2020; pp. 1–6. [Google Scholar]
  52. Lin, S.B.; Wang, K.; Wang, Y.; Zhou, D.X. Universal consistency of deep convolutional neural networks. arXiv 2021, arXiv:2106.12498. [Google Scholar] [CrossRef]
  53. Zihao, Z.; Geng, J.; Jiang, W. A time series classification method based on 1DCNN-FNN. In Proceedings of the 33rd Chinese Control and Decision Conference (CCDC), Kunming, China, 22–24 May 2021; pp. 1566–1571. [Google Scholar]
  54. Garcia, K.D.; de Sá, C.R.; Poel, M.; Carvalho, T.; Mendes-Moreira, J.; Cardoso, J.M.; de Carvalho, A.C.P.L.F.; Kok, J.N. An ensemble of autonomous auto-encoders for human activity recognition. Neurocomputing 2021, 439, 271–280. [Google Scholar] [CrossRef]
  55. Lima, W.S.; Bragança, H.L.; Souto, E.J. NOHAR-NOvelty discrete data stream for human activity recognition based on smartphones with inertial sensors. Expert Syst. Appl. 2021, 166, 114093. [Google Scholar] [CrossRef]
  56. Sena, J.; Barreto, J.; Caetano, C.; Cramer, G.; Schwartz, W.R. Human activity recognition based on smartphone and wearable sensors using multiscale DCNN ensemble. Neurocomputing 2021, 444, 226–243. [Google Scholar] [CrossRef]
  57. Ramesh, A.K.; Gajjala, K.S.; Nakano, K.; Chakraborty, B. Person authentication by gait data from smartphone sensors using convolutional autoencoder. In Proceedings of the International Conference on Intelligence Science, Durgapur, India, 24–27 February 2021; pp. 149–158. [Google Scholar]
  58. Dhammi, L.; Tewari, P. Classification of human activities using data captured through a smartphone using deep learning techniques. In Proceedings of the 3rd International Conference on Signal Processing and Communication (ICPSC), Coimbatore, India, 13–14 May 2021; pp. 1–6. [Google Scholar]
  59. Wenzheng, Z. Human activity recognition based on acceleration sensor and neural network. In Proceedings of the 8th International Conference on Orange Technology (ICOT), Daegu, Korea, 18–21 December 2020; pp. 1–5. [Google Scholar]
  60. Peppas, K.; Tsolakis, A.C.; Krinidis, S.; Tzovaras, D. Real-time physical activity recognition on smart mobile devices using convolutional neural networks. Appl. Sci. 2020, 10, 8482. [Google Scholar] [CrossRef]
  61. Mehmood, K.; Imran, H.A.; Latif, U. HARDenseNet: A 1D DenseNet inspired convolutional neural network for human activity recognition with inertial sensors. In Proceedings of the 23rd International Multitopic Conference (INMIC), Bahawalpur, Pakistan, 5–7 November 2020; pp. 1–6. [Google Scholar]
  62. Aswal, V.; Sreeram, V.; Kuchik, A.; Ahuja, S.; Patel, H. Real-time human activity generation using bidirectional long short term memory networks. In Proceedings of the 4th International Conference on Intelligent Computing and Control Systems (ICICCS), Madurai, India, 13–15 May 2020; pp. 775–780. [Google Scholar]
  63. Xia, K.; Huang, J.; Wang, H. LSTM-CNN architecture for human activity recognition. IEEE Access 2020, 8, 56855–56866. [Google Scholar] [CrossRef]
  64. Agarwal, P.; Alam, M. A lightweight deep learning model for human activity recognition on edge devices. Procedia Comput. Sci. 2020, 167, 2364–2373. [Google Scholar] [CrossRef]
  65. Zhang, H.; Xiao, Z.; Wang, J.; Li, F.; Szczerbicki, E. A novel IoT-perceptive human activity recognition (HAR) approach using multihead convolutional attention. IEEE Internet Things J. 2019, 7, 1072–1080. [Google Scholar] [CrossRef]
  66. Huang, J.; Lin, S.; Wang, N.; Dai, G.; Xie, Y.; Zhou, J. TSE-CNN: A two-stage end-to-end CNN for human activity recognition. IEEE J. Biomed. Health Inf. 2019, 24, 292–299. [Google Scholar] [CrossRef] [PubMed]
  67. Mukherjee, D.; Mondal, R.; Singh, P.K.; Sarkar, R.; Bhattacharjee, D. EnsemConvNet: A deep learning approach for human activity recognition using smartphone sensors for healthcare applications. Multimed. Tools Appl. 2020, 79, 31663–31690. [Google Scholar] [CrossRef]
  68. Tang, Y.; Teng, Q.; Zhang, L.; Min, F.; He, J. Efficient convolutional neural networks with smaller filters for human activity recognition using wearable sensors. arXiv 2020, arXiv:2005.03948v1. [Google Scholar]
  69. Beirami, M.J.; Shojaedini, S.V. Residual network of residual network: A new deep learning modality to improve human activity recognition by using smart sensors exposed to unwanted shocks. J. Health Manag. Inf. 2020, 7, 228–239. [Google Scholar]
  70. Zhang, Y.; Zhang, Z.; Zhang, Y.; Bao, J.; Zhang, Y.; Deng, H. Human activity recognition based on motion sensor using U-Net. IEEE Access 2019, 7, 75213–75226. [Google Scholar] [CrossRef]
  71. Pienaar, S.W.; Malekian, R. Human activity recognition using LSTM-RNN deep neural network architecture. In Proceedings of the IEEE 2nd Wireless Africa Conference, Pretoria, South Africa, 18–20 August 2019; pp. 1–5. [Google Scholar]
  72. Manu, R.D.; Kumar, S.; Snehashish, S.; Rekha, K.S. Smart home automation using IoT and deep learning. Int. Res. J. Eng. Technol. 2019, 6, 1–4. [Google Scholar]
  73. Varamin, A.A.; Abbasnejad, E.; Shi, Q.; Ranasinghe, D.C.; Rezatofighi, H. Deep auto-set: A deep auto-encoder-set network for activity recognition using wearables. In Proceedings of the 15th EAI International Conference on Mobile and Ubiquitous Systems: Computing, Networking and Services, New York, NY, USA, 5–7 November 2018; pp. 246–253. [Google Scholar]
  74. Chandini, U. A Machine learning based activity recognition for ambient assisted living. Int. J. Future Revolut. Comput. Sci. Commun. Eng. 2018, 4, 323–326. [Google Scholar]
  75. Dungkaew, T.; Suksawatchon, J.; Suksawatchon, U. Impersonal smartphone-based activity recognition using the accelerometer sensory data. In Proceedings of the 2nd International Conference on Information Technology (INCIT), Nakhonpathom, Thailand, 2–3 November 2017; pp. 1–6. [Google Scholar]
  76. Chen, Y.; Zhong, K.; Zhang, J.; Sun, Q.; Zhao, X. LSTM networks for mobile human activity recognition. In Proceedings of the International Conference on Artificial Intelligence: Technologies and Applications, Bangkok, Thailand, 24–25 January 2016; pp. 50–53. [Google Scholar]
  77. Abdallah, Z.S.; Gaber, M.M.; Srinivasan, B.; Krishnaswamy, S. Adaptive mobile activity recognition system with evolving data streams. Neurocomputing 2015, 150, 304–317. [Google Scholar] [CrossRef]
  78. Mekruksavanich, S.; Jitpattanakul, A.; Youplao, P.; Yupapin, P. Enhanced hand-oriented activity recognition based on smartwatch sensor data using LSTMs. Symmetry 2020, 12, 1570. [Google Scholar] [CrossRef]
  79. Han, D.; Lee, C.; Kang, H. Gravity control-based data augmentation technique for improving VR user activity recognition. Symmetry 2021, 13, 845. [Google Scholar] [CrossRef]
  80. Su, J.-Y.; Cheng, S.-C.; Huang, D.-K. Unsupervised object modeling and segmentation with symmetry detection for human activity recognition. Symmetry 2015, 7, 427–449. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Examples of symmetric and asymmetric activities.
Figure 1. Examples of symmetric and asymmetric activities.
Symmetry 14 02022 g001
Figure 2. An example of sensor data, which cannot be interpretable by humans.
Figure 2. An example of sensor data, which cannot be interpretable by humans.
Symmetry 14 02022 g002
Figure 3. Explainable and understandable sensor data for each human activity.
Figure 3. Explainable and understandable sensor data for each human activity.
Symmetry 14 02022 g003aSymmetry 14 02022 g003b
Figure 4. The general workflow of the proposed HARSI approach.
Figure 4. The general workflow of the proposed HARSI approach.
Symmetry 14 02022 g004
Figure 5. Loss values in the training and validation processes.
Figure 5. Loss values in the training and validation processes.
Symmetry 14 02022 g005
Figure 6. Confusion matrix obtained by the proposed HARSI method.
Figure 6. Confusion matrix obtained by the proposed HARSI method.
Symmetry 14 02022 g006
Figure 7. The execution time of the proposed method on different CNNs.
Figure 7. The execution time of the proposed method on different CNNs.
Symmetry 14 02022 g007
Table 1. Comparison of the recent HAR studies with our study.
Table 1. Comparison of the recent HAR studies with our study.
RefYearMethodDescriptionSensor
Types
DataSensor LocationNumber
of Activities
Sensor-Data-BasedSignal-Image-BasedXAI
CNNDNNRNNLSTMAE
[8]2022 Channel state information (CSI) based HARWireless signalCSIRoom6XX
[9]2022 Gait pattern analysisA, G, MSG * Center of mass7XX
[10]2022 Personalization in HARA, G, MUniMiB SHARPocket17XX
Motion Sense6
MobiAct15
[11]2022 HAR from piezoelectric-based kinetic energy signalsKEH transducersKEHHand, waist5XX
[12]2022 Feature extraction-based approachAWISDMPocket6XX
[13]2022 Comparative study on classifying human activitiesA, GUCI-HARWaist6XX
[14]2022 Gesture recognition in videosCameraSGRoom4XX
[1]2021 HAR from highly sparse body sensor dataA, RFIDRoomset1Chest4XX
Roomset2
[2]2021 Feature extraction-based approachA, G, MUniMiB-SHARPocket17XX
[3]2021 Biometric user identificationA, GUCI-HARWaist6XX
USC-HAD12
[4]2021 HAR in smart homesEnv. sensorsOrange4HomeRoom24XX
[5]2021 Industry 4.0-oriented approachAWISDMPocket6XX
[6]2021 Human pose and motion estimationCameraSGRoom5XX
Env. sensors8
[15]2021 Hybrid deep-learning-based modelMotion Kinect sensorSGRoom12XX
[16]2021 HAR using skeleton datasetsCameraUTD-MHADRoom27XX
MSR-Action3D20
[17]2021 Feature fusion-based approachA, G, MMHEALTHAnkle, arm, chest12XX
[18]2021 Causality feature extraction based approachEnv. sensorsAruba Room10XX
Milan15
Cairo13
[19]2021 HAR based on the Inception-ResNet modelA, GUCI-HARWaist6XX
Env. and body sensorsOpportunityRoom, body18
ADaphnetLegs, hip2
A, G, MPAMAP2Chest, ankle18
[20]2021 Multiple domain DL frameworkA, G, MSGHead, wrist, leg12XX
[21]2021 Feature-fusion-based approachA, GSGWaist6XX
UCI-HAR
[22]2021 Recognizing transitional activitiesA, GHAPTWaist12XX
HAD5
[23]2021 Optimal deep-learning-based approachA, GUCI-HARWaist6XX
USC-HAD12
[24]2021 HAR using time-series dataAUniMiB SHARPocket17XX
AWISDMPocket6
AUCI-HARWaist6
[25]2021 HAR on microcontrollersA, G, MPAMAP2Hand, chest, ankle12XX
[26]2021 Hybrid deep-learning-based approachA, GUCI-HARWaist6XX
AWISDMPocket
[27]2021 Feature-fusion-based approachA, GSGPocket6XX
UCI-HARWaist
[28]2021 Attention-based mechanismA, GHHARHand, chest, ankle6XX
A, G, MPAMAP212
A, GUSC-HAD12
[29]2021 HAR using multimodal sensorsMultimodalCMU-MMACRoom11XX
[30]2021 Multimodal complex HARA, G, MLifelogPocket, wrist, chest9XX
PAMAP2Wrist, arm, chest,18
[31]2021 Hierarchical hybrid deep-learning-based approachA, GUCI-HAPTWaist12XX
MobiActPocket11
[32]2021 Resource-constrained HAREMG sensorsMyo-TLElbow, wrist9XX
Db518
Our
Approach
Human activity recognition on signal images (HARSI)AWISDMPocket6
* SG: self-generated.
Table 2. Parameter settings and the number of parameters.
Table 2. Parameter settings and the number of parameters.
ModelLearning RateActivation FunctionOptimizerLoss FunctionTotal
Parameters
Total Trainable ParametersTotal
Nontrainable
Parameters
HARSI-ResNet342 × 10−3ReLU
(Rectified Linear
Unit)
Adam
(Adaptive
Moment
Estimation)
Cross
Entropy
21,815,104547,45621,267,648
HARSI-ResNet506 × 10−425,617,4722,162,56023,454,912
HARSI-ResNet1011 × 10−344,609,6002,214,78442,394,816
HARSI-AlexNet2 × 10−32,736,960267,2642,469,696
HARSI-DenseNet1214 × 10−48,010,6241,140,4166,870,208
HARSI-SqueezeNet_v1.02 × 10−31,265,856530,432735,424
HARSI-SqueezeNet_v1.11 × 10−31,252,928530,432722,496
HARSI-VGG161 × 10−315,253,568538,88014,714,688
HARSI-VGG192 × 10−320,565,824541,44020,024,384
Table 3. The performance of the proposed HARSI method on different CNN architectures. The best values are highlighted in bold.
Table 3. The performance of the proposed HARSI method on different CNN architectures. The best values are highlighted in bold.
ModelAccuracy (%)PrecisionRecallF-Measure
HARSI-ResNet3497.330.974330.973330.97326
HARSI-ResNet5096.000.960540.960000.96006
HARSI-ResNet10197.170.971940.971660.97161
HARSI-AlexNet89.170.890180.891660.89077
HARSI-DenseNet12196.670.966950.966660.96671
HARSI-SqueezeNet_v1.089.670.900250.896660.89725
HARSI-SqueezeNet_v1.193.000.929150.930000.92937
HARSI-VGG1696.830.968710.968330.96826
HARSI-VGG1998.000.979990.980000.97999
Table 4. Comparison of the proposed HARSI method against the classical machine learning methods on the same dataset.
Table 4. Comparison of the proposed HARSI method against the classical machine learning methods on the same dataset.
Ref.YearMethodAccuracy (%)
[12]2022Support Vector Machines87.40
Random Forest86.10
[34]2021Decision Tree82.00
Logistic Regression68.00
Multilayer Perceptron80.00
Neural Networks94.00
[35]2021Decision Tree 89.76
Linear Discriminant Analysis86.64
Gradients Boosting89.65
K-Nearest Neighbors 92.54
Bagging92.48
Random Forest92.71
Linear Kernel SVM78.55
RBF Kernel SVM89.07
Polynomial Kernel SVM92.48
[36]2021Random Forest 79.38
K-Nearest Neighbors75.04
Decision Tree77.60
Gradient Boosting 74.80
[37]2020Random Forest83.35
Neural Networks77.02
Decision Tree (J48) 75.96
Reduced-Error Pruning (REP) Tree 74.64
K-Nearest Neighbors72.08
KStar 71.84
Naive Bayes 63.89
Support Vector Machines55.45
[38]2020Random Forest92.78
Support Vector Machines91.39
[39]2020Neural Networks89.10
Decision Tree 87.45
Support Vector Machines95.13
Linear Support Vector Classifier 86.20
Logistic Regression81.10
Random Forest82.10
[40]2020Support Vector Machines82.00
[41] 2020Multilayer Perceptron86.95
[42]2019K-Nearest Neighbors92.00
Support Vector Machines93.50
Bagging93.80
[43]2018Random Forest82.66
K-Nearest Neighbors66.19
[44]2018Support Vector Machines82.27
[45]2018Naive Bayes 80.12
Decision Tree81.02
K-Nearest Neighbors77.58
Support Vector Machines80.93
[46]2017Naive Bayes Tree87.70
Multilayer Perceptron77.52
DT + LR + MLP91.62
NB Tree + MLP96.35
[47] 2016AdaBoost + J4897.83
AdaBoost + REP Tree97.33
AdaBoost + Random Tree95.69
AdaBoost + Random Forest94.44
AdaBoost + Hoeffding Tree87.84
AdaBoost + Decision Stump57.31
[48] 2015Decision Tree (J48)86.08
Logistic Regression77.52
Multilayer Perceptron88.81
J48 + LR + MLP91.62
[49]2015Decision Tree (J48)92.40
Logistic Regression84.30
Multilayer Perceptron91.70
J48 + LR + MLP93.00
[50]2015Neural Networks with Dropout85.36
Random Forest83.46
[33]2010Logistic Regression78.10
Decision Tree (J48)85.10
Multilayer Perceptron91.70
Average84.28
Our ApproachHuman Activity Recognition on Signal Images (HARSI)98.00
Table 5. Comparison of the proposed HARSI method against the state-of-the-art methods on the same dataset.
Table 5. Comparison of the proposed HARSI method against the state-of-the-art methods on the same dataset.
Ref.YearMethodAccuracy(%)
[12]2022CNN—Transfer Learning 90.40
Convolutional Neural Networks88.20
[5]2021CNN + Long Short-Term Memory97.76
Long Short-Term Memory96.61
Convolutional Neural Networks94.51
[13]2021Deep Neural Networks93.00
[24]2021Vanilla RNN + LSTM + GRU97.13
[26]2021CNN + Random Forest97.77
Deep Neural Networks74.00
Deep Neural Networks + LSTM81.00
Deep Neural Networks + Gated Recurrent Unit (GRU)80.00
Convolutional Neural Networks88.00
Convolutional Neural Networks + LSTM94.00
Convolutional Neural Networks + GRU82.00
[34]2021Deep Neural Networks95.00
[51]2021Residual Network95.66
Convolutional Neural Networks92.19
[52]2021Deep Convolutional Neural Networks91.25
[53]20211D Convolutional Neural Networks 91.12
1D CNN + Fuzzy Neural Network92.96
[54]2021Ensemble of Autoencoders (EAE) 82.00
KNN + Very Fast Decision Tree + Naive Bayes (EkVN)73.00
[55]2021NOvelty discrete data stream for Human Activity Recognition (NOHAR)93.00
[56]2021Deep Convolutional Neural Networks Ensemble89.01
[57]2021Convolutional AutoEncoder (CAE)95.60
[58]2021Convolutional Neural Networks95.00
Long Short-Term Memory97.50
[59]2020Convolutional Neural Networks93.25
[60]2020Deep Convolutional Neural Networks94.18
Region-based CNN93.68
[61]2020CNN—DenseNet94.65
[62]2020Bidirectional Long Short-Term Memory94.10
[38]2020Genetic algorithm-based classifier95.37
[39]2020Convolutional Neural Networks83.98
Long Short-Term Memory95.45
[63]2020LSTM–Convolutional Neural Networks95.75
[64]2020Lightweight Recurrent Neural Network—LSTM95.78
[65]2020Multihead Convolutional Attention 95.40
[66]2020Two-Stage End-to-end CNN with data augmentation (TSE + CNN + Aug)95.70
[41] 2020Gramian Angular Field + Multidilated Kernel Residual Network96.83
Long Short-Term Memory87.53
1D Convolutional Neural Network93.66
[6]2020EnsemConvNet (CNN-Net + Encoded-Net + CNN-LSTM)97.20
[68]2020Convolutional Neural Networks97.51
[69]2020Convolutional Neural Networks94.11
Residual Network 95.72
Residual Network of Residual Network 96.73
[40]2020Convolutional Neural Networks81.70
Recurrent Convolutional Network (RCN)94.00
Recurrent Convolutional Network + SVM91.50
[70]2019U-Net96.40
Mask Region-based CNN (R-CNN)86.20
SegNet: A Deep Convolutional Encoder-Decoder Arch.95.70
Full Convolutional Network (FCN)87.90
Deep Convolutional and LSTM 94.80
Long Short-Term Memory93.80
Convolutional Neural Networks 94.10
[71]2019LSTM–Recurrent Neural Networks93.81
[42]2019Supervised Regularization-based Robust Subspace (SRRS)93.50
Robust Principal Component Analysis 85.70
Latent Low-Rank Representation (LLRR)91.90
Joint Embedding Learning and Sparse Regression (JELSR)73.40
Principal Component Analysis (PCA)92.30
Linear Discriminant Analysis (LDA)71.50
[72]2019Long Short-Term Memory97.00
[44]2018Convolutional Neural Networks91.97
[45]2018Multivariate Bag-Of-SFA-Symbols 83.35
[73]2018Deep Autoencoder-Set Network94.90
[74]2018Long Short-Term Memory97.00
[43]2018Convolutional Neural Networks93.32
[75]2017Impersonal Smartphone-based Activity Recognition (ISAR)75.21
[76]2016Long Short-Term Memory92.10
[77]2015STream learning for mobile Activity Recognition (STAR)71.20
Average90.94
OurApproachHuman Activity Recognition on Signal Images (HARSI)98.00
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cengiz, A.B.; Birant, K.U.; Cengiz, M.; Birant, D.; Baysari, K. Improving the Performance and Explainability of Indoor Human Activity Recognition in the Internet of Things Environment. Symmetry 2022, 14, 2022. https://doi.org/10.3390/sym14102022

AMA Style

Cengiz AB, Birant KU, Cengiz M, Birant D, Baysari K. Improving the Performance and Explainability of Indoor Human Activity Recognition in the Internet of Things Environment. Symmetry. 2022; 14(10):2022. https://doi.org/10.3390/sym14102022

Chicago/Turabian Style

Cengiz, Ayse Betul, Kokten Ulas Birant, Mehmet Cengiz, Derya Birant, and Kemal Baysari. 2022. "Improving the Performance and Explainability of Indoor Human Activity Recognition in the Internet of Things Environment" Symmetry 14, no. 10: 2022. https://doi.org/10.3390/sym14102022

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