Next Article in Journal
SWAR: A Deep Multi-Model Ensemble Forecast Method with Spatial Grid and 2-D Time Structure Adaptability for Sea Level Pressure
Next Article in Special Issue
Semantic Features-Based Discourse Analysis Using Deceptive and Real Text Reviews
Previous Article in Journal
User-Generated Content in Social Media: A Twenty-Year Bibliometric Analysis in Hospitality
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Comparative Study of Machine Learning and Deep Learning Techniques for Fake News Detection

1
School of Information and Physical Sciences, College of Engineering Science and Environment, The University of Newcastle, Newcastle 2308, Australia
2
Department of Computer Science, King Khalid University, Abha 62521, Saudi Arabia
*
Author to whom correspondence should be addressed.
Information 2022, 13(12), 576; https://doi.org/10.3390/info13120576
Submission received: 27 October 2022 / Revised: 22 November 2022 / Accepted: 3 December 2022 / Published: 12 December 2022
(This article belongs to the Special Issue Advanced Natural Language Processing and Machine Translation)

Abstract

:
Efforts have been dedicated by researchers in the field of natural language processing (NLP) to detecting and combating fake news using an assortment of machine learning (ML) and deep learning (DL) techniques. In this paper, a review of the existing studies is conducted to understand and curtail the dissemination of fake news. Specifically, we conducted a benchmark study using a wide range of (1) classical ML algorithms such as logistic regression (LR), support vector machines (SVM), decision tree (DT), naive Bayes (NB), random forest (RF), XGBoost (XGB) and an ensemble learning method of such algorithms, (2) advanced ML algorithms such as convolutional neural networks (CNNs), bidirectional long short-term memory (BiLSTM), bidirectional gated recurrent units (BiGRU), CNN-BiLSTM, CNN-BiGRU and a hybrid approach of such techniques and (3) DL transformer-based models such as BERT b a s e and RoBERTa b a s e . The experiments are carried out using different pretrained word embedding methods across four well-known real-world fake news datasets—LIAR, PolitiFact, GossipCop and COVID-19—to examine the performance of different techniques across various datasets. Furthermore, a comparison is made between context-independent embedding methods (e.g., GloVe) and the effectiveness of BERT b a s e —contextualised representations in detecting fake news. Compared with the state of the art’s results across the used datasets, we achieve better results by solely relying on news text. We hope this study can provide useful insights for researchers working on fake news detection.

1. Introduction

Traditionally, people fundamentally consume news and information through newspapers and television (TV) channels. However, with the advent of the Internet and its intrusion into our lifestyle, the former has become less prominent [1]. Today, social media and live streaming platforms play a fundamental role compared with television as major news sources, as 62 percent of U.S. people gained news from social media in 2016, while 49 percent watched news through social media in 2012 (http://www.journalism.org/2016/05/26/news-use-across-social-media-platforms-2016/) (accessed on 3 March 2022).
Nowadays, the role of online social networks (OSNs) has significantly increased due to their convenient access. As a result, it is no longer limited to being a window for communication between individuals; rather, it has become an important tool for exchanging information and influencing and shaping public opinions [2].
The other side of the coin is fake news dissemination, specifically on OSNs, which poses a great concern to the individual and society due to the lack of control, supervision and automatic fact-checking, leading to low-quality and fake content generation. As a result, the general public is prone to countless disinformation and misinformation on OSNs, including fake news (i.e., news stories created intentionally with falsified information to mislead readers) [3,4]. Figure 1 illustrates the increase in fake news cases over the last two years.
It is not surprising to see falsehoods in information disseminated rapidly on OSNs, generating some degree of anonymity. For example, fake news has led to significant impacts on real-world events, where a piece of fake news from Reddit caused a real shooting (https://www.rollingstone.com/politics/politics-news/anatomy-of-a-fake-news-scandal-125877/) (accessed on 15 March 2022). In the 2016 U.S. presidential election, for instance, over one million posts were found to be related to a piece of fake news known as Pizzagate (https://tinyurl.com/z38z5zh) (accessed on 15 June 2022). Furthermore, during this period, the top 20 fake news pieces were reported to be larger than the top 20 most discussed real stories (https://tinyurl.com/y8dckwhr) (accessed on 15 June 2022). As stated by research on fake news velocity, tweets including falsified information on Twitter reach people six times faster than trustworthy tweets [6], resulting in fear, panic and financial loss in society [7]. According to a report in China, fake information constitutes more than one third of trending events on microblogs [8]. All these indicate how terribly fake news disseminates and how it can have an adverse social impact. According to [9], on Twitter, fake news, particularly political news, is usually retweeted by more users and disseminates extremely rapidly. In fact, some popular sources of information considered genuine, such as Wikipedia, are also subject to false information or fake news [10].
Without news verification, fake news would spread rapidly through OSNs, resulting in real consequences [11]. Therefore, researchers in the field of NLP dedicate their efforts to detecting fake news using various ML and DL algorithms. To better understand the performance of these algorithms and explore the directions for future study, in this paper, using four real-world fake news datasets, we compare the performance of seven classical ML algorithms—LR, SVM, NB, DT, RF, XGB and a voting ensemble ML method—with two scenarios of word representation methods: statistical (sparse) word vector representation methods and context-free (dense) pretrained word representation models. In addition, we also compare the performance of eight advanced ML models and two advanced transformer-based models: CNN, BiLSTM, BiGRU, CNN-BiLSTM, CNN-BiGRU, a hybrid model of CNN-BiLSTM-BiGRU (with two types of text representation methods: context-free and context-aware embedding models), BERT b a s e and RoBERTa b a s e . Note that the presented approaches concentrate only on the structure without considering word sense (i.e., resolving ambiguities in context).
The key contributions of this paper are as follows:
  • This paper surveys various feature-based methods and an assortment of ML and state-of-the-art transformer-based models used in the literature for fake news detection.
  • This paper provides a benchmark study for a wide range of classical and advanced ML algorithms with pretrained word-embedding methods (i.e., context-free and context-aware) as well as advanced pretrained transformer-based models using various datasets.
  • The experimental results indicate that no single technique can deliver the best performance scores across all datasets.
  • In general, advanced PLMs such as BERT b a s e and RoBERTa b a s e are effective at detecting fake news.

2. Problem Definition

We consider the problem of fake news detection as a binary classification, where a model is created to predict the credibility of a piece of news as either fake or real based on given attributes. Formally, assume A = { a 1 , a 2 , , a L } is a set of news articles, statements or tweets (where L is the length of the input). The goal of fake news detection is to predict whether the input text represents fake or real content.

3. Related Works

ML and DL techniques have significantly contributed to detecting and categorising fake news. Earlier research in fake news detection mainly relied on linguistic features to engineer relevant information. The textually based features were divided broadly into two groups, according to [12]: (1) general features and (2) latent features.

3.1. Classical ML Algorithms

General textual features are those typically used with a traditional ML context. Such features are the catalyst for describing content style w.r.t. four linguistic language levels: lexicon, syntax, discourse and semantic [13]. At the lexicon level, some statistical techniques, such as the bag-of-word (BoW) model, are mainly applied to evaluate the frequency statistics of lexicons [14]. The main task of the syntax level is to assess part-of-speech (POS) (e.g., nouns and verbs) frequencies using part-of-speech (POS)-taggers [14,15]. At the discourse level, rhetorical parsing tools and rhetorical structure theory (RST) can be used to capture rhetorical relations as features based on their frequencies among sentences [14,16].
For example, Pisarevskaya [17] modelled language markers and rhetorical relations using SVM and random forest classifiers and achieved an F-score of 0.65. At the semantic level, these frequencies can be assigned to lexicons or expressions corresponding to each psycho-linguistic category [12], such as those determined in Linguistic Inquiry and Word Count (LIWC) [18]. The underlying assumption of text-based methods is that textually and linguistic-based features differ between fake and real claims.
It is common for fake news stories to contain inflammatory and opinionated language, since the purpose of their creation is usually financial or political gain, as opposed to reporting objective information. In addition, fake news content often contains clickbait (i.e., to incentivise consumers to read the full article by clicking on a provided link) or incites confusion [19]. Linguistic-based features (e.g., lexical and syntactic features) refer to those features that can capture particular writing styles and sensational headlines [20].
Different linguistic-based features were extracted from the news text content and explored by existing works for fake news detection. In [18], the authors used a set of linguistic-based features and conducted learning experiments to develop robust fake news detectors. First, they analysed the various features of the news articles (including n-grams, LIWC, punctuation, grammar and readability). Then, based on these features, a linear SVM classifier was trained. In this study, computational linguistics was shown to be useful in detecting fake news automatically. Fuller et al. [21] proposed a linguistic-based method for deception detection consisting of 31 linguistic features, where 3 classifiers were to be used to refine them to only 8 cues. Such clues were based on the different feature sets in the linguistic field proposed earlier [22,23], in addition to others.
The use of a relatively simple approach based on term frequency (TF) and term frequency–inverse document frequency (TF-IDF) has been shown to be effective in some previous studies. Riedel et al. [24] applied a multilayer perceptron (MLP) in the context of the fake news challenge dataset, and they showed that using a relatively simple approach based on TF and TF-IDF yielded a good accuracy score of 88.5%. Ahmed et al. [25] applied a linear support vector machine (LSVM) using TF-IDF and with unigram features on a dataset of 2000 news pieces, yielding an accuracy of 92%. Bharadwaj [26] experimented with different features such as TF and TF-IDF with n-gram features, and the results showed that random forest (RF) with bigram features achieved the best accuracy of 95.66%. In [27], the authors experimented with character and word n-grams to study their effect on detecting fake news and concluded that the former contributed more towards improving fake news detection performance compared with word n-grams. The TF-IDF method and CountVectorizer (CV) were used by [28] as a feature extraction technique. They demonstrated that their approach was more accurate than state-of-the-art approaches such as Transformers’ Bidirectional Encoder Representations (BERT). Linguistic features have been used in both supervised and unsupervised learning approaches. It is essential to recognise that prospective deceivers use short sentences, certain phrasal verbs as well as certain verb tenses as part of their language, which has been revealed by some cooperative experiments between psychologists, linguists and computer scientists [22,29,30,31,32]. For example, 16 language variables were investigated by Burgoon et al. [29] in order to see if they may help distinguish between deceptive and truthful communications. They conducted two experiments in order to construct a database, in which they set up face-to-face or computer-based discussions with one volunteer acting as the deceiver and the other acting genuinely. Then, such discussions were transcribed for further processing, and they came up with certain linguistic cue classes that could reveal the deceiver. The authors utilised the C4.5 decision tree (DT) technique with 15-fold cross-validation to cluster and construct a hierarchical tree structure of the proposed features. Their proposed method achieved an overall accuracy of 60.72 per cent on a short sample of 72 cases. They concluded that noncontent words (e.g., function words) should be included when studying social and personality processes. According to the authors, in addition to specific nouns and verbs, linguistic style markers such as pronouns, articles and prepositions are equally important in revealing what someone is thinking and feeling. Therefore, liars are more likely to tell less self-relevant, less complicated and more negative stories. Moreover, liars were more likely than truth-tellers to utilise negative emotion (more negative feeling) terms [31].
Liars may feel guilty, either for the lie they said or for the topic they lied about [33]. Knapp et al. [34] observed that liars were far more likely than truth-tellers to make disparaging statements about their communication partners. In addition, they observed that the use of other references is more common among liars than among truth-tellers. However, this is inconsistent with what [31] found, where they observed that compared with truth-tellers, liars used fewer third-person pronouns. Horne et al. [35] applied an SVM classifier with a linear kernel using several linguistic clues. The authors cast the problem as a multi-class classification to identify whether an article was real, fake or satire, where classes were equally distributed. After a 5-fold cross-validation with BuzzFeed News (https://www.buzzfeednews.com/article/craigsilverman/viral-fake-election-news-outperformed-real-news-on-facebook) (accessed on 5 May 2022) that was enriched with satire articles, they reached 78 percent accuracy. Their feature set mainly consisted of POS tags and specific LIWC word categories.
A wide range of stylistic, morphological and grammatical text and punctuation could serve as useful cues to detect news veracity. These sets of features were adopted by Papadopoulou et al. [36] using a two-level text-based classifier in order to detect clickbait. Similarly, Rubin et al. [37] used some of these features, such as punctuation and grammatical features. Sentiment analysis, opinion mining and opinionated and inflammatory language as factors for fake news detection have also been explored [19,37,38,39,40,41,42,43,44,45,46]. Bhelande et al. [47] applied a naive Bayesian classifier using a bag of positive and negative words (i.e., sentiment analysis). Syntactic features (e.g., sentence-level features such as POS and n-grams) [4,13,18,25,37,39,44,48,49,50], and lexical features for salient content words (i.e., character-level and word-level features) [8,18,20,25,35,37,40,44,51,52,53,54,55,56] have been adopted by existing research. Castillo et al. [40] analysed and assessed information credibility on Twitter by adopting a list of rudimentary content-based features. In [57], the authors aimed to gauge the credibility of web claims by exploiting language stylistic features. They claimed that the language style of the reporting articles plays a crucial role in understanding their credibility. Da Silva et al. [58] examined ML approaches to identify and detect fake news. They found that neural networks composed of statistical classification algorithms heavily concentrate on analysing lexicons as the main features for detection to perform best. Potthast et al. [20] used news content to extract style features for fake news prediction. Psycholinguistic cues (i.e., additional indicators of persuasive language) such as sadness, anger and others and signals of biased language have been adopted in detection as well [9,18,37,46,49,54]. More work on fake news detection relying on linguistic cues can be found in [52,59,60]. A systematic mapping study was conducted in [61] to analyse and synthesise studies regarding the use of ML techniques for detecting fake news. Klyuev [62] discussed different approaches and techniques to fight fake news. They also discussed how important it is to define text features using NLP methods to build a text document profile. However, it turns out that straightforward approaches such as n-grams and POS tags tend to ignore different writing style features and cannot capture cues across long news articles. Indeed, irrelevant (noisy) text inevitably exists in fake news datasets, particularly those extracted from OSNs such as Twitter. If not processed, this would end up with classification performance leading to inaccurate predictions. This poses the challenge of how to encode such data so as to mitigate such a problem. Various dedicated efforts have been put forward to respond to this challenge using neural network-based models. Each provides a unique set of informative and useful features that can help discriminate fake news from real news.

3.2. Advanced ML and DL Models

Latent textual features are those used for news text representations which can be applied at the word level [63,64], sentence level [65,66] or document level [66], resulting in vector representations that can be used for further processing (i.e., as an input to a classifier). For example, one study on detecting rumours on Twitter applied Word2Vec in order to create vector representation [67,68]. Another study on detecting fake news based on the content [69] applied convolutional neural networks (CNNs) and bidirectional long short-term memory (BiLSTM) for embedding the textual and speaker metadata information for fake news detection. Quian et al. proposed two-level CNNs, where the first level produces embedding for a sentence using words and the second level uses the sentence embedding to generate article embedding [70]. Using CNNs and pretrained word embeddings, Goldani et al. [71] proposed a capsule network model based on the ISOT and LIAR datasets for (binary and multi-class) fake news classification. Their results showed that the best accuracy obtained using binary classification on ISOT was 99.8%, while multi-class classification using the LIAR dataset yielded 39.5% accuracy. Similarly, Girgis et al. [72] performed fake news classification using the LIAR dataset. The authors employed three different models: a vanilla recurrent neural network (RNN), a gated recurrent unit (GRU) and long short-term memory (LSTM). Regarding accuracy, the GRU model resulted in 21.7% accuracy, slightly higher than the other two models (LSTM = 21.6% and RNN = 21.5%). The authors presented a DL model in [73] for automatically detecting fake news in Slovak. In addition, several local online news sources were used to gather data related to the COVID-19 pandemic to train and evaluate various DL models. A model combining a bidirectional long-short-term memory network with one-dimensional convolutional layers achieved an average macro F1 score of 94% on an independent test set. Thus, advanced ML methods perform substantially better when used to classify fake news on a binary scale but perform much worse when used to classify fake news on a more refined level.
In fake news detection, it is important to capture contextualised features. In contrast to those unidirectional models that read the text input sequentially, BERT offers clear advantages over encoding models in the literature, where it basically reads the entire sequence of words at once and flows it through a stack of transformer encoders. The transformer is simply an attention-based neural network model which alleviates the currency-related issues in RNNs by handling long-term dependency in the text more efficiently using an attention mechanism. The attention mechanism is a module that interprets how significant a word is for the subsequent representation of the current word by assigning each word a weight based on the context (relevance of all other words in different positions in a sentence against the current word at the current position). Generally speaking, unlike feature-based engineering methods, which failed to fully exploit the rich semantic and syntactic information in the content, the results obtained by DL-based methods show superior improvements. Furthermore, as a state-of-the-art model, BERT can capture deep contextualised information. The technique of learning how to transfer knowledge is known as transfer learning, which stores and applies the knowledge gained from performing a specific task to another problem. Learning this way is useful for training and evaluating models with relatively small amounts of data. In recent years, pretrained language models (PLMs) have become mainstream for downstream text classification [74] thanks to transformer-based structures. Major advances have been driven by the use of PLMs, such as ELMo [75], GPT [76] or BERT [74]. BERT and RoBERTa, the most commonly utilised PLMs, were trained on huge corpora, such as those containing over three billion words for BERT [74]. The success of such approaches raises the question of how such models can be used for downstream text classification tasks. Over the PLMs, task-specific layers are added for each downstream task, and then the new model is trained with only those layers from scratch [74,77,78] in a supervised manner. Specifically, these models use a two-step learning approach. First, in a self-supervised manner, they learn language representations by analysing a tremendous amount of text. This process is commonly called pretraining. Second, feature-based and fine-tuning approaches can then be used to apply these pretrained language representations to downstream NLP tasks. The former uses pretrained representations and includes them as additional features for learning a given task. The latter introduces minimal task-specific parameters, and all pretrained parameters are fine-tuned for the downstream tasks. These models are advantageous in that they can learn deep context-aware word representations from large unannotated text corpora, namely large-scale self-supervised pretraining. This is especially useful when learning a domain-specific language with insufficient available labelled data.
As we will further discuss in later sections, this paper exploits the power of BERT b a s e and its variations in building robust fake news predictive models. Few studies have been conducted using such models, despite using different methodologies and different scenarios, which have shown promising results. One recent example is a study conducted by Kula et al. [79], which presents a hybrid architecture based on a combination of BERT and an RNN. Alghamdi et al. [80] presented a computational framework for automatic fake news detection using BERT b a s e . The authors applied BERT b a s e to encode the input text, and then the resulting output sequence was fed into a CNN network to extract the salient local features. The metadata were then encoded using a CNN followed by a BiLSTM and then a classification layer. Their study achieved state-of-the-art performance on the LIAR dataset. Aggarwal et al. [81] showed that BERT, even with minimal text preprocessing, provided better performance compared with that of the LSTM and gradient-boosted tree models. Jwa et al. [82] adopted BERT for fake news detection by analysing the relationship between the headline and the body text of news using the FNC dataset, where they achieved an F1 score of 0.746. In an attempt to automatically detect fake news spreaders, Baruah et al. [83] proposed BERT for the classification task, achieving an accuracy of 0.690.
Given that BERT is more complex to train (depending on how large the number of parameters being used is), a variation of BERT, the so-called DistilBERT method [84], provides a simpler and more reasonable number of parameters compared with that of BERT (reducing BERT by 40% in size while retaining 97% of its language understanding abilities), thus leading to faster training (60% faster). With a larger dataset, larger batches and more iterations, a robust BERT model was developed: the so-called RoBERTa [78]. A benchmark study of ML models for fake news detection was provided in [85], where the authors formulated the problem of fake news detection using three different datasets, including the LIAR dataset [69] (see Section 4.6.1 for more details about the dataset), as a binary classification. Their experimental results showed the power of advanced PLMs such as BERT and RoBERTa. A summary of the previous related studies on the LIAR dataset (binary classification) can be seen in Table 1.
Using a composite fake news dataset called FakeNewsNet (see Section 4.6.2 for more details about the dataset), which includes two datasets—PolitiFact and GossipCop—the authors of [87] employed an autoencoder with LSTM of two layers for both encoders and decoders and also used another network of LSTM with two layers to capture the temporal pattern of user engagements. The authors later presented a DL model based on the hierarchical attention network for fake news detection using the same datasets where their proposed framework achieved state-of-the-art results [88]. Using the PolitiFact dataset, a deep neural network model with various representations was proposed in [89]. A summary of the previous related work on the FakeNewsNet dataset can be seen in Table 2 and Table 3.
In [92], the authors applied a pretrained transformer model, the so-called XLNet model, combined with Latent Dirichlet Allocation (LDA) by integrating contextualised representations generated from the former with topical distributions produced by the latter. Their model achieved an F1 score of 0.967. In the same vein, a fine-tuned transformer-based ensemble model was proposed in [93]. The proposed model achieved a 0.979 F1 score on the Constraint@AAAI2021-COVID19 fake news dataset (see Section 4.6.3 for more details about the dataset). Similarly, the authors of [94] carried out several experiments on the same dataset, and they proposed a framework for detecting fake news using the BERT language model by considering content information and prior knowledge as well as the credibility of the source. According to the results, the highest F1 scores obtained ranged from 97.57 to 98.13. By applying several supervised advanced ML models such as CNNs, LSTM and BERT to detect COVID-19 fake news, the authors of [95] achieved the best accuracy of 98.41% using the BERT-based version. A summary of the previous related studies on the COVID-19 dataset can be seen in Table 4.
To summarise, classical ML algorithms are easy to comprehend and perform well on small datasets, but they (1) require complex feature engineering and (2) fail to capture substantial semantical contextual knowledge for a specific input text. To overcome this, advanced ML techniques such as CNN- and RNN-based methods are well suited for complicated classification problems powered by massive data and can learn more complicated (latent) features. However, CNNs typically struggle with capturing long-term contextual dependencies, while RNN-based methods perform suboptimally in handling such dependencies. As such, a combination of these two architectures may overcome some of their inherent limitations. Aside from the fact that surface-level features cannot effectively capture semantical patterns in text, the lack of sufficient data constitutes a bottleneck for DL models. Thus, to address this, the power of the DL transformer-based models, such as BERT and its variations, can be effectively leveraged to build robust fake news predictive models.

4. Comparative Study

Based on thorough literature review analysis, we conducted this comprehensive comparative study by selecting several popular, well-discussed classical and advanced ML and DL models for fake news detection. In this section, we will first present different word-embedding techniques developed in NLP and then introduce different traditional and advanced ML algorithms as well as DL models for fake news detection and the benchmark datasets, which will be used to conduct the comparative study. The objective is to examine the performance of a wide range of ML techniques, including classical and advanced ML models as well as DL approaches, across various datasets for fake news detection.

4.1. Embeddings

Word embedding is considered the centre of any NLP task and is basically a form of word representation that bridges the human understanding of language to that of a machine. Word embeddings are typically learned from a large corpus of text. The distributed representational vector that captures the semantics of the input text can be obtained using two main alternatives that have been used for several text classification tasks in NLP. Modelling text is challenging because it is messy, and techniques such as ML algorithms require well-defined, fixed-length inputs and outputs. ML algorithms require the text to be converted into numbers (specifically vectors of numbers) as they cannot directly work with raw text. These vectors capture more linguistic properties of the text, as “in language processing, the vectors x are derived from textual data, in order to reflect various linguistic properties of the text” [99] (p. 65). Textual data requires a numerical representation and, most importantly, an efficient representation for computation and manipulation. Various statistically and contextually based methods have been developed to represent text numerically. The former is based only on statistical metrics, which typically generate a sparse vector space, while the latter is based on the concept of word context, which produces a rather dense vector space.
Several approaches have applied such methods (e.g., the bag-of-words [100], n-gram [101] and TF–IDF methods [102]) as input for text classification using ML (e.g., NB classifiers [103], K-NN algorithms [104] and SVM [105]). Nevertheless, the contextualised representation that allows for efficient computation and captures the underlying patterns is urgently needed, especially with the massive amounts of textual data. These statistically based representation methods are computationally efficient and have achieved promising results where they are traditionally considered the centre of any text classification task. However, these methods focus entirely on capturing the frequency features of a particular word, and the contextual information of a text is fully disregarded, making it difficult to capture semantic-based information. To capture more semantical information, pretrained language representation models (a.k.a context-independent models) were developed, such as Word2Vec and GloVe, which captured those semantics patterns but did little to capture context information. A significant amount of attention has been devoted to developing context-aware representations of textual data, including transformer-based models such as BERT, which has led to outstanding results in most NLP mainstream tasks. Overall, these techniques try to model the following problem. Assume we have a corpus of n documents S = { d 1 , d 2 , , d n } , each of which consists of m words W = { w 1 , w 2 , , w m } and a vocabulary V. The embedding vector representation w i is defined as mapping each word w i in a specific document d i into a continuous space R d , where d is the dimension of the vector space. Mathematically, the words in a document can be mapped as follows:
w i w i , w i R d

4.1.1. Non-Contextualised Embeddings: Sparse Vector Representation-Based

In this subsection, we present two statistical methods that generate sparse vector representations of documents. The first one is a popular and simple feature extraction method with text data: the bag-of-words (BoW) method. The second one is the TF-IDF method, which overcomes the problem of the former.
Bag of words (BoW): The bag-of-words model, or BoW for short, is very popular, simple and flexible, and it can be used in a myriad of ways for extracting features from the text in order to be used for modelling using ML algorithms. It is a representation of text that describes the occurrence of words within a specific document. The idea of this distributional representation of features was investigated and proposed by Harris [106]. The BoW model, as the name implies, discards any information about the order or structure of words in the document and is only concerned with whether known words occur in the document and not where they occur in the document, thus failing to capture semantic patterns. To illustrate this, given two sentences (the NLP is difficult, not easy at all, and the NLP is easy, not difficult at all) with completely opposite semantic meanings, the BoW model would give them the exact same representation just because they have the exact same words but in a different order, which is not effective. Another issue of such a model is that as the vocabulary size increases, so does the dimension of the vector space since, in this model, the number of words in the vocabulary forms the dimension of the vector representation of a document, resulting in what are called sparse representations. Such large dimensions, however, are bound to result in the so-called curse of dimensionality, making it easy to fall into overfitting and resulting in terrible out-of-sample performance.
Term frequency–inverse document frequency (TF-IDF): The problem with the BoW method is that it treats all words as equally important, and this is attributed to the fact that BoW methods score each word based on its frequency in a document. Thus, frequently occurring words dominate others in the document with a larger score, which is problematic, especially when such words are not as informative to the model as less-occurring words. Rescaling the frequency of words by penalising the scores of the most frequent words across all documents is one approach to dealing with this issue. This approach is the so-called TF-IDF metric proposed in [107], and it has been widely used in many NLP tasks. The TF-IDF method allows for quantifying words by reflecting how important a word is to a document in a corpus of documents. This method is premised on the idea that each word is assigned its own weight w i j based on its appearance in the document and across all of the documents. These weights highlight words that are distinct and contain useful information in a given document. “Thus the idf of a rare term is high, whereas the idf of a frequent term is likely to be low” [108] (p. 118). For each word in a document j, the TF-IDF value is calculated by first calculating the term frequency (TF), which counts the number of occurrences of words in a document, and then the inverse document frequency (IDF), which is the catalyst for ensuring that words appear less commonly are assigned more weights compared with those occurring more frequently (e.g., stop words), which is calculated as follows:
l o g ( | D | d f i )
where d f i denotes the number of documents that contains a word i and | D | refers to the number of documents in the corpus. The TF-IDF metric is calculated as follows:
w i = t f i j · l o g ( | D | d f i )
where t f i j , d f i and | D | refer to the number of appearances of a word i in the document j, the number of documents containing a word i and the number of documents, respectively. However, this method cannot capture semantic patterns, making it only useful for lexical features.

4.1.2. Non-Contextualized Embeddings: Dense Vector Representation-Based

Word embedding: Perhaps one of the key breakthroughs for the remarkable performance of ML methods in a suit of NLP tasks is a way of representing words (i.e., a learned representation) in a given text by allowing words with similar meanings to have a similar representation. It is this approach that generates a dense vector that carries more informative information. This can have many advantages, as “one of the benefits of using dense and low-dimensional vectors is computational: the majority of neural network toolkits do not play well with very high-dimensional, sparse vectors” [99] (p. 92). This also allows the model to generalise well. Contrary to classical word representation methods such as BoW, which generate sparse word representations using thousands or millions of dimensions, the rationale behind the word-embedding approach is premised on the idea of assigning each word a densely distributed representation (i.e., a real-valued vector) with often tens or hundreds of dimensions.
Word2Vec: Contrary to the previously discussed word representation methods that fail to capture the semantic and syntactic relations between the words, producing sparse vector representations, Word2Vec and GloVe (discussed next) can generate dense semantic representations. Word2Vec, developed by Tomas Mikolov et al. at Google in 2013 [109], is a statistical method that leverages the use of neural networks for efficiently learning a standalone word embedding from a given text corpus. This approach is considered a de facto standard for developing pretrained word embedding. The learned vectors of such an approach can be analysed, and interesting results can be found. Two different learning approaches have been proposed to model the algorithm architecture: the continuous bag-of-words (CBOW) and the skip-gram models. The CBOW model uses the context of a current word in order to predict that word. In other words, it learns the embedding by predicting the current word based on its context. Alternatively, the continuous skip-gram model learns by predicting the surrounding words given a current word. The Word2Vec approach is advantageous in that efficient and high-quality word embeddings can thankfully be learned with less space and time complexity, and this shows the key benefit of this approach, where it can handle larger corpora of a text by allowing larger dimensional embeddings to be learned (more dimensions) from such corpora. Here, the Gensim library is used to retrieve Word2Vec embeddings. The text is first cleaned, then stemming is used, and stop words are removed. The models were initialised with Word2Vec embeddings of a vector size set to 300-d and maximum length tokens set to 128.
GloVe: By extending the previous word-embedding approach (Word2Vec), the Global Vectors for Word Representation method, or GloVe for short, was developed by Pennington et al. [64] in order to learn word vectors more efficiently. This approach generally results in better word embeddings. This is because GloVe combines global statistics from matrix factorisation techniques such as latent semantic analysis (LSA) and local context-based learning methods such as Word2Vec. In short, “GloVe, is a new global log-bilinear regression model for the unsupervised learning of word representations that outperforms other models on word analogy, word similarity, and named entity recognition tasks.” [64]. Here, the models are initialised with 100-dimensional pretrained GloVe embeddings.

4.1.3. Contextualized Embeddings: Context-Aware Embeddings

Bi-Directional Encoder Representations from Transformers (BERT): PLMs (vector representations of words and embeddings) trained on massive amounts of textual data have formed the basis of many language-based tasks nowadays. As context-independent neural embeddings, Word2Vec and Glove extracted from shallow neural networks are examples of the most frequently pretrained word embedding techniques prior to the advent of recent trends (PLMs shined in 2018). Yet, nevertheless, these techniques failed to capture deeper contextual relations, since they mostly model indirect relationships by capturing only short-range context based on a specific co-occurrence window. In fact, since 2018, the interest of the NLP community in these kinds of pretrained word-embedding techniques has constantly been fading in favour of the most recent trend of transfer learning. Examples include Universal Language Model Fine-Tuning (ULMFiT) [110], Embedding from Language Models (ELMo) [75], OpenAI Generative Pretrained Transformer (GPT) [76], and Google’s BERT model [74]. ULMFiT [110] is pretrained on a universal language model on a general domain corpus which then can be fine-tuned on the target task data. Radford et al. [76] generated a transformer-based language model, the so-called OpenAI GPT model, which is a unidirectional language model. Unlike OpenAI GPT, BERT, generated by Devlin et al. [74], is the first deeply bidirectional and unsupervised language representation that employs a multi-layer bidirectional transformer encoder which jointly conditions both the left and the right contexts in all layers. The transformer architecture comprises two blocks, an encoder and a decoder, to read the text and produce a prediction, respectively. BERT uses only the encoder portion of the transformer. BERT [74] is a language representation model which Google AI introduced. Before discussing how BERT works, we first discuss some key points behind its prominent success. As the first-of-its-kind language representation, BERT contains several transformer encoders stacked together that can be used to pretrain deep bidirectional representations. The concept of bidirectionality in BERT allows it to consider left and right contexts. In other words, BERT is based on a self-attention layer jointly conditioning both the left and the right contexts in all layers, and thus, BERT generates context-aware embeddings. This is the key differentiator between BERT and its predecessor OpenAI GPT, where the former is deeply bidirectional whereas the latter is a unidirectional pretrained model (left-to-right language model pretraining). More details about its architecture will be introduced later.

4.2. ML Algorithms

In this subsection, we mainly describe the classification models we used.

4.2.1. Classical ML Models

The six models that were chosen for investigation, as previously stated in the introduction, were SVM, LR, MNB, DT, RF and XGB with both TF-IDF features and pretrained Word2Vec representations. A plethora of ML algorithms have been explored and tested in the literature for fake news detection. Below is the list of those approaches implemented in this work:
  • Logistic Regression (LR): LR is a statistical model applied as a great baseline algorithm in a wide range of text classification tasks.
  • Support Vector Machine (SVM): The SVM classifier is a strong classifier that yields promising results in a suite of NLP tasks.
  • Multinomial Naive Bayes (MNB): MNB is a kind of probabilistic algorithm (a Bayesian learning approach) that is also popular and yields great results in different NLP tasks.
  • Decision Tree (DT): This is a tree-based algorithm whose end nodes represent high-level features. A branch represents an output, while a leaf represents a label class. There are internal nodes that test one attribute and branch from a node that selects one value for the attribute. The leaf node is used to predict the class label. Classification is carried out based on supervised learning, which involves mapping the features and values to desired outcomes.
  • Random Forest (RF): RF models consist of a set of decision trees, each trained from a random selection of features.
  • XGBoost (XGB): This is an ensemble ML algorithm. The XGB algorithm uses a gradient-boosting framework whose algorithm is based on decision trees. Through boosting, the trees are constructed sequentially, with each one (i.e., weak learners) aimed at reducing the errors of the previous one. With the help of these weak learners, the boosting technique is able to combine these weak learners to produce a strong learner.
  • Ensemble: This is a hard voting ensemble learning method that combines the ML algorithms, including LR, SVM, DT, MNB, RF and XGB, which is built for better performance.

4.2.2. Advanced ML Models

Developing an automatic fake news detection model is more important than ever before, given how much data a single person can curate daily. Just think of solutions for detecting and categorising social users’ tweets on social media to understand the characteristics of fake and real content. Moreover, detecting fake content from text has already shown its importance in the literature, where adding features extracted from text is essential for good performance in fake news classification. Advanced ML and DL methods have proven their effectiveness in the field of NLP. Here, we investigate the power of 10 different advanced ML models—CNNs, BiLSTM, BiGRU, CNN-BiLSTM and CNN-BiGRU, an ensemble of a CNN, BiLSTM and BiGRU—as well as the transformer-based models BERT b a s e and RoBERTa which are used for fake news detection. The workflow of processing advanced ML models is shown in Figure 2. As shown in the figure, the embedding layer (initialised using Word2Vec, GloVe or BERT b a s e ) was used to embed the data. Then, the resultant representations were fed into the model (CNN, BiLSTM, etc.), followed by a fully connected layer (FC) comprising a single neuron that was activated using a sigmoid function for classification. The models are described below with their experimental set-ups. Table 5 lists the hyperparameters used in the advanced ML models. Binary cross-entropy and an Adam optimiser were used to train the models:
  • CNN: A one-dimensional convolutional neural network is a powerful ML approach for automatically extracting features from text inputs. A CNN can extract local features automatically but is less computationally expensive than other ML algorithms. Here, the architecture includes a single CNN layer with 128 filters with a kernel size of 5, which are activated with ReLU as an activation function. The generated feature map is then refined and reduced using the max-pooling layer, resulting in the most relevant information. After that, the output is flattened and passed to a dense output layer, with a single unit activated with a sigmoid as an activation function.
  • LSTM: The LSTM model [111] improves on the RNN’s flaws by adding an additive and multiplicative interaction to the recurrence formula and a distinct memory state. A model’s complexity can also be increased by stacking LSTM layers. With three gates—an input gate, a forget gate and an output gate—LSTM models eliminate the gradient vanishing and explosion concerns brought by RNNs. An important characteristic of LSTM models is their ability to capture long-term dependency. The LSTM method has been proven to be effective when used for long sentences [112]. Mathematically, the LSTM components can be formulated as follows [113]:
    f t = σ ( W x f · x t + W h f · h t 1 + b f ) ,
    i t = σ ( W x i · x t + W h i · h t 1 + b i ) ,
    C t ˜ = t a n h ( W x c · x t + W h c · h t 1 + b i ) ,
    C t = f t C t 1 + i C t ˜ ,
    o t = σ ( W x o · x t + W h o · h h 1 + b o ) ,
    h t = o t t a n h ( C t ) .
    In the formulas above, σ represents the logistic sigmoid activation function. W, b and C t represent the weight matrix, the bias and the state of the memory unit at time t, respectively. Here, a single BiLSTM layer with 128 units is used to encode the input text.
  • GRU: In a GRU variant, there are only two gates: an update gate and a reset gate. The update gate combines the forget and input gates and decides what information will be passed to the current state. The reset gates determine when to ignore the previously hidden state [114]. As with LSTM, the update and reset gates are computed as follows [114]:
    r t = δ ( W r h t 1 + U r x t + b r ) ,
    z t = δ ( W z h t 1 + U z x t + b z ) ,
    h t = ( 1 z t ) h t 1 + z t h t ˜ ,
    h t ˜ = t a n h ( W h t ˜ ( h t 1 r t ) + U h t ˜ x t ) .
    In the formulas above, δ ( . ) denotes the logistic sigmoid function and W and U show the weight matrices of gates h t and b, respectively, referring to the hidden state and bias vectors. A basic RNN considers the context of the past but cannot consider the context of the future. Hence, to account for future and previous contexts, bidirectional LSTM (BiLSTM) and bidirectional GRU (BiGRU) are excellent choices thanks to their breakthrough designs. To accomplish this, the forward and backward hidden layers are combined, thereby controlling the temporal information flow in both directions and leading to better learning.
    Here, we used a single BiGRU layer with 128 neurons. Even though BiLSTM and BiGRU have shown their superiority in a suite of NLP problems, they are not free from two shortcomings: (1) as the high-dimensional input space increases, so does the complexity of these models, leading to further complexity in optimizing such models, and (2) as these models can capture succeeding and proceeding contextual information (bidirectionality concept), they are not able to focus on the most salient parts of the contextual information of the text. Therefore, to overcome the former issue, a CNN can be used to reduce the dimensionality of the feature space while retaining the informative features from the text. In addition, a CNN can capture and extract local patterns.
  • CNN-BiLSTM: Hybridising recurrence-based models with a CNN helps extract salient features, capturing local contextualised patterns and improving the model’s accuracy. First, a single CNN layer of 128 filters with a kernel size of 5 is used to process the input vectors and extract the local features. The resultant feature maps of the CNN layer are then fed to a single BiLSTM layer with 128 units to learn the long-term dependencies of the local features of news articles. This is followed by an output layer with a single unit activated with a sigmoid function. The temporal and contextual features and long-term dependencies of the input text can be learned and captured from the text by using an RNN, and important local features can be detected by harnessing the power of the CNN in handling the spatial relations [115,116].
  • CNN-BiGRU: Similar to the CNN-BiLSTM model, the architecture with a BiLSTM layer was replaced with a BiGRU layer.
  • Hybrid: This is a hybrid model that combines three models: a single CNN layer with 128 neurons of a kernel size of 5, followed by a max-pooling and then a BiLSTM layer with 128 units and then a BiGRU layer with 128 units.

4.2.3. Transformer-Based Models

By analysing existing related works, only a few studies have used transformer-based models to detect fake news, and little research has explored how to best harness such models to detect fake news. It becomes prohibitively challenging to process massive amounts of user-generated content manually. Therefore, automated systems capable of detecting fake content are essential. However, fake news on social media is a non-trivial task, since fake news is written deliberately to mislead readers, and user-generated content is typically of poor quality. To address these challenges, researchers proposed various methods for interpreting the meaning of a word through embedding vectors. Neural network-based methods such as Word2Vec and GloVe are commonly used to learn word embeddings from large word corpora. However, these embedding models have the disadvantage of being context-free since context is neglected, and static embeddings for words are generated regardless of their contexts. Therefore, to achieve finer-grained performance, a model must be able to capture semantic and contextual patterns. Moreover, the advanced ML models can automatically extract semantic information from a given input to detect fake content, but they cannot accurately recognise fake content without a deep understanding of the text. As such, there has been growing interest in the attention paradigm in recent years.
There is an overall paradigm shift taking place in the NLP community, which aims to develop a set of models that not only improve accuracy but also address the problem of lacking labelled data, which has been a long-standing issue in the scientific community. In addition, there is an urgent need to detect fake news automatically. However, this is a challenging task, since existing ML models (prior to the advent of transformer models) fail to provide a deeper semantic understanding of text inputs. This has caused NLP research to make great strides by introducing pretrained transformer-based language models. Using PLMs trained on massive unlabeled data for text classification tasks is becoming increasingly popular. To adapt for the downstream task, new neural network layers are layered on top of the pretrained layers in the PLM [117]. Here, a fully contented (FC) layer is added on top of the PLMs for classification purposes. A sophisticated approach is needed to detect fake news, since it has become increasingly difficult to distinguish between fake and real content.
This section evaluates two models—BERT [74] and RoBERTa [78]—which have been considered the key breakthroughs of the impressive performance in a suite of NLP tasks, largely due to their powerful language representations being learned from massive amounts of a text corpus. Moreover, such models can be easily fine-tuned to a specific downstream task through so-called transfer learning:
  • BERT: Originally introduced by Devlin et al. [74], BERT stands for Bidirectional Encoder Representation from Transformers. As the first deeply bidirectional and unsupervised language representation, this model uses a multi-layer bidirectional transformer encoder that simultaneously conditions the left and right contexts. As a result, BERT generates embeddings that are context-aware. BERT further eliminates the unidirectional constraint by performing pretraining using an unsupervised prediction task that includes a masked language model (MLM) which understands context and predicts words. A vector representation can therefore be generated by the model that captures the general information of the input text. The semantic representation of each word in the input text can be improved using an attention mechanism by boosting semantic representation based on the context of the word. The attention mechanism plays an important role in transformer architecture in that it assigns varying weights to different parts of text according to their contributions to the output. An Attention function maps queries and follows key-value and output-vector pairs, which can be seen in Equation (14):
    A t t e n t i o n ( Q , K , V ) = S o f t m a x ( Q K T d k ) V
    where Q, K and V denote the query, key and value itself, respectively. d k denotes the dimension of the key vector k and query vector q. Attention uses a Softmax activation function that normalises the inputs to a value between 0 and 1. BERT uses a multi-head attention mechanism which can be seen in Equation (15), where each specific head and the associated weight matrices are denoted with the subscript i:
    M u l t i H e a d ( Q , K , V ) = C o n c a t ( h e a d 1 , , h e a d h ) W O
    where W O denotes a weight matrix that was trained jointly with the model and each h e a d i is calculated as follows:
    h e a d i = A t t e n t i o n ( Q W i Q , K W i K , V W i V )
    Even though BERT contains millions of parameters (i.e., BERT b a s e contains 110 million parameters, while BERT l a r g e has 340 million parameters) [74], in contrast to pretraining, BERT is relatively inexpensive to apply to downstream tasks using jointly fine-tuned parameters based on a pretrained model. In this work, we use BERT b a s e .
  • RoBERTa [78]: An optimised version of the BERT approach was introduced by Facebook. In this method, BERT is retrained with an improved training methodology by (1) removing the Next Sentence Prediction task from pretraining; (2) using 10 times as much data as BERT to train RoBERTa and (3) introducing dynamic masking with larger batch sizes so that the masked tokens change during training, as opposed to the static masking pattern used in BERT. In this way, RoBERTa differs from BERT in the way it approaches pretraining. We experiment with RoBERTa b a s e in this paper.

4.3. Preprocessing

Preprocessing of the input text included stop word and punctuation removal. The text was then passed through a process of tokenising it into tokens and building a dictionary as a result of learning the vocabulary included in the given text corpus. Such a dictionary was then used to map every token into a unique integer, resulting in sequences of integers. Then, we padded or truncated such sequences to a fixed number of entries, as we needed to feed the models with vectors of the same length. The downside here was that any vectors that were too long would be truncated, resulting in the loss of some (useful) information. These sequences would then be transformed into fixed-length vectors of word embeddings (i.e., the dimension of such word vector embeddings being used). We then initialised the neural networks with (1) context-independent pretrained word-embedding models such as Word2Vec and GloVe and (2) the context-informed pretrained language model BERT b a s e . For classical ML algorithms, in order to convert tokenised texts into features, CountVectorizer (CV) and TF-IDF were used as statistical feature extractors for the ML models. Additionally, we used pretrained Word2Vec and GloVe to produce word representations as features for classical ML algorithms.

4.4. Experimental Set-Up

All the work for the experiments was carried out using an Intel Core i5 2.3 GHz, 8 GB RAM system running macOS. Classical ML classifiers were implemented using the scikit-learn package and Keras library to implement the advanced ML models. The HuggingFace library was used to implement the PLMs. For the advanced ML models, we carried out the experiments with two scenarios: (1) with context-independent embedding models such as Word2Vec and GloVe and (2) with context-informed embedding models such as BERT b a s e . In our final configuration, we used an Adam optimiser with a default learning rate value for the advanced ML models, namely 2 × 10 5 for BERT b a s e models and 1 × 10 5 for training RoBERTa b a s e models. Again, based on a trial-and-error examination, these values performed the best. For the advanced ML models and PLMs, the sigmoid function was used in the output layer to reduce the error during training, while binary cross-entropy was used to calculate the loss during backpropagation. All models were trained for 10 epochs with a batch size of 32.

4.5. Evaluation Metrics

We employed four evaluation criteria extensively used in text classification tasks, namely accuracy, precision, recall and F1 score (calculated as in the equations below), to assess the performance of the models:
  • Accuracy (A): Accuracy is a measure of the classifier’s ability to correctly classify a piece of information as either fake or real. The accuracy can be estimated using Equation (17):
    A c c u r a c y = T P + T N T P + T N + F P + F N
  • Precsion (P): Precision is a measure for the classifier’s exactness such that a low value indicates a large number of false positives. The precsion represents the number of positive predictions divided by the total number of positive class values predicted and is calculated using Equation (18):
    P r e c i s i o n = T P T P + F P
  • Recall (R): Recall is considered a measure of a classifier’s completeness (e.g., a low value of recall indicates many false negatives), where the number of true positives is divided by the number of true positives and the number of false negatives, as can be clearly seen in Equation (19):
    R e c a l l = T P T P + F N
  • F1 score (F1): The F1 score is calculated as the weighted harmonic mean of the precision and recall measures of the classifier using Equation (20):
    F 1 = 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 = 2 * T P 2 * T P + F P + F N
    where T P , T N , F P and F N are true positive, true negative, false positive and false negative, respectively.

4.6. Datasets

Our study made use of the following datasets described below. These datasets best reflect the real world, where the percentage of fake news that is fabricated and shared is smaller than the volume of news produced on a daily basis. The baseline models on the LIAR, FakeNewsNet-PolitiFact, FakeNewsNet-GossipCop and COVID-19 datasets are summarised in Table 1, Table 2, Table 3 and Table 4, respectively.

4.6.1. LIAR

This is a large-scale, publicly available fake news dataset proposed in [69]. The dataset has about 12,800 entries and contains two main components: user profiles and short political statements. The user profile features include the speaker’s name, job, party affiliation, state, credit history and context. The statements (reported during the time interval from 2007 to 2016) have been labelled by the editors of Politifact.com (accessed on 18 March 2022) using six fine-grained categories, namely true, mostly true, half true, barely true, false and pants on fire. These six labels are relatively balanced in size. Overall, each statement has its associated label and information about the speaker of that statement. The authors considered a multi-class classification task on the Liar dataset. Several textual and metadata elements, such as the speaker’s affiliation and the source newspaper, were included, and labels were assigned based on the six degrees of truth that PolitiFact provides. In our study, we cast the problem as a binary classification problem where we set true, mostly true, and half true to “true” and the rest to “false”. The statistics of the dataset are shown in Table 6.

4.6.2. FakeNewsNet

This comprehensive dataset (https://github.com/KaiDMML/FakeNewsNet) (accessed on 20 March 2022) consists of full-text news articles collected from the politifact.com and gossipcop.com (accessed on 18 March 2022) websites. Each of these includes tagged news content (e.g., news articles) and social context information (e.g., relevant social user interactions for news articles). Nevertheless, this contextual information may not always be available in the real world. Thus, we solely relied on content-based features (i.e., news article text). The statistics of the dataset are tabulated in Table 7.

4.6.3. COVID-19

The goal of the Constraint@AAAI2021 COVID19 fake news detection challenge [118] was to build a model that has the potential to identify whether a message about COVID-19 is fake or real news. This dataset is a collection of COVID-19-related social media posts, comments and news, classified as real or fake based on their truthfulness. The dataset [98] was collected from various social media platforms, such as Twitter and YouTube. In addition, the challenge organisers collected 10,700 social media posts and news articles about COVID-19 in the form of an annotated dataset in English. As the dataset was separated in advance by the task organisers into training, validation and testing sets, we opted to evaluate our models using the original split. The statistics of the dataset are shown in Table 8.

5. Analysis of Experimental Results

Unlike traditional ML algorithms, which require human experts to (manually) encode domain knowledge through feature engineering, which is inefficient and impractical, advanced ML and DL models can learn relevant and important feature representations automatically, making them particularly well-suited for NLP tasks. This section presents the experimental results related to a set of datasets commonly used in fake news detection. First, a comparison of the results of classical ML classifiers using different word representation methods across the used performance metrics is shown in Table 9, Tables 11 and 13. Second, a comparison of the performance of an assortment of advanced ML models using two scenarios, context-independent embedding models and context-informed embedding models, is presented in Table 10, Tables 12 and 14.

5.1. Analysis of Results on the LIAR Dataset

This subsection presents the results of an assortment of ML and DL models using the LIAR dataset. It is apparent from Table 9 that the highest accuracy result achieved on the LIAR dataset was 63.9% for the ensemble with TF-IDF features, where this classifier also resulted in significantly high precision and recall values, while the state-of-the-art result achieved on LIAR (binary classification) was 62% (refer to Table 1).
In contrast, the lowest result was 51.78% for DT with GloVe embeddings. The classifier could not detect some patterns in the given input text. As presented in Table 10, the performance of different models was compared. What stands out from this table is that the majority of the models achieved an accuracy of no more than 62% with BERT b a s e and the CNN-BiLSTM using BERT b a s e embeddings reporting an accuracy of 63.06%. It is clear from the table that the lowest result was reported for the recurrence-based models with Word2Vec embeddings. As the LIAR dataset contains short political statements, it is difficult to obtain useful clues that could help discriminate fake news from real news, specifically when using complex models such as LSTM, since this may increase the risk of overfitting. Based on that, we agree with [85] that a news article’s information and the length of the dataset are critical factors that affect the performance of recurrence-based models. Such models are more likely to overcome overfitting when sufficient information is provided in a news article. The obtained results indicate that the classical ML methods outperformed sophisticated DL methods on the LIAR dataset, including the most recent advanced PLMs such as BERT b a s e and RoBERTa b a s e .

5.2. Analysis of Results on FakeNewsNet Dataset

Previous studies used the accuracy, precision, recall and F1 scores to evaluate the performance of the proposed classifiers on the FakeNewsNet dataset. Table 2 and Table 3 display the summary of the results obtained from previous work on the PolitiFact and GossipCop datasets, respectively. What is striking about the resulting accuracies in these tables is that relying on modelling the unigram features (weighted by TF-IDF) showed a constantly high accuracy, outperforming the SAF and CNN baselines even with the combinations of social clues and news content on both datasets. As demonstrated in Table 11, the highest accuracy value (89.93%) on the PolitiFact dataset was achieved using the ensemble with CV as a feature extractor method. This model (Ensemble+CV) was found to be the second-best model for the GossipCop dataset, with SVM+TF-IDF (F1 score of 91.55%) being found to be the best-performing model among the other models on the GossipCop dataset. DL approaches, particularly the most recent advanced PLMs, achieved state-of-the-art performance in many NLP tasks. As can be observed from Table 12, RoBERTa b a s e performed the best on the PolitiFact dataset with an F1 score of 93.17, while the accuracy results (92%) were on par with the ensemble advanced ML model using BERT b a s e representations. For the GossipCop dataset, CNN-BERT b a s e was found to be the best-performing model (F1 score of 91.45%). This indicates that context-aware transformer-based models help uncover useful patterns to distinguish fake news from real news. Since the GossipCop dataset is highly imbalanced, different upsampling and downsampling techniques can help balance the dataset and further increase detection performance. By relying only on news text and having minimal text preprocessing, we achieved better results than the state-of-the-art results on the FakeNewsNet dataset.

5.3. Analysis of Results on COVID-19 Dataset

BERT can effectively detect fake content as it has the power to encode deep semantic contextual information. Based on the results in Table 13 and Table 14, we have shown that the BERT b a s e model, with a linear layer being added on top of it for classification, achieved the best results compared with its counterparts and other classical ML and DL approaches, followed by CNN-BERT b a s e , which was ranked as the second-best-performing model on the COVID-19 dataset. Due to the incapability of vector space methods such as CV and TF-IDF to take context into account, using these representations with ML classifiers relies on the appearance of tokens in making final decisions, regardless of their context. The results of our study indicate that vector space models are ineffective at capturing the deeper semantics and contextual patterns contained in user-generated content on Twitter. A major advantage of BERT (and its variations) in the case of Twitter (where user-generated content often contains misspellings, noise and abbreviations) is the use of sub-tokens rather than fixed tokens. It is thus ideal for use with such data [119] instead of standard context-independent word embeddings that operate at the word level.

6. Discussion

In this section, we will have an overall discussion of the experimental results in the previous subsections. The experiments were carried out using a set of datasets commonly used in fake news detection by applying an assortment of classical and advanced ML classifiers using different word representation methods as well as different transformer-based architectures. The results of the models on the LIAR dataset are tabulated in Table 9 and Table 10. The TF-IDF-based method for extracting text features performed better, with the ensemble approach exhibiting the highest accuracy while the CNN-LSTM/CNN-GRU hybrid model with GloVe embeddings had the highest F1 score. Despite their ability to capture automatic features, advanced ML models may produce unsatisfactory results when used individually at this level of analysis, where hybrid algorithms generally perform better. For instance, when a CNN and LSTM were combined with BERT b a s e , the classification accuracy was comparable to that of the best-performing model. As such, we agree with [80] that combining different models based on CNNs and recurrence-based models helps uncover useful patterns to discriminate fake from real content, where a model based on this combination would not only be able to extract the most relevant and salient local features but also capture long-term dependencies. Additionally, BERT b a s e alone achieved an accuracy result that was on par with the best-performing models, demonstrating the power of such a model in capturing useful clues from a short text sample. Moreover, we observed that the CNN network with context-free embedding models beat the recurrence-based models, whereas the opposite was true for the embedding models that took context into account.
Using the FakeNewsNet dataset, as shown in Table 11 and Table 12, we found that the transformer-based DL architectures outperformed the state-of-the-art approaches on the PolitiFact dataset. In contrast, classical ML models, such as SVM with TF-IDF features, beat the DL models on the GossipCop dataset, including the official baselines. On the COVID-19 dataset (see Table 13 and Table 14), BERT b a s e was ranked as the top-performing model, and CNN-BERT b a s e was the second-best model. The results show that transformer-based models are powerful in capturing the intricacies of such (noisy and sparse) datasets.

7. Conclusions

This paper presents an overview of fake news detection approaches found in the literature and compares seven classical ML algorithms, namely LR, SVM, NB, DT, RF, XGB and an ensemble that combines all such algorithms, with two scenarios of word representation methods: statistical (sparse) word vector representation methods and context-free (dense) pretrained word representation models. In addition, this paper compares eight advanced ML models, namely the CNN, BiLSTM, BiGRU, CNN-BiLSTM, CNN-BiGRU and different hybrid models with two types of text-representation models—context-free and context-aware embedding models—and two advanced pretrained transformer-based models: BERT b a s e and RoBERTa b a s e . We found that the ensemble of the classical ML methods with TF-IDF features outperformed the other methods on the LIAR dataset, including advanced ML models. As the LIAR dataset contains short political statements, it is difficult to obtain useful clues that could help discriminate fake news from real news. However, BERT b a s e alone was able to achieve an accuracy result that was on par with the best-performing models, demonstrating the power of such a model in capturing useful clues from a short piece of text. RoBERTa b a s e performed the best on the PolitiFact dataset with an F1 score of 93.17. In contrast, classical ML methods, such as SVM with TF-IDF features, beat the DL models on the GossipCop dataset, including the official baselines. For the COVID-19 dataset, the best-performing model was BERT b a s e . By relying only on news text with minimal text preprocessing, we achieved better results than the state-of-the-art results across the used datasets. It may be possible to improve the performance if context-based information, in addition to news text, is considered and if other factors, such as style and sentiment, are included. Generally, the experimental results indicate that no single technique can deliver the best performance scores across all used datasets.
To this end, we hope our comparative study will contribute to future research in this area and identify the best model for detecting fake news.

Author Contributions

Conceptualisation, experiments and formal analysis, J.A.; paper writing, J.A.; writing—review and editing, Y.L.; supervision, S.L. and Y.L. 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 LIAR data is available from [69], FakeNewsNet data is available from https://github.com/KaiDMML/FakeNewsNet (accessed on 20 March 2022) and COVID-19 data is available from [118].

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bondielli, A.; Marcelloni, F. A survey on fake news and rumour detection techniques. Inf. Sci. 2019, 497, 38–55. [Google Scholar] [CrossRef]
  2. Kwak, H.; Lee, C.; Park, H.; Moon, S. What is Twitter, a Social Network or a News Media? In Proceedings of the 19th International Conference on World Wide Web; Association for Computing Machinery: New York, NY, USA, 2010; pp. 591–600. [Google Scholar] [CrossRef] [Green Version]
  3. Allcott, H.; Gentzkow, M. Social media and fake news in the 2016 election. J. Econ. Perspect. 2017, 31, 211–236. [Google Scholar] [CrossRef] [Green Version]
  4. Shu, K.; Sliva, A.; Wang, S.; Tang, J.; Liu, H. Fake News Detection on Social Media: A Data Mining Perspective. arXiv 2017, arXiv:1708.01967. [Google Scholar] [CrossRef]
  5. Trends, G. “Fake News—Explore—Google Trends”. Available online: https://trends.google.com/trends/explore?date=2010-01-01%202022-07-14&q=%22fake%20news%22 (accessed on 20 July 2022).
  6. Langin, K. Fake news spreads faster than true news on Twitter—Thanks to people, not bots. Sci. Mag. 2018. Available online: https://www.science.org/content/article/fake-news-spreads-faster-true-news-twitter-thanks-people-not-bots (accessed on 20 February 2022).
  7. Zubiaga, A.; Aker, A.; Bontcheva, K.; Liakata, M.; Procter, R. Detection and Resolution of Rumours in Social Media: A Survey. ACM Comput. Surv. 2018, 51, 1–36. [Google Scholar] [CrossRef] [Green Version]
  8. Zhao, Z.; Resnick, P.; Mei, Q. Enquiring Minds: Early Detection of Rumors in Social Media from Enquiry Posts. In Proceedings of the 24th International Conference on World Wide Web; International World Wide Web Conferences Steering Committee: Geneva, Switzerland, 2015; pp. 1395–1405. [Google Scholar] [CrossRef]
  9. Vosoughi, S.; Roy, D.; Aral, S. The spread of true and false news online. Science 2018, 359, 1146–1151. [Google Scholar] [CrossRef]
  10. Kumar, S.; West, R.; Leskovec, J. Disinformation on the web: Impact, characteristics, and detection of wikipedia hoaxes. In Proceedings of the 25th International Conference on World Wide Web, Montreal, QC, Canada, 11–15 April 2016; pp. 591–602. [Google Scholar]
  11. Friggeri, A.; Adamic, L.; Eckles, D.; Cheng, J. Rumor cascades. In Proceedings of the International AAAI Conference on Web and Social Media, Ann Arbor, MI, USA, 1–4 June 2014; Volume 8. [Google Scholar]
  12. Zhou, X.; Zafarani, R. A Survey of Fake News: Fundamental Theories, Detection Methods, and Opportunities. ACM Comput. Surv. 2020, 53, 1–40. [Google Scholar] [CrossRef]
  13. Conroy, N.K.; Rubin, V.L.; Chen, Y. Automatic deception detection: Methods for finding fake news. Proc. Assoc. Inf. Sci. Technol. 2015, 52, 1–4. [Google Scholar] [CrossRef] [Green Version]
  14. Zhou, X.; Jain, A.; Phoha, V.V.; Zafarani, R. Fake News Early Detection: An Interdisciplinary Study. arXiv 2019, arXiv:1904.11679. [Google Scholar]
  15. Feng, S.; Banerjee, R.; Choi, Y. Syntactic Stylometry for Deception Detection. In Proceedings of the 50th Annual Meeting of the Association for Computational Linguistics (Volume 2: Short Papers); Association for Computational Linguistics: Jeju Island, Republic of Korea, 2012; pp. 171–175. [Google Scholar]
  16. Karimi, H.; Tang, J. Learning hierarchical discourse-level structure for fake news detection. arXiv 2019, arXiv:1903.07389. [Google Scholar]
  17. Pisarevskaya, D. Deception detection in news reports in the russian language: Lexics and discourse. In Proceedings of the 2017 EMNLP Workshop: Natural Language Processing Meets Journalism, Copenhagen, Denmark, 7 September 2017; pp. 74–79. [Google Scholar]
  18. Pérez-Rosas, V.; Kleinberg, B.; Lefevre, A.; Mihalcea, R. Automatic detection of fake news. arXiv 2017, arXiv:1708.07104. [Google Scholar]
  19. Chen, Y.; Conroy, N.J.; Rubin, V.L. Misleading Online Content: Recognizing Clickbait as “False News”. In Proceedings of the 2015 ACM on Workshop on Multimodal Deception Detection; Association for Computing Machinery: New York, NY, USA, 2015; pp. 15–19. [Google Scholar] [CrossRef]
  20. Potthast, M.; Kiesel, J.; Reinartz, K.; Bevendorff, J.; Stein, B. A stylometric inquiry into hyperpartisan and fake news. arXiv 2017, arXiv:1702.05638. [Google Scholar]
  21. Fuller, C.M.; Biros, D.P.; Wilson, R.L. Decision support for determining veracity via linguistic-based cues. Decis. Support Syst. 2009, 46, 695–703. [Google Scholar] [CrossRef]
  22. Zhou, L.; Burgoon, J.K.; Nunamaker, J.F.; Twitchell, D. Automating linguistics-based cues for detecting deception in text-based asynchronous computer-mediated communications. Group Decis. Negot. 2004, 13, 81–106. [Google Scholar] [CrossRef]
  23. Pennebaker, J.W.; Francis, M.E.; Booth, R.J. Linguistic inquiry and word count: LIWC 2001. Mahway Lawrence Erlbaum Assoc. 2001, 71, 2001. [Google Scholar]
  24. Riedel, B.; Augenstein, I.; Spithourakis, G.P.; Riedel, S. A simple but tough-to-beat baseline for the Fake News Challenge stance detection task. arXiv 2018, arXiv:cs.CL/1707.03264. [Google Scholar]
  25. Ahmed, H.; Traore, I.; Saad, S. Detection of online fake news using n-gram analysis and machine learning techniques. In Proceedings of the International Conference on Intelligent, Secure, and Dependable Systems in Distributed and Cloud Environments; Springer: Berlin/Heidelberg, Germany, 2017; pp. 127–138. [Google Scholar]
  26. Bharadwaj, P.; Shao, Z. Fake news detection with semantic features and text mining. Int. J. Nat. Lang. Comput. (IJNLC) 2019, 8, 17–22. [Google Scholar] [CrossRef]
  27. Wynne, H.E.; Wint, Z.Z. Content based fake news detection using n-gram models. In Proceedings of the 21st International Conference on Information Integration and Web-Based Applications & Services, Munich, Germany, 2–4 December 2019; pp. 669–673. [Google Scholar]
  28. Gravanis, G.; Vakali, A.; Diamantaras, K.; Karadais, P. Behind the cues: A benchmarking study for fake news detection. Expert Syst. Appl. 2019, 128, 201–213. [Google Scholar] [CrossRef]
  29. Burgoon, J.K.; Blair, J.P.; Qin, T.; Nunamaker, J.F. Detecting deception through linguistic analysis. In Proceedings of the International Conference on Intelligence and Security Informatics; Springer: Berlin/Heidelberg, Germany, 2003; pp. 91–101. [Google Scholar]
  30. Hancock, J.T.; Curry, L.E.; Goorha, S.; Woodworth, M. On lying and being lied to: A linguistic analysis of deception in computer-mediated communication. Discourse Process. 2007, 45, 1–23. [Google Scholar] [CrossRef]
  31. Newman, M.L.; Pennebaker, J.W.; Berry, D.S.; Richards, J.M. Lying Words: Predicting Deception from Linguistic Styles. Personal. Soc. Psychol. Bull. 2003, 29, 665–675. [Google Scholar] [CrossRef]
  32. Tausczik, Y.R.; Pennebaker, J.W. The psychological meaning of words: LIWC and computerized text analysis methods. J. Lang. Soc. Psychol. 2010, 29, 24–54. [Google Scholar] [CrossRef]
  33. Vrij, A. Detecting Lies and Deceit: The Psychology of Lying and Implications for Professional Practice; Wiley: Hoboken, NJ, USA, 2000. [Google Scholar]
  34. Knapp, M.L.; Hart, R.P.; Dennis, H.S. An exploration of deception as a communication construct. Hum. Commun. Res. 1974, 1, 15–29. [Google Scholar] [CrossRef]
  35. Horne, B.; Adali, S. This just in: Fake news packs a lot in title, uses simpler, repetitive content in text body, more similar to satire than real news. In Proceedings of the International AAAI Conference on Web and Social Media, Montreal, QC, Canada, 15–18 May 2017; Volume 11, pp. 759–766. [Google Scholar]
  36. Papadopoulou, O.; Zampoglou, M.; Papadopoulos, S.; Kompatsiaris, I. A two-level classification approach for detecting clickbait posts using text-based features. arXiv 2017, arXiv:1710.08528. [Google Scholar]
  37. Rubin, V.L.; Conroy, N.; Chen, Y.; Cornwell, S. Fake news or truth? using satirical cues to detect potentially misleading news. In Proceedings of the Second Workshop on Computational Approaches to Deception Detection, San Diego, CA, USA, 17 June 2016; pp. 7–17. [Google Scholar]
  38. Chowdhury, G.G. Natural language processing. Annu. Rev. Inf. Sci. Technol. 2003, 37, 51–89. [Google Scholar] [CrossRef] [Green Version]
  39. Qazvinian, V.; Rosengren, E.; Radev, D.R.; Mei, Q. Rumor has it: Identifying Misinformation in Microblogs. In Proceedings of the 2011 Conference on Empirical Methods in Natural Language Processing; Association for Computational Linguistics: Edinburgh, UK, 2011; pp. 1589–1599. [Google Scholar]
  40. Castillo, C.; Mendoza, M.; Poblete, B. Information Credibility on Twitter. In Proceedings of the 20th International Conference on World Wide Web; Association for Computing Machinery: New York, NY, USA, 2011; pp. 675–684. [Google Scholar] [CrossRef]
  41. Hamidian, S.; Diab, M.T. Rumor Detection and Classification for Twitter Data. arXiv 2019, arXiv:1912.08926. [Google Scholar]
  42. Yang, F.; Liu, Y.; Yu, X.; Yang, M. Automatic Detection of Rumor on Sina Weibo. In Proceedings of the ACM SIGKDD Workshop on Mining Data Semantics; Association for Computing Machinery: New York, NY, USA, 2012. [Google Scholar] [CrossRef]
  43. Ajao, O.; Bhowmik, D.; Zargari, S. Sentiment Aware Fake News Detection on Online Social Networks. In Proceedings of the ICASSP 2019—2019 IEEE International Conference on Acoustics, Speech and Signal Processing (ICASSP), Brighton, UK, 12–17 May 2019; pp. 2507–2511. [Google Scholar]
  44. Ratkiewicz, J.; Conover, M.; Meiss, M.; Gonçalves, B.; Flammini, A.; Menczer, F. Detecting and tracking political abuse in social media. In Proceedings of the International AAAI Conference on Web and Social Media, Barcelona, Spain, 17–21 July 2011; Volume 5. [Google Scholar]
  45. Samonte, M.J.C. Polarity analysis of editorial articles towards fake news detection. In Proceedings of the 2018 International Conference on Internet and e-Business, Singapore, 25–27 April 2018; pp. 108–112. [Google Scholar]
  46. Volkova, S.; Shaffer, K.; Jang, J.Y.; Hodas, N. Separating facts from fiction: Linguistic models to classify suspicious and trusted news posts on twitter. In Proceedings of the 55th Annual Meeting of the Association for Computational Linguistics (Volume 2: Short Papers), Vancouver, BC, Canada, 30 July–4 August 2017; pp. 647–653. [Google Scholar]
  47. Bhelande, M.; Sanadhya, A.; Purao, M.; Waldia, A.; Yadav, V. Identifying controversial news using sentiment analysis. Imp. J. Interdiscip. Res. 2017, 3. Available online: https://www.semanticscholar.org/paper/Identifying-Controversial-News-using-Sentiment-Bhelande-Sanadhya/23862325ff7b53e7851cd4398553d82cbca483f4 (accessed on 22 May 2022).
  48. Qin, Y.; Wurzer, D.; Lavrenko, V.; Tang, C. Spotting Rumors via Novelty Detection. arXiv 2016, arXiv:1611.06322. [Google Scholar]
  49. Kwon, S.; Cha, M.; Jung, K. Rumor Detection over Varying Time Windows. PLoS ONE 2017, 12, e0168344. [Google Scholar] [CrossRef] [Green Version]
  50. Wei, W.; Wan, X. Learning to identify ambiguous and misleading news headlines. arXiv 2017, arXiv:1705.06031. [Google Scholar]
  51. Chakraborty, A.; Paranjape, B.; Kakarla, S.; Ganguly, N. Stop Clickbait: Detecting and Preventing Clickbaits in Online News Media. arXiv 2016, arXiv:1610.09786. [Google Scholar]
  52. Feng, V.W.; Hirst, G. Detecting Deceptive Opinions with Profile Compatibility. In Proceedings of the Sixth International Joint Conference on Natural Language Processing; Asian Federation of Natural Language Processing: Nagoya, Japan, 2013; pp. 338–346. [Google Scholar]
  53. Potthast, M.; Köpsel, S.; Stein, B.; Hagen, M. Clickbait Detection. In Proceedings of the ECIR, Padua, Italy, 20–23 March 2016. [Google Scholar]
  54. Gupta, A.; Kumaraguru, P.; Castillo, C.; Meier, P. Tweetcred: Real-time credibility assessment of content on twitter. In Proceedings of the International Conference on Social Informatics; Springer: Berlin/Heidelberg, Germany, 2014; pp. 228–243. [Google Scholar]
  55. Bhattacharjee, S.D.; Talukder, A.; Balantrapu, B.V. Active learning based news veracity detection with feature weighting and deep-shallow fusion. In Proceedings of the 2017 IEEE International Conference on Big Data (Big Data), Boston, MA, USA, 11–14 December 2017; IEEE: Piscataway, NJ, USA, 2017; pp. 556–565. [Google Scholar]
  56. Ribeiro, M.H.; Calais, P.H.; Almeida, V.A.; Meira, W., Jr. “Everything I Disagree With is# FakeNews”: Correlating Political Polarization and Spread of Misinformation. arXiv 2017, arXiv:1706.05924. [Google Scholar]
  57. Popat, K. Assessing the Credibility of Claims on the Web. In Proceedings of the 26th International Conference on World Wide Web Companion; International World Wide Web Conferences Steering Committee: Geneva, Switzerland, 2017; pp. 735–739. [Google Scholar] [CrossRef] [Green Version]
  58. Cardoso Durier da Silva, F.; Vieira, R.; Garcia, A.C. Can machines learn to detect fake news? A survey focused on social media. In Proceedings of the 52nd Hawaii International Conference on System Sciences, Maui, HI, USA, 8–11 January 2019. [Google Scholar]
  59. Markowitz, D.M.; Hancock, J.T. Linguistic Traces of a Scientific Fraud: The Case of Diederik Stapel. PLoS ONE 2014, 9, e105937. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Ruchansky, N.; Seo, S.; Liu, Y. CSI: A Hybrid Deep Model for Fake News Detection. In Proceedings of the 2017 ACM on Conference on Information and Knowledge Management; Association for Computing Machinery: New York, NY, USA, 2017; pp. 797–806. [Google Scholar]
  61. Lahby, M.; Aqil, S.; Yafooz, W.M.S.; Abakarim, Y. Online Fake News Detection Using Machine Learning Techniques: A Systematic Mapping Study. In Combating Fake News with Computational Intelligence Techniques; Lahby, M., Pathan, A.S.K., Maleh, Y., Yafooz, W.M.S., Eds.; Springer International Publishing: Cham, Switzerland, 2022; pp. 3–37. [Google Scholar] [CrossRef]
  62. Klyuev, V. Fake news filtering: Semantic approaches. In Proceedings of the 2018 7th International Conference on Reliability, Infocom Technologies and Optimization (Trends and Future Directions) (ICRITO), Noida, India, 29–31 August 2018; IEEE: Piscataway, NJ, USA, 2018; pp. 9–15. [Google Scholar]
  63. Mikolov, T.; Chen, K.; Corrado, G.; Dean, J. Efficient estimation of word representations in vector space. arXiv 2013, arXiv:1301.3781. [Google Scholar]
  64. Pennington, J.; Socher, R.; Manning, C.D. Glove: Global vectors for word representation. In Proceedings of the 2014 Conference on Empirical Methods in Natural Language Processing (EMNLP), Doha, Qatar, 25–29 October 2014; pp. 1532–1543. [Google Scholar]
  65. Arora, S.; Liang, Y.; Ma, T. A simple but tough-to-beat baseline for sentence embeddings. In Proceedings of the International Conference on Learning Representations, Toulon, France, 24–26 April 2017. [Google Scholar]
  66. Le, Q.; Mikolov, T. Distributed representations of sentences and documents. In Proceedings of the International Conference on Machine Learning, PMLR, Beijing, China, 21–26 June 2014; pp. 1188–1196. [Google Scholar]
  67. Zubiaga, A.; Liakata, M.; Procter, R. Exploiting context for rumour detection in social media. In Proceedings of the International Conference on Social Informatics; Springer: Berlin/Heidelberg, Germany, 2017; pp. 109–123. [Google Scholar]
  68. Mikolov, T.; Sutskever, I.; Chen, K.; Corrado, G.; Dean, J. Distributed Representations of Words and Phrases and Their Compositionality. In Proceedings of the 26th International Conference on Neural Information Processing Systems; Curran Associates Inc.: Red Hook, NY, USA, 2013; Volume 2, pp. 3111–3119. [Google Scholar]
  69. Wang, W.Y. “Liar, Liar Pants on Fire”: A New Benchmark Dataset for Fake News Detection. arXiv 2017, arXiv:1705.00648. [Google Scholar]
  70. Qian, F.; Gong, C.; Sharma, K.; Liu, Y. Neural User Response Generator: Fake News Detection with Collective User Intelligence. In Proceedings of the IJCAI, Stockholm, Sweden, 13–19 July 2018; Volume 18, pp. 3834–3840. [Google Scholar]
  71. Goldani, M.H.; Momtazi, S.; Safabakhsh, R. Detecting fake news with capsule neural networks. Appl. Soft Comput. 2021, 101, 106991. [Google Scholar] [CrossRef]
  72. Girgis, S.; Amer, E.; Gadallah, M. Deep learning algorithms for detecting fake news in online text. In Proceedings of the 2018 13th International Conference on Computer Engineering and Systems (ICCES), Cairo, Egypt, 18–19 December 2018; IEEE: Piscataway, NJ, USA, 2018; pp. 93–97. [Google Scholar]
  73. Sarnovskỳ, M.; Maslej-Krešňáková, V.; Ivancová, K. Fake News Detection Related to the COVID-19 in Slovak Language Using Deep Learning Methods. Acta Polytech. Hung. 2022, 19, 43–57. [Google Scholar] [CrossRef]
  74. Devlin, J.; Chang, M.W.; Lee, K.; Toutanova, K. Bert: Pre-training of deep bidirectional transformers for language understanding. arXiv 2018, arXiv:1810.04805. [Google Scholar]
  75. Peters, M.E.; Neumann, M.; Iyyer, M.; Gardner, M.; Clark, C.; Lee, K.; Zettlemoyer, L. Deep contextualized word representations. arXiv 2018, arXiv:1802.05365. [Google Scholar]
  76. Radford, A.; Narasimhan, K.; Salimans, T.; Sutskever, I. Improving Language Understanding by Generative Pre-Training. 2018. Available online: https://www.cs.ubc.ca/~amuham01/LING530/papers/radford2018improving.pdf (accessed on 10 August 2022).
  77. Müller, M.; Salathé, M.; Kummervold, P.E. Covid-twitter-bert: A natural language processing model to analyse COVID-19 content on twitter. arXiv 2020, arXiv:2005.07503. [Google Scholar]
  78. Liu, Y.; Ott, M.; Goyal, N.; Du, J.; Joshi, M.; Chen, D.; Levy, O.; Lewis, M.; Zettlemoyer, L.; Stoyanov, V. Roberta: A robustly optimized bert pretraining approach. arXiv 2019, arXiv:1907.11692. [Google Scholar]
  79. Kula, S.; Choraś, M.; Kozik, R. Application of the bert-based architecture in fake news detection. In Proceedings of the Computational Intelligence in Security for Information Systems Conference; Springer: Berlin/Heidelberg, Germany, 2019; pp. 239–249. [Google Scholar]
  80. Alghamdi, J.; Lin, Y.; Luo, S. Modeling Fake News Detection Using BERT-CNN-BiLSTM Architecture. In Proceedings of the 2022 IEEE 5th International Conference on Multimedia Information Processing and Retrieval (MIPR), Online, 2–4 August 2022; pp. 354–357. [Google Scholar] [CrossRef]
  81. Aggarwal, A.; Chauhan, A.; Kumar, D.; Mittal, M.; Verma, S. Classification of fake news by fine-tuning deep bidirectional transformers based language model. EAI Endorsed Trans. Scalable Inf. Syst. 2020, 7, e10. [Google Scholar] [CrossRef] [Green Version]
  82. Jwa, H.; Oh, D.; Park, K.; Kang, J.M.; Lim, H. exBAKE: Automatic fake news detection model based on bidirectional encoder representations from transformers (bert). Appl. Sci. 2019, 9, 4062. [Google Scholar] [CrossRef]
  83. Baruah, A.; Das, K.A.; Barbhuiya, F.A.; Dey, K. Automatic Detection of Fake News Spreaders Using BERT. In Proceedings of the CLEF (Working Notes), Thessaloniki, Greece, 22–25 September 2020. [Google Scholar]
  84. Sanh, V.; Debut, L.; Chaumond, J.; Wolf, T. DistilBERT, a distilled version of BERT: Smaller, faster, cheaper and lighter. arXiv 2019, arXiv:1910.01108. [Google Scholar]
  85. Khan, J.Y.; Khondaker, M.T.I.; Afroz, S.; Uddin, G.; Iqbal, A. A benchmark study of machine learning models for online fake news detection. Mach. Learn. Appl. 2021, 4, 100032. [Google Scholar] [CrossRef]
  86. Elhadad, M.K.; Li, K.F.; Gebali, F. A Novel Approach for Selecting Hybrid Features from Online News Textual Metadata for Fake News Detection. In Proceedings of the Advances on P2P, Parallel, Grid, Cloud and Internet Computing; Barolli, L., Hellinckx, P., Natwichai, J., Eds.; Springer International Publishing: Cham, Switzerland, 2020; pp. 914–925. [Google Scholar]
  87. Shu, K.; Mahudeswaran, D.; Wang, S.; Lee, D.; Liu, H. FakeNewsNet: A Data Repository with News Content, Social Context and Spatialtemporal Information for Studying Fake News on Social Media. arXiv 2018, arXiv:1809.01286. [Google Scholar] [CrossRef]
  88. Shu, K.; Cui, L.; Wang, S.; Lee, D.; Liu, H. DEFEND: Explainable Fake News Detection. In Proceedings of the 25th ACM SIGKDD International Conference on Knowledge Discovery & Data Mining; Association for Computing Machinery: New York, NY, USA, 2019; pp. 395–405. [Google Scholar] [CrossRef]
  89. Koloski, B.; Perdih, T.S.; Robnik-Šikonja, M.; Pollak, S.; Škrlj, B. Knowledge graph informed fake news classification via heterogeneous representation ensembles. Neurocomputing 2022, 496, 208–226. [Google Scholar] [CrossRef]
  90. Oriola, O. Exploring N-gram, word embedding and topic models for content-based fake news detection in FakeNewsNet evaluation. Int. J. Comput. Appl. 2021, 975, 8887. [Google Scholar] [CrossRef]
  91. Sadeghi, F.; Jalaly Bidgoly, A.; Amirkhani, H. Fake News Detection on Social Media Using A Natural Language Inference Approach. Multimed. Tools Appl. 2020. [Google Scholar] [CrossRef]
  92. Gautam, A.; Venktesh, V.; Masud, S. Fake News Detection System using XLNet model with Topic Distributions: CONSTRAINT@AAAI2021 Shared Task. arXiv 2021, arXiv:2101.11425. [Google Scholar] [CrossRef]
  93. Shifath, S.M.S.U.R.; Khan, M.F.; Islam, M.S. A transformer based approach for fighting COVID-19 fake news. arXiv 2021, arXiv:2101.12027. [Google Scholar] [CrossRef]
  94. Veyseh, A.P.B.; Thai, M.T.; Nguyen, T.H.; Dou, D. Rumor detection in social networks via deep contextual modeling. In Proceedings of the 2019 IEEE/ACM International Conference on Advances in Social Networks Analysis and Mining, Vancouver, BC, Canada, 27–30 August 2019; pp. 113–120. [Google Scholar]
  95. Wani, A.; Joshi, I.; Khandve, S.; Wagh, V.; Joshi, R. Evaluating Deep Learning Approaches for COVID-19 Fake News Detection. In Combating Online Hostile Posts in Regional Languages during Emergency Situation; Springer International Publishing: Berlin/Heidelberg, Germany, 2021; pp. 153–163. [Google Scholar] [CrossRef]
  96. Shushkevich, E.; Cardiff, J. TUDublin team at Constraint@ AAAI2021–COVID19 Fake News Detection. arXiv 2021, arXiv:2101.05701. [Google Scholar]
  97. Felber, T. Constraint 2021: Machine learning models for COVID-19 fake news detection shared task. arXiv 2021, arXiv:2101.03717. [Google Scholar]
  98. Patwa, P.; Sharma, S.; Pykl, S.; Guptha, V.; Kumari, G.; Akhtar, M.S.; Ekbal, A.; Das, A.; Chakraborty, T. Fighting an infodemic: COVID-19 fake news dataset. In Proceedings of the International Workshop on Combating Online Hostile Posts in Regional Languages during Emergency Situation; Springer: Berlin/Heidelberg, Germany, 2021; pp. 21–29. [Google Scholar]
  99. Goldberg, Y. Neural network methods for natural language processing. Synth. Lect. Hum. Lang. Technol. 2017, 10, 1–309. [Google Scholar]
  100. Wallach, H.M. Topic modeling: Beyond bag-of-words. In Proceedings of the 23rd International Conference on Machine Learning, Pittsburgh, PA, USA, 25–29 June 2006; pp. 977–984. [Google Scholar]
  101. Damashek, M. Gauging similarity with n-grams: Language-independent categorization of text. Science 1995, 267, 843–848. [Google Scholar] [CrossRef] [PubMed]
  102. Joachims, T. A Probabilistic Analysis of the Rocchio Algorithm with TFIDF for Text Categorization. Technical Report, Carnegie-Mellon univ Pittsburgh pa Dept of Computer Science. 1996. Available online: https://www.cs.cornell.edu/people/tj/publications/joachims_97a.pdf (accessed on 11 January 2022).
  103. McCallum, A.; Nigam, K. A comparison of event models for naive bayes text classification. In Proceedings of the AAAI-98 Workshop on Learning for Text Categorization; Citeseer: State College, PA, USA, 1998; Volume 752, pp. 41–48. [Google Scholar]
  104. Trstenjak, B.; Mikac, S.; Donko, D. KNN with TF-IDF based framework for text categorization. Procedia Eng. 2014, 69, 1356–1364. [Google Scholar] [CrossRef] [Green Version]
  105. Joachims, T. Text categorization with support vector machines: Learning with many relevant features. In Proceedings of the European Conference on Machine Learning; Springer: Berlin/Heidelberg, Germany, 1998; pp. 137–142. [Google Scholar]
  106. Harris, Z.S. Distributional structure. Word 1954, 10, 146–162. [Google Scholar] [CrossRef]
  107. Salton, G. MCGILL, Michael. In Introduction to Modern Information Retrieval; McGraw-Hill, Inc.: New York, NY, USA, 1986. [Google Scholar]
  108. Schütze, H.; Manning, C.D.; Raghavan, P. Introduction to Information Retrieval; Cambridge University Press: Cambridge, UK, 2008; Volume 39. [Google Scholar]
  109. Mikolov, T.; Yih, W.t.; Zweig, G. Linguistic regularities in continuous space word representations. In Proceedings of the 2013 Conference of the North American Chapter of the Association for Computational Linguistics: Human Language Technologies, Atlanta, GA, USA, 9–14 June 2013; pp. 746–751. [Google Scholar]
  110. Howard, J.; Ruder, S. Universal language model fine-tuning for text classification. arXiv 2018, arXiv:1801.06146. [Google Scholar]
  111. Hochreiter, S.; Schmidhuber, J. Long short-term memory. Neural Comput. 1997, 9, 1735–1780. [Google Scholar] [CrossRef]
  112. Tang, D.; Qin, B.; Liu, T. Document Modeling with Gated Recurrent Neural Network for Sentiment Classification. In Proceedings of the 2015 Conference on Empirical Methods in Natural Language Processing, Lisbon, Portugal, 17–21 September 2015; pp. 1422–1432. [Google Scholar]
  113. Graves, A. Supervised sequence labelling. In Supervised Sequence Labelling with Recurrent Neural Networks; Springer: Berlin/Heidelberg, Germany, 2012; pp. 5–13. [Google Scholar]
  114. Cho, K.; van Merrienboer, B.; Gulcehre, C.; Bahdanau, D.; Bougares, F.; Schwenk, H.; Bengio, Y. Learning Phrase Representations using RNN Encoder-Decoder for Statistical Machine Translation. arXiv 2014, arXiv:1406.1078. [Google Scholar]
  115. Zhang, X.; Chen, F.; Huang, R. A combination of RNN and CNN for attention-based relation classification. Procedia Comput. Sci. 2018, 131, 911–917. [Google Scholar] [CrossRef]
  116. Zhou, C.; Sun, C.; Liu, Z.; Lau, F. A C-LSTM neural network for text classification. arXiv 2015, arXiv:1511.08630. [Google Scholar]
  117. Gururangan, S.; Marasović, A.; Swayamdipta, S.; Lo, K.; Beltagy, I.; Downey, D.; Smith, N.A. Don’t stop pretraining: Adapt language models to domains and tasks. arXiv 2020, arXiv:2004.10964. [Google Scholar]
  118. Patwa, P.; Bhardwaj, M.; Guptha, V.; Kumari, G.; Sharma, S.; Pykl, S.; Das, A.; Ekbal, A.; Akhtar, M.S.; Chakraborty, T. Overview of constraint 2021 shared tasks: Detecting english COVID-19 fake news and hindi hostile posts. In Proceedings of the International Workshop on Combating Online Hostile Posts in Regional Languages during Emergency Situation; Springer: Berlin/Heidelberg, Germany, 2021; pp. 42–53. [Google Scholar]
  119. Horne, L.; Matti, M.; Pourjafar, P.; Wang, Z. GRUBERT: A GRU-Based Method to Fuse BERT Hidden Layers for Twitter Sentiment Analysis. In Proceedings of the 1st Conference of the Asia-Pacific Chapter of the Association for Computational Linguistics and the 10th International Joint Conference on Natural Language Processing: Student Research Workshop; Association for Computational Linguistics: Suzhou, China, 2020; pp. 130–138. [Google Scholar]
Figure 1. Fake news trends (2019–2022) [5].
Figure 1. Fake news trends (2019–2022) [5].
Information 13 00576 g001
Figure 2. The structure of advanced ML models.
Figure 2. The structure of advanced ML models.
Information 13 00576 g002
Table 1. A summary of exiting work on the LIAR dataset.
Table 1. A summary of exiting work on the LIAR dataset.
ModelsMetrics
AccuracyPrecisionRecallF1
RoBERTa b a s e  [85]0.620.630.620.62
SVM [86]0.62NANANA
Table 2. A summary of exiting work on (FakeNewsNet) PolitiFact dataset.
Table 2. A summary of exiting work on (FakeNewsNet) PolitiFact dataset.
ModelsMetrics
AccuracyPrecisionRecallF1
Social Article Fusion [87]0.690.640.790.71
Logistic Regression (N-Gram) [90]0.800.790.780.78
BiLSTM-BERT [91]0.8558NANANA
LNN-KG [89]0.8800.90110.8800.8892
DEFEND [88]0.9040.9020.9560.928
Table 3. A summary of exiting work on (FakeNewsNet) GossipCop dataset.
Table 3. A summary of exiting work on (FakeNewsNet) GossipCop dataset.
ModelsMetrics
AccuracyPrecisionRecallF1
CNN [87]0.7230.7510.7010.725
Logistic Regression (N-Gram) [90]0.820.750.790.77
DEFEND [88]0.8080.7290.7820.755
Table 4. A summary of exiting work on the COVID-19 dataset.
Table 4. A summary of exiting work on the COVID-19 dataset.
ModelsMetrics
AccuracyPrecisionRecallF1
SVM+LR+NB+biLSTM [96]NANANA0.94
SVM [97]0.95700.95710.95700.9570
SNN(LM+KG) [89]0.95700.95330.96520.9569
BERT-Based [95]0.9841NANANA
Ensemble Transformer Models [93]0.97990.97990.97990.9799
XLNet with Topic Distributions [92]NA0.9680.9670.967
SVM [98]0.93320.93330.93320.9332
Table 5. The hyperparameters of the advanced ML models.
Table 5. The hyperparameters of the advanced ML models.
ModelHyperparameters
Hidden
Layer(s)
Activation
Function
Neurons or
Filters
Kernel
Size
Dropout
CNN1ReLU12850.3
BiLSTM1N/A128N/A0.3
BiGRU1N/A128N/A0.3
Table 6. The statistics of the LIAR dataset.
Table 6. The statistics of the LIAR dataset.
# Candidate News12,791
# True news7134
# Fake news5657
Table 7. The statistics of the FakeNewsNet dataset.
Table 7. The statistics of the FakeNewsNet dataset.
DatasetPolitiFactGossipCop
# Candidate news69418,676
# True news35614,129
# Fake news3384547
Table 8. The statistics of the COVID-19 dataset.
Table 8. The statistics of the COVID-19 dataset.
# Candidate News6420
# True news3360
# Fake news3060
Table 9. Fake news detection using classical ML classification models on LIAR dataset.
Table 9. Fake news detection using classical ML classification models on LIAR dataset.
Dataset
LIAR
ModelFeatureA (%)P (%)R (%)F1 (%)
LRCV0.61960.65300.69330.6726
SVMCV0.63460.64410.78570.7079
MNBCV0.62430.65300.71150.6810
DTCV0.5730.0.61370.65410.6332
RFCV0.62750.63600.79270.7057
XGBCV0.60930.60850.85990.7127
EnsembleCV0.63060.65040.74510.6945
LRTFIDF0.63540.64930.76750.8655
SVMTFIDF0.61010.64210.69610.6680
MNBTFIDF0.60930.63110.73810.6804
DTTFIDF0.54140.59170.60080.5962
RFTFIDF0.56590.56580.98740.7194
XGBTFIDF0.60930.61210.83750.7073
EnsembleTFIDF0.63930.65030.77870.7087
LRWord2Vec0.63610.64620.78290.7080
SVMWord2Vec0.63140.64380.77450.7031
MNBWord2Vec0.57620.57780.92020.7099
DTWord2Vec0.54930.60320.58540.5942
RFWord2Vec0.56750.57070.93840.7097
XGBWord2Vec0.61560.63000.77030.6931
EnsembleWord2Vec0.62670.63550.79130.7049
LRGloVe0.61720.62650.79410.7004
SVMGloVe0.62120.63060.79130.7019
MNGloVe0.56350.56350.56350.7208
DTGloVe0.51780.57240.57000.5712
RFGloVe0.60060.60420.84450.7044
XGBGloVe0.59430.61090.77170.6819
EnsembleGloVe0.59830.60990.79690.6910
Table 10. Fake news detection using advanced ML and DL classification models on LIAR dataset.
Table 10. Fake news detection using advanced ML and DL classification models on LIAR dataset.
Dataset
LIAR
ModelFeatureA (%)P (%)R (%)F1 (%)
CNNWord2Vec0.58250.62190.66110.6409
BiLSTMWord2Vec0.55330.60280.60780.6053
BiGRUWord2Vec0.55720.60730.60640.6069
CNN-LSTMWord2Vec0.56350.56351.00000.7208
CNN-GRUWord2Vec0.56270.56320.99860.7202
CNN-BiLSTMWord2Vec0.57620.62810.60780.6178
CNN-BiGRUWord2Vec0.57140.60670.68070.6416
HybridWord2Vec0.57850.61750.66250.6392
CNNGloVe0.61720.63880.73810.6849
BiLSTMGloVe0.58880.64420.60360.6233
BiGRUGloVe0.59270.63710.64430.6407
CNN-LSTMGloVe0.56350.56351.00000.7208
CNN-GRUGloVe0.56350.56351.00000.7208
CNN-BiLSTMGloVe0.59120.63650.64010.6383
CNN-BiGRUGloVe0.60140.62400.73670.6757
HybridGloVe0.58090.63320.60920.6210
CNNBERT b a s e 0.59750.65550.60220.6277
BiLSTMBERT b a s e 0.62040.63920.74930.6899
BiGRUBERT b a s e 0.61800.63860.74230.6865
CNN-LSTMBERT b a s e 0.63060.66020.71010.6842
CNN-GRUBERT b a s e 0.60770.65430.64430.6493
CNN-BiLSTMBERT b a s e 0.57930.61000.70310.6532
CNN-BiGRUBERT b a s e 0.61400.63130.75770.6887
HybridBERT b a s e 0.59120.61950.71150.6623
BERT b a s e BERT b a s e 0.63060.66620.69050.6781
RoBERTa b a s e RoBERTa b a s e 0.61170.64680.68490.6653
Table 11. Fake news detection using classical ML classification models on FakeNewsNet dataset.
Table 11. Fake news detection using classical ML classification models on FakeNewsNet dataset.
Datasets
PolitiFactGossipCop
ModelFeatureA (%)P (%)R (%)F1 (%)A (%)P (%)R (%)F1 (%)
LRCV0.83110.73740.97710.84050.83750.86440.93330.8976
SVMCV0.83300.72480.99780.83970.81830.81360.98810.8924
MNBCV0.72200.53931.00000.70070.83350.84170.96280.8982
DTCV0.74190.80500.77580.79010.78750.86070.86070.8607
RFCV0.81310.69180.99770.81710.85360.85050.98030.9108
XGBCV0.85480.77200.98400.86520.84290.84420.97370.9043
EnsembleCV0.89931.00000.82500.90410.85280.85210.97650.9100
LRTFIDF0.85670.99790.76420.86550.85870.86050.97230.9130
SVMTFIDF0.84160.98550.74840.85080.86350.86670.97020.9155
MNBTFIDF0.75240.99730.59120.74230.82120.85740.91820.8868
DTTFIDF0.68410.82720.60220.69700.78610.86120.85780.8595
RFTFIDF0.79601.0000.66190.79660.84960.84350.98560.9090
XGBTFIDF0.87480.98460.80500.88580.84100.84450.97020.9030
EnsembleTFIDF0.87770.95650.82500.88590.86110.86500.96910.9141
LRWord2Vec0.78840.65250.99520.78820.83830.84210.96980.9015
SVMWord2Vec0.79220.66510.98600.79440.83670.83250.98390.9019
MNBWord2Vec0.65750.97910.44180.60890.76260.76261.00000.8653
DTWord2Vec0.69070.53930.91220.67790.75780.85290.82450.8385
RFWord2Vec0.69070.49840.97840.66040.83940.83880.97720.9027
XGBWord2Vec0.69830.50470.99070.66880.83320.84060.96420.8982
EnsembleWord2Vec0.88490.93240.86250.89610.84100.83950.97860.9037
LRGloVe0.69920.98190.51100.67220.76280.76320.99890.8653
SVMGloVe0.72390.98590.55030.70640.76260.76261.00000.8653
MNGloVe0.65750.97580.44340.60970.76260.76261.00000.8653
DTGloVe0.62620.81510.49210.61370.72300.82650.80590.8161
RFGloVe0.65460.97890.43710.60430.82360.82280.97960.8944
XGBGloVe0.71060.98530.52830.68780.80620.81370.96740.8839
EnsembleGloVe0.79860.91940.71250.80280.79440.79120.99230.8804
Table 12. Fake news detection using advanced ML and DL classification models on FakeNewsNet dataset.
Table 12. Fake news detection using advanced ML and DL classification models on FakeNewsNet dataset.
Datasets
PolitiFactGossipCop
ModelFeatureA (%)P (%)R (%)F1 (%)A (%)P (%)R (%)F1 (%)
CNNWord2Vec0.64750.91890.42500.58120.84150.85660.95160.9016
BiLSTMWord2Vec0.77700.91530.67500.77700.82390.87020.90380.8867
BiGRUWord2Vec0.75540.81940.73750.77630.82630.86160.92000.8898
CNN-LSTMWord2Vec0.72660.70590.90000.79120.83030.87000.91400.8915
CNN-GRUWord2Vec0.66910.68480.78750.73260.82660.86940.90910.8888
CNN-BiLSTMWord2Vec0.70500.75320.72500.73890.82600.86830.90980.8886
CNN-BiGRUWord2Vec0.73380.74170.82500.78110.80970.87910.87010.8746
HybridWord2Vec0.70500.81970.62500.70920.82980.86390.92210.8920
CNNGloVe0.85610.87500.87500.87500.80970.89680.84800.8717
BiLSTMGloVe0.81290.89710.76250.82430.82840.85480.93370.8925
BiGRUGloVe0.82010.85710.82500.84080.82710.85800.92660.8910
CNN-LSTMGloVe0.75540.83820.71250.77030.81850.87090.89470.8826
CNN-GRUGloVe0.71940.73560.80000.76650.81370.86880.89010.8793
CNN-BiLSTMGloVe0.79860.86110.77500.81580.82900.85900.92800.8922
CNN-BiGRUGloVe0.81290.83750.83750.83750.81800.86750.89860.8828
HybridGloVe0.80850.83540.82500.83020.83490.85980.93610.8963
CNNBERT b a s e 0.89930.93420.88750.91030.86160.86840.97020.9145
BiLSTMBERT b a s e 0.90650.95890.87500.91500.83910.87690.91790.8969
BiGRUBERT b a s e 0.90650.91360.92500.91930.85330.87540.94170.9073
CNN-LSTMBERT b a s e 0.85610.87500.87500.87500.84180.87290.92770.8994
CNN-GRUBERT b a s e 0.84170.85370.87500.86420.85140.87840.93960.9061
CNN-BiLSTMBERT b a s e 0.89930.91250.91250.91250.83700.87560.91650.8956
CNN-BiGRUBERT b a s e 0.87770.94370.83750.88740.83220.89150.88800.8897
HybridBERT b a s e 0.92090.96000.90000.92900.84370.87190.93190.9009
BERT b a s e BERT b a s e 0.91370.93590.91250.92410.85440.87240.94770.9085
RoBERTa b a s e RoBERTa b a s e 0.92090.92590.93750.93170.83830.87940.91330.8960
Table 13. Fake news detection using classical ML classification models on COVID-19 dataset.
Table 13. Fake news detection using classical ML classification models on COVID-19 dataset.
Dataset
COVID-19
ModelFeatureA (%)P (%)R (%)F1 (%)
LRCV0.93130.93630.93210.9342
SVMCV0.93180.93330.93660.9349
MNBCV0.90510.91270.90540.9090
DTCV0.88740.89210.89290.8925
RFCV0.92710.92210.94020.9310
XGBCV0.89440.89420.90540.8997
EnsembleCV0.93270.94440.92590.9351
LRTFIDF0.92940.92090.94640.9335
SVMTFIDF0.94300.93540.95710.9462
MNBTFIDF0.91870.90850.93930.9236
DTTFIDF0.86960.87780.87230.8751
RFTFIDF0.86540.84670.90710.8759
XGBTFIDF0.88830.89300.89380.8934
EnsembleTFIDF0.93600.94080.93660.9387
LRWord2Vec0.90280.90500.90980.9074
SVMWord2Vec0.90510.90180.91870.9102
MNBWord2Vec0.83880.80830.90710.8549
DTWord2Vec0.79160.79930.80360.8014
RFWord2Vec0.87340.87940.87860.8790
XGBWord2Vec0.89160.89430.89910.8967
EnsembleWord2Vec0.90230.91750.89380.9055
LRGloVe0.81820.81700.84110.8289
SVMGloVe0.81730.81450.84290.8284
MNGloVe0.64860.60240.69610.7421
DTGloVe0.73790.74500.75890.7519
RFGloVe0.82800.81440.86960.8411
XGBGloVe0.84160.83810.86430.8510
EnsembleGloVe0.84490.84930.85540.8523
Table 14. Fake news detection using advanced ML and DL classification models on COVID-19 dataset.
Table 14. Fake news detection using advanced ML and DL classification models on COVID-19 dataset.
Dataset
COVID-19
ModelFeatureA (%)P (%)R (%)F1 (%)
CNNWord2Vec0.91870.90220.94730.9242
BiLSTMWord2Vec0.92570.93090.92680.9289
BiGRUWord2Vec0.92940.92460.94200.9332
CNN-LSTMWord2Vec0.92900.92910.93570.9324
CNN-GRUWord2Vec0.93320.93420.93840.9363
CNN-BiLSTMWord2Vec0.93040.93620.93040.9333
CNN-BiGRUWord2Vec0.90700.96560.85270.9056
HybridWord2Vec0.92940.91450.95450.9340
CNNGloVe0.95420.95540.95710.9563
BiLSTMGloVe0.93550.91610.96520.9400
BiGRUGloVe0.94210.92200.97140.9461
CNN-LSTMGloVe0.85930.79910.97680.8791
CNN-GRUGloVe0.52340.52340.99910.6869
CNN-BiLSTMGloVe0.94770.93080.97230.9511
CNN-BiGRUGloVe0.94720.93820.96250.9502
HybridGloVe0.93740.91430.97140.9420
CNNBERT b a s e 0.97520.97250.98040.9764
BiLSTMBERT b a s e 0.96500.94850.98660.9672
BiGRUBERT b a s e 0.97060.96160.98300.9722
CNN-LSTMBERT b a s e 0.95980.94340.98210.9624
CNN-GRUBERT b a s e 0.96260.97100.95710.9640
CNN-BiLSTMBERT b a s e 0.96210.95290.97590.9643
CNN-BiGRUBERT b a s e 0.95650.94850.96960.9589
HybridBERT b a s e 0.96170.96420.96250.9634
BERT b a s e BERT b a s e 0.97710.97350.98300.9782
RoBERTa b a s e RoBERTa b a s e 0.96680.95410.98390.9688
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Alghamdi, J.; Lin, Y.; Luo, S. A Comparative Study of Machine Learning and Deep Learning Techniques for Fake News Detection. Information 2022, 13, 576. https://doi.org/10.3390/info13120576

AMA Style

Alghamdi J, Lin Y, Luo S. A Comparative Study of Machine Learning and Deep Learning Techniques for Fake News Detection. Information. 2022; 13(12):576. https://doi.org/10.3390/info13120576

Chicago/Turabian Style

Alghamdi, Jawaher, Yuqing Lin, and Suhuai Luo. 2022. "A Comparative Study of Machine Learning and Deep Learning Techniques for Fake News Detection" Information 13, no. 12: 576. https://doi.org/10.3390/info13120576

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