• Open supplemental data
  • Reference Manager
  • Simple TEXT file

People also looked at

Original research article, explainable automated essay scoring: deep learning really has pedagogical value.

automated essay scoring dataset

  • School of Computing and Information Systems, Faculty of Science and Technology, Athabasca University, Edmonton, AB, Canada

Automated essay scoring (AES) is a compelling topic in Learning Analytics for the primary reason that recent advances in AI find it as a good testbed to explore artificial supplementation of human creativity. However, a vast swath of research tackles AES only holistically; few have even developed AES models at the rubric level, the very first layer of explanation underlying the prediction of holistic scores. Consequently, the AES black box has remained impenetrable. Although several algorithms from Explainable Artificial Intelligence have recently been published, no research has yet investigated the role that these explanation models can play in: (a) discovering the decision-making process that drives AES, (b) fine-tuning predictive models to improve generalizability and interpretability, and (c) providing personalized, formative, and fine-grained feedback to students during the writing process. Building on previous studies where models were trained to predict both the holistic and rubric scores of essays, using the Automated Student Assessment Prize’s essay datasets, this study focuses on predicting the quality of the writing style of Grade-7 essays and exposes the decision processes that lead to these predictions. In doing so, it evaluates the impact of deep learning (multi-layer perceptron neural networks) on the performance of AES. It has been found that the effect of deep learning can be best viewed when assessing the trustworthiness of explanation models. As more hidden layers were added to the neural network, the descriptive accuracy increased by about 10%. This study shows that faster (up to three orders of magnitude) SHAP implementations are as accurate as the slower model-agnostic one. It leverages the state-of-the-art in natural language processing, applying feature selection on a pool of 1592 linguistic indices that measure aspects of text cohesion, lexical diversity, lexical sophistication, and syntactic sophistication and complexity. In addition to the list of most globally important features, this study reports (a) a list of features that are important for a specific essay (locally), (b) a range of values for each feature that contribute to higher or lower rubric scores, and (c) a model that allows to quantify the impact of the implementation of formative feedback.

Automated essay scoring (AES) is a compelling topic in Learning Analytics (LA) for the primary reason that recent advances in AI find it as a good testbed to explore artificial supplementation of human creativity. However, a vast swath of research tackles AES only holistically; only a few have even developed AES models at the rubric level, the very first layer of explanation underlying the prediction of holistic scores ( Kumar et al., 2017 ; Taghipour, 2017 ; Kumar and Boulanger, 2020 ). None has attempted to explain the whole decision process of AES, from holistic scores to rubric scores and from rubric scores to writing feature modeling. Although several algorithms from XAI (explainable artificial intelligence) ( Adadi and Berrada, 2018 ; Murdoch et al., 2019 ) have recently been published (e.g., LIME, SHAP) ( Ribeiro et al., 2016 ; Lundberg and Lee, 2017 ), no research has yet investigated the role that these explanation models (trained on top of predictive models) can play in: (a) discovering the decision-making process that drives AES, (b) fine-tuning predictive models to improve generalizability and interpretability, and (c) providing teachers and students with personalized, formative, and fine-grained feedback during the writing process.

One of the key anticipated benefits of AES is the elimination of human bias such as rater fatigue, rater’s expertise, severity/leniency, scale shrinkage, stereotyping, Halo effect, rater drift, perception difference, and inconsistency ( Taghipour, 2017 ). At its turn, AES may suffer from its own set of biases (e.g., imperfections in training data, spurious correlations, overrepresented minority groups), which has incited the research community to look for ways to make AES more transparent, accountable, fair, unbiased, and consequently trustworthy while remaining accurate. This required changing the perception that AES is merely a machine learning and feature engineering task ( Madnani et al., 2017 ; Madnani and Cahill, 2018 ). Hence, researchers have advocated that AES should be seen as a shared task requiring several methodological design decisions along the way such as curriculum alignment, construction of training corpora, reliable scoring process, and rater performance evaluation, where the goal is to build and deploy fair and unbiased scoring models to be used in large-scale assessments and classroom settings ( Rupp, 2018 ; West-Smith et al., 2018 ; Rupp et al., 2019 ). Unfortunately, although these measures are intended to design reliable and valid AES systems, they may still fail to build trust among users, keeping the AES black box impenetrable for teachers and students.

It has been previously recognized that divergence of opinion among human and machine graders has been only investigated superficially ( Reinertsen, 2018 ). So far, researchers investigated the characteristics of essays through qualitative analyses which ended up rejected by AES systems (requiring a human to score them) ( Reinertsen, 2018 ). Others strived to justify predicted scores by identifying essay segments that actually caused the predicted scores. In spite of the fact that these justifications hinted at and quantified the importance of these spatial cues, they did not provide any feedback as to how to improve those suboptimal essay segments ( Mizumoto et al., 2019 ).

Related to this study and the work of Kumar and Boulanger (2020) is Revision Assistant, a commercial AES system developed by Turnitin ( Woods et al., 2017 ; West-Smith et al., 2018 ), which in addition to predicting essays’ holistic scores provides formative, rubric-specific, and sentence-level feedback over multiple drafts of a student’s essay. The implementation of Revision Assistant moved away from the traditional approach to AES, which consists in using a limited set of features engineered by human experts representing only high-level characteristics of essays. Like this study, it rather opted for including a large number of low-level writing features, demonstrating that expert-designed features are not required to produce interpretable predictions. Revision Assistant’s performance was reported on two essay datasets, one of which was the Automated Student Assessment Prize (ASAP) 1 dataset. However, performance on the ASAP dataset was reported in terms of quadratic weighted kappa and this for holistic scores only. Models predicting rubric scores were trained only with the other dataset which was hosted on and collected through Revision Assistant itself.

In contrast to feature-based approaches like the one adopted by Revision Assistant, other AES systems are implemented using deep neural networks where features are learned during model training. For example, Taghipour (2017) in his doctoral dissertation leverages a recurrent neural network to improve accuracy in predicting holistic scores, implement rubric scoring (i.e., organization and argument strength), and distinguish between human-written and computer-generated essays. Interestingly, Taghipour compared the performance of his AES system against other AES systems using the ASAP corpora, but he did not use the ASAP corpora when it came to train rubric scoring models although ASAP provides two corpora provisioning rubric scores (#7 and #8). Finally, research was also undertaken to assess the generalizability of rubric-based models by performing experiments across various datasets. It was found that the predictive power of such rubric-based models was related to how much the underlying feature set covered a rubric’s criteria ( Rahimi et al., 2017 ).

Despite their numbers, rubrics (e.g., organization, prompt adherence, argument strength, essay length, conventions, word choices, readability, coherence, sentence fluency, style, audience, ideas) are usually investigated in isolation and not as a whole, with the exception of Revision Assistant which provides feedback at the same time on the following five rubrics: claim, development, audience, cohesion, and conventions. The literature reveals that rubric-specific automated feedback includes numerical rubric scores as well as recommendations on how to improve essay quality and correct errors ( Taghipour, 2017 ). Again, except for Revision Assistant which undertook a holistic approach to AES including holistic and rubric scoring and provision of rubric-specific feedback at the sentence level, AES has generally not been investigated as a whole or as an end-to-end product. Hence, the AES used in this study and developed by Kumar and Boulanger (2020) is unique in that it uses both deep learning (multi-layer perceptron neural network) and a huge pool of linguistic indices (1592), predicts both holistic and rubric scores, explaining holistic scores in terms of rubric scores, and reports which linguistic indices are the most important by rubric. This study, however, goes one step further and showcases how to explain the decision process behind the prediction of a rubric score for a specific essay, one of the main AES limitations identified in the literature ( Taghipour, 2017 ) that this research intends to address, at least partially.

Besides providing explanations of predictions both globally and individually, this study not only goes one step further toward the automated provision of formative feedback but also does so in alignment with the explanation model and the predictive model, allowing to better map feedback to the actual characteristics of an essay. Woods et al. (2017) succeeded in associating sentence-level expert-derived feedback with strong/weak sentences having the greatest influence on a rubric score based on the rubric, essay score, and the sentence characteristics. While Revision Assistant’s feature space consists of counts and binary occurrence indicators of word unigrams, bigrams and trigrams, character four-grams, and part-of-speech bigrams and trigrams, they are mainly textual and locational indices; by nature they are not descriptive or self-explanative. This research fills this gap by proposing feedback based on a set of linguistic indices that can encompass several sentences at a time. However, the proposed approach omits locational hints, leaving the merging of the two approaches as the next step to be addressed by the research community.

Although this paper proposes to extend the automated provision of formative feedback through an interpretable machine learning method, it rather focuses on the feasibility of automating it in the context of AES instead of evaluating the pedagogical quality (such as the informational and communicational value of feedback messages) or impact on students’ writing performance, a topic that will be kept for an upcoming study. Having an AES system that is capable of delivering real-time formative feedback sets the stage to investigate (1) when feedback is effective, (2) the types of feedback that are effective, and (3) whether there exist different kinds of behaviors in terms of seeking and using feedback ( Goldin et al., 2017 ). Finally, this paper omits describing the mapping between the AES model’s linguistic indices and a pedagogical language that is easily understandable by students and teachers, which is beyond its scope.

Methodology

This study showcases the application of the PDR framework ( Murdoch et al., 2019 ), which provides three pillars to describe interpretations in the context of the data science life cycle: P redictive accuracy, D escriptive accuracy, and R elevancy to human audience(s). It is important to note that in a broader sense both terms “explainable artificial intelligence” and “interpretable machine learning” can be used interchangeably with the following meaning ( Murdoch et al., 2019 ): “the use of machine-learning models for the extraction of relevant knowledge about domain relationships contained in data.” Here “predictive accuracy” refers to the measurement of a model’s ability to fit data; “descriptive accuracy” is the degree at which the relationships learned by a machine learning model can be objectively captured; and “relevant knowledge” implies that a particular audience gets insights into a chosen domain problem that guide its communication, actions, and discovery ( Murdoch et al., 2019 ).

In the context of this article, formative feedback that assesses students’ writing skills and prescribes remedial writing strategies is the relevant knowledge sought for, whose effectiveness on students’ writing performance will be validated in an upcoming study. However, the current study puts forward the tools and evaluates the feasibility to offer this real-time formative feedback. It also measures the predictive and descriptive accuracies of AES and explanation models, two key components to generate trustworthy interpretations ( Murdoch et al., 2019 ). Naturally, the provision of formative feedback is dependent on the speed of training and evaluating new explanation models every time a new essay is ingested by the AES system. That is why this paper investigates the potential of various SHAP implementations for speed optimization without compromising the predictive and descriptive accuracies. This article will show how the insights generated by the explanation model can serve to debug the predictive model and contribute to enhance the feature selection and/or engineering process ( Murdoch et al., 2019 ), laying the foundation for the provision of actionable and impactful pieces of knowledge to educational audiences, whose relevancy will be judged by the human stakeholders and estimated by the magnitude of resulting changes.

Figure 1 overviews all the elements and steps encompassed by the AES system in this study. The following subsections will address each facet of the overall methodology, from hyperparameter optimization to relevancy to both students and teachers.

www.frontiersin.org

Figure 1. A flow chart exhibiting the sequence of activities to develop an end-to-end AES system and how the various elements work together to produce relevant knowledge to the intended stakeholders.

Automated Essay Scoring System, Dataset, and Feature Selection

As previously mentioned, this paper reuses the AES system developed by Kumar and Boulanger (2020) . The AES models were trained using the ASAP’s seventh essay corpus. These narrative essays were written by Grade-7 students in the setting of state-wide assessments in the United States and had an average length of 171 words. Students were asked to write a story about patience. Kumar and Boulanger’s work consisted in training a predictive model for each of the four rubrics according to which essays were graded: ideas, organization, style, and conventions. Each essay was scored by two human raters on a 0−3 scale (integer scale). Rubric scores were resolved by adding the rubric scores assigned by the two human raters, producing a resolved rubric score between 0 and 6. This paper is a continuation of Boulanger and Kumar (2018 , 2019 , 2020) and Kumar and Boulanger (2020) where the objective is to open the AES black box to explain the holistic and rubric scores that it predicts. Essentially, the holistic score ( Boulanger and Kumar, 2018 , 2019 ) is determined and justified through its four rubrics. Rubric scores, in turn, are investigated to highlight the writing features that play an important role within each rubric ( Kumar and Boulanger, 2020 ). Finally, beyond global feature importance, it is not only indispensable to identify which writing indices are important for a particular essay (local), but also to discover how they contribute to increase or decrease the predicted rubric score, and which feature values are more/less desirable ( Boulanger and Kumar, 2020 ). This paper is a continuation of these previous works by adding the following link to the AES chain: holistic score, rubric scores, feature importance, explanations, and formative feedback. The objective is to highlight the means for transparent and trustable AES while empowering learning analytics practitioners with the tools to debug these models and equip educational stakeholders with an AI companion that will semi-autonomously generate formative feedback to teachers and students. Specifically, this paper analyzes the AES reasoning underlying its assessment of the “style” rubric, which looks for command of language, including effective and compelling word choice and varied sentence structure, that clearly supports the writer’s purpose and audience.

This research’s approach to AES leverages a feature-based multi-layer perceptron (MLP) deep neural network to predict rubric scores. The AES system is fed by 1592 linguistic indices quantitatively measured by the Suite of Automatic Linguistic Analysis Tools 2 (SALAT), which assess aspects of grammar and mechanics, sentiment analysis and cognition, text cohesion, lexical diversity, lexical sophistication, and syntactic sophistication and complexity ( Kumar and Boulanger, 2020 ). The purpose of using such a huge pool of low-level writing features is to let deep learning extract the most important ones; the literature supports this practice since there is evidence that features automatically selected are not less interpretable than those engineered ( Woods et al., 2017 ). However, to facilitate this process, this study opted for a semi-automatic strategy that consisted of both filter and embedded methods. Firstly, the original ASAP’s seventh essay dataset consists of a training set of 1567 essays and a validation and testing sets of 894 essays combined. While the texts of all 2461 essays are still available to the public, only the labels (the rubric scores of two human raters) of the training set have been shared with the public. Yet, this paper reused the unlabeled 894 essays of the validation and testing sets for feature selection, a process that must be carefully carried out by avoiding being informed by essays that will train the predictive model. Secondly, feature data were normalized, and features with variances lower than 0.01 were pruned. Thirdly, the last feature of any pair of features having an absolute Pearson correlation coefficient greater than 0.7 was also pruned (the one that comes last in terms of the column ordering in the datasets). After the application of these filter methods, the number of features was reduced from 1592 to 282. Finally, the Lasso and Ridge regression regularization methods (whose combination is also called ElasticNet) were applied during the training of the rubric scoring models. Lasso is responsible for pruning further features, while Ridge regression is entrusted with eliminating multicollinearity among features.

Hyperparameter Optimization and Training

To ensure a fair evaluation of the potential of deep learning, it is of utmost importance to minimally describe this study’s exploration of the hyperparameter space, a step that is often found to be missing when reporting the outcomes of AES models’ performance ( Kumar and Boulanger, 2020 ). First, a study should list the hyperparameters it is going to investigate by testing for various values of each hyperparameter. For example, Table 1 lists all hyperparameters explored in this study. Note that L 1 and L 2 are two regularization hyperparameters contributing to feature selection. Second, each study should also report the range of values of each hyperparameter. Finally, the strategy to explore the selected hyperparameter subspace should be clearly defined. For instance, given the availability of high-performance computing resources and the time/cost of training AES models, one might favor performing a grid (a systematic testing of all combinations of hyperparameters and hyperparameter values within a subspace) or a random search (randomly selecting a hyperparameter value from a range of values per hyperparameter) or both by first applying random search to identify a good starting candidate and then grid search to test all possible combinations in the vicinity of the starting candidate’s subspace. Of particular interest to this study is the neural network itself, that is, how many hidden layers should a neural network have and how many neurons should compose each hidden layer and the neural network as a whole. These two variables are directly related to the size of the neural network, with the number of hidden layers being a defining trait of deep learning. A vast swath of literature is silent about the application of interpretable machine learning in AES and even more about measuring its descriptive accuracy, the two components of trustworthiness. Hence, this study pioneers the comprehensive assessment of deep learning impact on AES’s predictive and descriptive accuracies.

www.frontiersin.org

Table 1. Hyperparameter subspace investigated in this article along with best hyperparameter values per neural network architecture.

Consequently, the 1567 labeled essays were divided into a training set (80%) and a testing set (20%). No validation set was put aside; 5-fold cross-validation was rather used for hyperparameter optimization. Table 1 delineates the hyperparameter subspace from which 800 different combinations of hyperparameter values were randomly selected out of a subspace of 86,248,800 possible combinations. Since this research proposes to investigate the potential of deep learning to predict rubric scores, several architectures consisting of 2 to 6 hidden layers and ranging from 9,156 to 119,312 parameters were tested. Table 1 shows the best hyperparameter values per depth of neural networks.

Again, the essays of the testing set were never used during the training and cross-validation processes. In order to retrieve the best predictive models during training, every time the validation loss reached a record low, the model was overwritten. Training stopped when no new record low was reached during 100 epochs. Moreover, to avoid reporting the performance of overfit models, each model was trained five times using the same set of best hyperparameter values. Finally, for each resulting predictive model, a corresponding ensemble model (bagging) was also obtained out of the five models trained during cross-validation.

Predictive Models and Predictive Accuracy

Table 2 delineates the performance of predictive models trained previously by Kumar and Boulanger (2020) on the four scoring rubrics. The first row lists the agreement levels between the resolved and predicted rubric scores measured by the quadratic weighted kappa. The second row is the percentage of accurate predictions; the third row reports the percentages of predictions that are either accurate or off by 1; and the fourth row reports the percentages of predictions that are either accurate or at most off by 2. Prediction of holistic scores is done merely by adding up all rubric scores. Since the scale of rubric scores is 0−6 for every rubric, then the scale of holistic scores is 0−24.

www.frontiersin.org

Table 2. Rubric scoring models’ performance on testing set.

While each of these rubric scoring models might suffer from its own systemic bias and hence cancel off each other’s bias by adding up the rubric scores to derive the holistic score, this study (unlike related works) intends to highlight these biases by exposing the decision making process underlying the prediction of rubric scores. Although this paper exclusively focuses on the Style rubric, the methodology put forward to analyze the local and global importance of writing indices and their context-specific contributions to predicted rubric scores is applicable to every rubric and allows to control for these biases one rubric at a time. Comparing and contrasting the role that a specific writing index plays within each rubric context deserves its own investigation, which has been partly addressed in the study led by Kumar and Boulanger (2020) . Moreover, this paper underscores the necessity to measure the predictive accuracy of rubric-based holistic scoring using additional metrics to account for these rubric-specific biases. For example, there exist several combinations of rubric scores to obtain a holistic score of 16 (e.g., 4-4-4-4 vs. 4-3-4-5 vs. 3-5-2-6). Even though the predicted holistic score might be accurate, the rubric scores could all be inaccurate. Similarity or distance metrics (e.g., Manhattan and Euclidean) should then be used to describe the authenticity of the composition of these holistic scores.

According to what Kumar and Boulanger (2020) report on the performance of several state-of-the-art AES systems trained on ASAP’s seventh essay dataset, the AES system they developed and which will be reused in this paper proved competitive while being fully and deeply interpretable, which no other AES system does. They also supply further information about the study setting, essay datasets, rubrics, features, natural language processing (NLP) tools, model training, and evaluation against human performance. Again, this paper showcases the application of explainable artificial intelligence in automated essay scoring by focusing on the decision process of the Rubric #3 (Style) scoring model. Remember that the same methodology is applicable to each rubric.

Explanation Model: SHAP

SH apley A dditive ex P lanations (SHAP) is a theoretically justified XAI framework that can provide simultaneously both local and global explanations ( Molnar, 2020 ); that is, SHAP is able to explain individual predictions taking into account the uniqueness of each prediction, while highlighting the global factors influencing the overall performance of a predictive model. SHAP is of keen interest because it unifies all algorithms of the class of additive feature attribution methods, adhering to a set of three properties that are desirable in interpretable machine learning: local accuracy, missingness, and consistency ( Lundberg and Lee, 2017 ). A key advantage of SHAP is that feature contributions are all expressed in terms of the outcome variable (e.g., rubric scores), providing a same scale to compare the importance of each feature against each other. Local accuracy refers to the fact that no matter the explanation model, the sum of all feature contributions is always equal to the prediction explained by these features. The missingness property implies that the prediction is never explained by unmeasured factors, which are always assigned a contribution of zero. However, the converse is not true; a contribution of zero does not imply an unobserved factor, it can also denote a feature irrelevant to explain the prediction. The consistency property guarantees that a more important feature will always have a greater magnitude than a less important one, no matter how many other features are included in the explanation model. SHAP proves superior to other additive attribution methods such as LIME (Local Interpretable Model-Agnostic Explanations), Shapley values, and DeepLIFT in that they never comply with all three properties, while SHAP does ( Lundberg and Lee, 2017 ). Moreover, the way SHAP assesses the importance of a feature differs from permutation importance methods (e.g., ELI5), measured as the decrease in model performance (accuracy) as a feature is perturbated, in that it is based on how much a feature contributes to every prediction.

Essentially, a SHAP explanation model (linear regression) is trained on top of a predictive model, which in this case is a complex ensemble deep learning model. Table 3 demonstrates a scale explanation model showing how SHAP values (feature contributions) work. In this example, there are five instances and five features describing each instance (in the context of this paper, an instance is an essay). Predictions are listed in the second to last column, and the base value is the mean of all predictions. The base value constitutes the reference point according to which predictions are explained; in other words, reasons are given to justify the discrepancy between the individual prediction and the mean prediction (the base value). Notice that the table does not contain the actual feature values; these are SHAP values that quantify the contribution of each feature to the predicted score. For example, the prediction of Instance 1 is 2.46, while the base value is 3.76. Adding up the feature contributions of Instance 1 to the base value produces the predicted score:

www.frontiersin.org

Table 3. Array of SHAP values: local and global importance of features and feature coverage per instance.

Hence, the generic equation of the explanation model ( Lundberg and Lee, 2017 ) is:

where g(x) is the prediction of an individual instance x, σ 0 is the base value, σ i is the feature contribution of feature x i , x i ∈ {0,1} denotes whether feature x i is part of the individual explanation, and j is the total number of features. Furthermore, the global importance of a feature is calculated by adding up the absolute values of its corresponding SHAP values over all instances, where n is the total number of instances and σ i ( j ) is the feature contribution for instance i ( Lundberg et al., 2018 ):

Therefore, it can be seen that Feature 3 is the most globally important feature, while Feature 2 is the least important one. Similarly, Feature 5 is Instance 3’s most important feature at the local level, while Feature 2 is the least locally important. The reader should also note that a feature shall not necessarily be assigned any contribution; some of them are just not part of the explanation such as Feature 2 and Feature 3 in Instance 2. These concepts lay the foundation for the explainable AES system presented in this paper. Just imagine that each instance (essay) will be rather summarized by 282 features and that the explanations of all the testing set’s 314 essays will be provided.

Several implementations of SHAP exist: KernelSHAP, DeepSHAP, GradientSHAP, and TreeSHAP, among others. KernelSHAP is model-agnostic and works for any type of predictive models; however, KernelSHAP is very computing-intensive which makes it undesirable for practical purposes. DeepSHAP and GradientSHAP are two implementations intended for deep learning which takes advantage of the known properties of neural networks (i.e., MLP-NN, CNN, or RNN) to accelerate up to three orders of magnitude the processing time to explain predictions ( Chen et al., 2019 ). Finally, TreeSHAP is the most powerful implementation intended for tree-based models. TreeSHAP is not only fast; it is also accurate. While the three former implementations estimate SHAP values, TreeSHAP computes them exactly. Moreover, TreeSHAP not only measures the contribution of individual features, but it also considers interactions between pairs of features and assigns them SHAP values. Since one of the goals of this paper is to assess the potential of deep learning on the performance of both predictive and explanation models, this research tested the former three implementations. TreeSHAP is recommended for future work since the interaction among features is critical information to consider. Moreover, KernelSHAP, DeepSHAP, and GradientSHAP all require access to the whole original dataset to derive the explanation of a new instance, another constraint TreeSHAP is not subject to.

Descriptive Accuracy: Trustworthiness of Explanation Models

This paper reuses and adapts the methodology introduced by Ribeiro et al. (2016) . Several explanation models will be trained, using different SHAP implementations and configurations, per deep learning predictive model (for each number of hidden layers). The rationale consists in randomly selecting and ignoring 25% of the 282 features feeding the predictive model (e.g., turning them to zero). If it causes the prediction to change beyond a specific threshold (in this study 0.10 and 0.25 were tested), then the explanation model should also reflect the magnitude of this change while ignoring the contributions of these same features. For example, the original predicted rubric score of an essay might be 5; however, when ignoring the information brought in by a subset of 70 randomly selected features (25% of 282), the prediction may turn to 4. On the other side, if the explanation model also predicts a 4 while ignoring the contributions of the same subset of features, then the explanation is considered as trustworthy. This allows to compute the precision, recall, and F1-score of each explanation model (number of true and false positives and true and false negatives). The process is repeated 500 times for every essay to determine the average precision and recall of every explanation model.

Judging Relevancy

So far, the consistency of explanations with predictions has been considered. However, consistent explanations do not imply relevant or meaningful explanations. Put another way, explanations only reflect what predictive models have learned during training. How can the black box of these explanations be opened? Looking directly at the numerical SHAP values of each explanation might seem a daunting task, but there exist tools, mainly visualizations (decision plot, summary plot, and dependence plot), that allow to make sense out of these explanations. However, before visualizing these explanations, another question needs to be addressed: which explanations or essays should be picked for further scrutiny of the AES system? Given the huge number of essays to examine and the tedious task to understand the underpinnings of a single explanation, a small subset of essays should be carefully picked that should represent concisely the state of correctness of the underlying predictive model. Again, this study applies and adapts the methodology in Ribeiro et al. (2016) . A greedy algorithm selects essays whose predictions are explained by as many features of global importance as possible to optimize feature coverage. Ribeiro et al. demonstrated in unrelated studies (i.e., sentiment analysis) that the correctness of a predictive model can be assessed with as few as four or five well-picked explanations.

For example, Table 3 reveals the global importance of five features. The square root of each feature’s global importance is also computed and considered instead to limit the influence of a small group of very influential features. The feature coverage of Instance 1 is 100% because all features are engaged in the explanation of the prediction. On the other hand, Instance 2 has a feature coverage of 61.5% because only Features 1, 4, and 5 are part of the prediction’s explanation. The feature coverage is calculated by summing the square root of each explanation’s feature’s global importance together and dividing by the sum of the square roots of all features’ global importance:

Additionally, it can be seen that Instance 4 does not have any zero-feature value although its feature coverage is only 84.6%. The algorithm was constrained to discard from the explanation any feature whose contribution (local importance) was too close to zero. In the case of Table 3 ’s example, any feature whose absolute SHAP value is less than 0.10 is ignored, hence leading to a feature coverage of:

In this paper’s study, the real threshold was 0.01. This constraint was actually a requirement for the DeepSHAP and GradientSHAP implementations because they only output non-zero SHAP values contrary to KernelSHAP which generates explanations with a fixed number of features: a non-zero SHAP value indicates that the feature is part of the explanation, while a zero value excludes the feature from the explanation. Without this parameter, all 282 features would be part of the explanation although a huge number only has a trivial (very close to zero) SHAP value. Now, a much smaller but variable subset of features makes up each explanation. This is one way in which Ribeiro et al.’s SP-LIME algorithm (SP stands for Submodular Pick) has been adapted to this study’s needs. In conclusion, notice how Instance 4 would be selected in preference to Instance 5 to explain Table 3 ’s underlying predictive model. Even though both instances have four features explaining their prediction, Instance 4’s features are more globally important than Instance 5’s features, and therefore Instance 4 has greater feature coverage than Instance 5.

Whereas Table 3 ’s example exhibits the feature coverage of one instance at a time, this study computes it for a subset of instances, where the absolute SHAP values are aggregated (summed) per candidate subset. When the sum of absolute SHAP values per feature exceeds the set threshold, the feature is then considered as covered by the selected set of instances. The objective in this study was to optimize the feature coverage while minimizing the number of essays to validate the AES model.

Research Questions

One of this article’s objectives is to assess the potential of deep learning in automated essay scoring. The literature has often claimed ( Hussein et al., 2019 ) that there are two approaches to AES, feature-based and deep learning, as though these two approaches were mutually exclusive. Yet, the literature also puts forward that feature-based AES models may be more interpretable than deep learning ones ( Amorim et al., 2018 ). This paper embraces the viewpoint that these two approaches can also be complementary by leveraging the state-of-the-art in NLP and automatic linguistic analysis and harnessing one of the richest pools of linguistic indices put forward in the research community ( Crossley et al., 2016 , 2017 , 2019 ; Kyle, 2016 ; Kyle et al., 2018 ) and applying a thorough feature selection process powered by deep learning. Moreover, the ability of deep learning of modeling complex non-linear relationships makes it particularly well-suited for AES given that the importance of a writing feature is highly dependent on its context, that is, its interactions with other writing features. Besides, this study leverages the SHAP interpretation method that is well-suited to interpret very complex models. Hence, this study elected to work with deep learning models and ensembles to test SHAP’s ability to explain these complex models. Previously, the literature has revealed the difficulty to have at the same time both accurate and interpretable models ( Ribeiro et al., 2016 ; Murdoch et al., 2019 ), where favoring one comes at the expense of the other. However, this research shows how XAI makes it now possible to produce both accurate and interpretable models in the area of AES. Since ensembles have been repeatedly shown to boost the accuracy of predictive models, they were included as part of the tested deep learning architectures to maximize generalizability and accuracy, while making these predictive models interpretable and exploring whether deep learning can even enhance their descriptive accuracy further.

This study investigates the trustworthiness of explanation models, and more specifically, those explaining deep learning predictive models. For instance, does the depth, defined as the number of hidden layers, of an MLP neural network increases the trustworthiness of its SHAP explanation model? The answer to this question will help determine whether it is possible to have very accurate AES models while having competitively interpretable/explainable models, the corner stone for the generation of formative feedback. Remember that formative feedback is defined as “any kind of information provided to students about their actual state of learning or performance in order to modify the learner’s thinking or behavior in the direction of the learning standards” and that formative feedback “conveys where the student is, what are the goals to reach, and how to reach the goals” ( Goldin et al., 2017 ). This notion contrasts with summative feedback which basically is “a justification of the assessment results” ( Hao and Tsikerdekis, 2019 ).

As pointed out in the previous section, multiple SHAP implementations are evaluated in this study. Hence, this paper showcases whether the faster DeepSHAP and GradientSHAP implementations are as reliable as the slower KernelSHAP implementation . The answer to this research question will shed light on the feasibility of providing immediate formative feedback and this multiple times throughout students’ writing processes.

This study also looks at whether a summary of the data produces as trustworthy explanations as those from the original data . This question will be of interest to AES researchers and practitioners because it could allow to significantly decrease the processing time of the computing-intensive and model-agnostic KernelSHAP implementation and test further the potential of customizable explanations.

KernelSHAP allows to specify the total number of features that will shape the explanation of a prediction; for instance, this study experiments with explanations of 16 and 32 features and observes whether there exists a statistically significant difference in the reliability of these explanation models . Knowing this will hint at whether simpler or more complex explanations are more desirable when it comes to optimize their trustworthiness. If there is no statistically significant difference, then AES practitioners are given further flexibility in the selection of SHAP implementations to find the sweet spot between complexity of explanations and speed of processing. For instance, the KernelSHAP implementation allows to customize the number of factors making up an explanation, while the faster DeepSHAP and GradientSHAP do not.

Finally, this paper highlights the means to debug and compare the performance of predictive models through their explanations. Once a model is debugged, the process can be reused to fine-tune feature selection and/or feature engineering to improve predictive models and for the generation of formative feedback to both students and teachers.

The training, validation, and testing sets consist of 1567 essays, each of which has been scored by two human raters, who assigned a score between 0 and 3 per rubric (ideas, organization, style, and conventions). In particular, this article looks at predictive and descriptive accuracy of AES models on the third rubric, style. Note that although each essay has been scored by two human raters, the literature ( Shermis, 2014 ) is not explicit about whether only two or more human raters participated in the scoring of all 1567 essays; given the huge number of essays, it is likely that more than two human raters were involved in the scoring of these essays so that the amount of noise introduced by the various raters’ biases is unknown while probably being at some degree balanced among the two groups of raters. Figure 2 shows the confusion matrices of human raters on Style Rubric. The diagonal elements (dark gray) correspond to exact matches, whereas the light gray squares indicate adjacent matches. Figure 2A delineates the number of essays per pair of ratings, and Figure 2B shows the percentages per pair of ratings. The agreement level between each pair of human raters, measured by the quadratic weighted kappa, is 0.54; the percentage of exact matches is 65.3%; the percentage of adjacent matches is 34.4%; and 0.3% of essays are neither exact nor adjacent matches. Figures 2A,B specify the distributions of 0−3 ratings per group of human raters. Figure 2C exhibits the distribution of resolved scores (a resolved score is the sum of the two human ratings). The mean is 3.99 (with a standard deviation of 1.10), and the median and mode are 4. It is important to note that the levels of predictive accuracy reported in this article are measured on the scale of resolved scores (0−6) and that larger scales tend to slightly inflate quadratic weighted kappa values, which must be taken into account when comparing against the level of agreement between human raters. Comparison of percentages of exact and adjacent matches must also be made with this scoring scale discrepancy in mind.

www.frontiersin.org

Figure 2. Summary of the essay dataset (1567 Grade-7 narrative essays) investigated in this study. (A) Number of essays per pair of human ratings; the diagonal (dark gray squares) lists the numbers of exact matches while the light-gray squares list the numbers of adjacent matches; and the bottom row and the rightmost column highlight the distributions of ratings for both groups of human raters. (B) Percentages of essays per pair of human ratings; the diagonal (dark gray squares) lists the percentages of exact matches while the light-gray squares list the percentages of adjacent matches; and the bottom row and the rightmost column highlight the distributions (frequencies) of ratings for both groups of human raters. (C) The distribution of resolved rubric scores; a resolved score is the addition of its two constituent human ratings.

Predictive Accuracy and Descriptive Accuracy

Table 4 compiles the performance outcomes of the 10 predictive models evaluated in this study. The reader should remember that the performance of each model was averaged over five iterations and that two models were trained per number of hidden layers, one non-ensemble and one ensemble. Except for the 6-layer models, there is no clear winner among other models. Even for the 6-layer models, they are superior in terms of exact matches, the primary goal for a reliable AES system, but not according to adjacent matches. Nevertheless, on average ensemble models slightly outperform non-ensemble models. Hence, these ensemble models will be retained for the next analysis step. Moreover, given that five ensemble models were trained per neural network depth, the most accurate model among the five is selected and displayed in Table 4 .

www.frontiersin.org

Table 4. Performance of majority classifier and average/maximal performance of trained predictive models.

Next, for each selected ensemble predictive model, several explanation models are trained per predictive model. Every predictive model is explained by the “Deep,” “Grad,” and “Random” explainers, except for the 6-layer model where it was not possible to train a “Deep” explainer apparently due to a bug in the original SHAP code caused by either a unique condition in this study’s data or neural network architecture. However, this was beyond the scope of this study to fix and investigate this issue. As it will be demonstrated, no statistically significant difference exists between the accuracy of these explainers.

The “Random” explainer serves as a baseline model for comparison purpose. Remember that to evaluate the reliability of explanation models, the concurrent impact of randomly selecting and ignoring a subset of features on the prediction and explanation of rubric scores is analyzed. If the prediction changes significantly and its corresponding explanation changes (beyond a set threshold) accordingly (a true positive) or if the prediction remains within the threshold as does the explanation (a true negative), then the explanation is deemed as trustworthy. Hence, in the case of the Random explainer, it simulates random explanations by randomly selecting 32 non-zero features from the original set of 282 features. These random explanations consist only of non-zero features because, according to SHAP’s missingness property, a feature with a zero or a missing value never gets assigned any contribution to the prediction. If at least one of these 32 features is also an element of the subset of the ignored features, then the explanation is considered as untrustworthy, no matter the size of a feature’s contribution.

As for the layer-2 model, six different explanation models are evaluated. Recall that layer-2 models generated the least mean squared error (MSE) during hyperparameter optimization (see Table 1 ). Hence, this specific type of architecture was selected to test the reliability of these various explainers. The “Kernel” explainer is the most computing-intensive and took approximately 8 h of processing. It was trained using the full distributions of feature values in the training set and shaped explanations in terms of 32 features; the “Kernel-16” and “Kernel-32” models were trained on a summary (50 k -means centroids) of the training set to accelerate the processing by about one order of magnitude (less than 1 h). Besides, the “Kernel-16” explainer derived explanations in terms of 16 features, while the “Kernel-32” explainer explained predictions through 32 features. Table 5 exhibits the descriptive accuracy of these various explanation models according to a 0.10 and 0.25 threshold; in other words, by ignoring a subset of randomly picked features, it assesses whether or not the prediction and explanation change simultaneously. Note also how each explanation model, no matter the underlying predictive model, outperforms the “Random” model.

www.frontiersin.org

Table 5. Precision, recall, and F1 scores of the various explainers tested per type of predictive model.

The first research question addressed in this subsection asks whether there exists a statistically significant difference between the “Kernel” explainer, which generates 32-feature explanations and is trained on the whole training set, and the “Kernel-32” explainer which also generates 32-feature explanations and is trained on a summary of the training set. To determine this, an independent t-test was conducted using the precision, recall, and F1-score distributions (500 iterations) of both explainers. Table 6 reports the p -values of all the tests and for the 0.10 and 0.25 thresholds. It reveals that there is no statistically significant difference between the two explainers.

www.frontiersin.org

Table 6. p -values of independent t -tests comparing whether there exist statistically significant differences between the mean precisions, recalls, and F1-scores of 2-layer explainers and between those of the 2-layer’s, 4-layer’s, and 6-layer’s Gradient explainers.

The next research question tests whether there exists a difference in the trustworthiness of explainers shaping 16 or 32-feature explanations. Again t-tests were conducted to verify this. Table 6 lists the resulting p -values. Again, there is no statistically significant difference in the average precisions, recalls, and F1-scores of both explainers.

This leads to investigating whether the “Kernel,” “Deep,” and “Grad” explainers are equivalent. Table 6 exhibits the results of the t-tests conducted to verify this and reveals that none of the explainers produce a statistically significantly better performance than the other.

Armed with this evidence, it is now possible to verify whether deeper MLP neural networks produce more trustworthy explanation models. For this purpose, the performance of the “Grad” explainer for each type of predictive model will be compared against each other. The same methodology as previously applied is employed here. Table 6 , again, confirms that the explanation model of the 2-layer predictive model is statistically significantly less trustworthy than the 4-layer’s explanation model; the same can be said of the 4-layer and 6-layer models. The only exception is the difference in average precision between 2-layer and 4-layer models and between 4-layer and 6-layer models; however, there clearly exists a statistically significant difference in terms of precision (and also recall and F1-score) between 2-layer and 6-layer models.

The Best Subset of Essays to Judge AES Relevancy

Table 7 lists the four best essays optimizing feature coverage (93.9%) along with their resolved and predicted scores. Notice how two of the four essays were picked by the adapted SP-LIME algorithm with some strong disagreement between the human and the machine graders, two were picked with short and trivial text, and two were picked exhibiting perfect agreement between the human and machine graders. Interestingly, each pair of longer and shorter essays exposes both strong agreement and strong disagreement between the human and AI agents, offering an opportunity to debug the model and evaluate its ability to detect the presence or absence of more basic (e.g., very small number of words, occurrences of sentence fragments) and more advanced aspects (e.g., cohesion between adjacent sentences, variety of sentence structures) of narrative essay writing and to appropriately reward or penalize them.

www.frontiersin.org

Table 7. Set of best essays to evaluate the correctness of the 6-layer ensemble AES model.

Local Explanation: The Decision Plot

The decision plot lists writing features by order of importance from top to bottom. The line segments display the contribution (SHAP value) of each feature to the predicted rubric score. Note that an actual decision plot consists of all 282 features and that only the top portion of it (20 most important features) can be displayed (see Figure 3 ). A decision plot is read from bottom to top. The line starts at the base value and ends at the predicted rubric score. Given that the “Grad” explainer is the only explainer common to all predictive models, it has been selected to derive all explanations. The decision plots in Figure 3 show the explanations of the four essays in Table 7 ; the dashed line in these plots represents the explanation of the most accurate predictive model, that is the ensemble model with 6 hidden layers which also produced the most trustworthy explanation model. The predicted rubric score of each explanation model is listed in the bottom-right legend. Explanation of the writing features follow in a next subsection.

www.frontiersin.org

Figure 3. Comparisons of all models’ explanations of the most representative set of four essays: (A) Essay 228, (B) Essay 68, (C) Essay 219, and (D) Essay 124.

Global Explanation: The Summary Plot

It is advantageous to use SHAP to build explanation models because it provides a single framework to discover the writing features that are important to an individual essay (local) or a set of essays (global). While the decision plots list features of local importance, Figure 4 ’s summary plot ranks writing features by order of global importance (from top to bottom). All testing set’s 314 essays are represented as dots in the scatterplot of each writing feature. The position of a dot on the horizontal axis corresponds to the importance (SHAP value) of the writing feature for a specific essay and its color indicates the magnitude of the feature value in relation to the range of all 314 feature values. For example, large or small numbers of words within an essay generally contribute to increase or decrease rubric scores by up to 1.5 and 1.0, respectively. Decision plots can also be used to find the most important features for a small subset of essays; Figure 5 demonstrates the new ordering of writing indices when aggregating the feature contributions (summing the absolute values of SHAP values) of the four essays in Table 7 . Moreover, Figure 5 allows to compare the contributions of a feature to various essays. Note how the orderings in Figures 3 −5 can differ from each other, sharing many features of global importance as well as having their own unique features of local importance.

www.frontiersin.org

Figure 4. Summary plot listing the 32 most important features globally.

www.frontiersin.org

Figure 5. Decision plot delineating the best model’s explanations of Essays 228, 68, 219, and 124 (6-layer ensemble).

Definition of Important Writing Indices

The reader shall understand that it is beyond the scope of this paper to make a thorough description of all writing features. Nevertheless, the summary and decision plots in Figures 4 , 5 allow to identify a subset of features that should be examined in order to validate this study’s predictive model. Supplementary Table 1 combines and describes the 38 features in Figures 4 , 5 .

Dependence Plots

Although the summary plot in Figure 4 is insightful to determine whether small or large feature values are desirable, the dependence plots in Figure 6 prove essential to recommend whether a student should aim at increasing or decreasing the value of a specific writing feature. The dependence plots also reveal whether the student should directly act upon the targeted writing feature or indirectly on other features. The horizontal axis in each of the dependence plots in Figure 6 is the scale of the writing feature and the vertical axis is the scale of the writing feature’s contributions to the predicted rubric scores. Each dot in a dependence plot represents one of the testing set’s 314 essays, that is, the feature value and SHAP value belonging to the essay. The vertical dispersion of the dots on small intervals of the horizontal axis is indicative of interaction with other features ( Molnar, 2020 ). If the vertical dispersion is widespread (e.g., the [50, 100] horizontal-axis interval in the “word_count” dependence plot), then the contribution of the writing feature is most likely at some degree dependent on other writing feature(s).

www.frontiersin.org

Figure 6. Dependence plots: the horizontal axes represent feature values while vertical axes represent feature contributions (SHAP values). Each dot represents one of the 314 essays of the testing set and is colored according to the value of the feature with which it interacts most strongly. (A) word_count. (B) hdd42_aw. (C) ncomp_stdev. (D) dobj_per_cl. (E) grammar. (F) SENTENCE_FRAGMENT. (G) Sv_GI. (H) adjacent_overlap_verb_sent.

The contributions of this paper can be summarized as follows: (1) it proposes a means (SHAP) to explain individual predictions of AES systems and provides flexible guidelines to build powerful predictive models using more complex algorithms such as ensembles and deep learning neural networks; (2) it applies a methodology to quantitatively assess the trustworthiness of explanation models; (3) it tests whether faster SHAP implementations impact the descriptive accuracy of explanation models, giving insight on the applicability of SHAP in real pedagogical contexts such as AES; (4) it offers a toolkit to debug AES models, highlights linguistic intricacies, and underscores the means to offer formative feedback to novice writers; and more importantly, (5) it empowers learning analytics practitioners to make AI pedagogical agents accountable to the human educator, the ultimate problem holder responsible for the decisions and actions of AI ( Abbass, 2019 ). Basically, learning analytics (which encompasses tools such as AES) is characterized as an ethics-bound, semi-autonomous, and trust-enabled human-AI fusion that recurrently measures and proactively advances knowledge boundaries in human learning.

To exemplify this, imagine an AES system that supports instructors in the detection of plagiarism, gaming behaviors, and the marking of writing activities. As previously mentioned, essays are marked according to a grid of scoring rubrics: ideas, organization, style, and conventions. While an abundance of data (e.g., the 1592 writing metrics) can be collected by the AES tool, these data might still be insufficient to automate the scoring process of certain rubrics (e.g., ideas). Nevertheless, some scoring subtasks such as assessing a student’s vocabulary, sentence fluency, and conventions might still be assigned to AI since the data types available through existing automatic linguistic analysis tools prove sufficient to reliably alleviate the human marker’s workload. Interestingly, learning analytics is key for the accountability of AI agents to the human problem holder. As the volume of writing data (through a large student population, high-frequency capture of learning episodes, and variety of big learning data) accumulate in the system, new AI agents (predictive models) may apply for the job of “automarker.” These AI agents can be quite transparent through XAI ( Arrieta et al., 2020 ) explanation models, and a human instructor may assess the suitability of an agent for the job and hire the candidate agent that comes closest to human performance. Explanations derived from these models could serve as formative feedback to the students.

The AI marker can be assigned to assess the writing activities that are similar to those previously scored by the human marker(s) from whom it learns. Dissimilar and unseen essays can be automatically assigned to the human marker for reliable scoring, and the AI agent can learn from this manual scoring. To ensure accountability, students should be allowed to appeal the AI agent’s marking to the human marker. In addition, the human marker should be empowered to monitor and validate the scoring of select writing rubrics scored by the AI marker. If the human marker does not agree with the machine scores, the writing assignments may be flagged as incorrectly scored and re-assigned to a human marker. These flagged assignments may serve to update predictive models. Moreover, among the essays that are assigned to the machine marker, a small subset can be simultaneously assigned to the human marker for continuous quality control; that is, to continue comparing whether the agreement level between human and machine markers remains within an acceptable threshold. The human marker should be at any time able to “fire” an AI marker or “hire” an AI marker from a pool of potential machine markers.

This notion of a human-AI fusion has been observed in previous AES systems where the human marker’s workload has been found to be significantly alleviated, passing from scoring several hundreds of essays to just a few dozen ( Dronen et al., 2015 ; Hellman et al., 2019 ). As the AES technology matures and as the learning analytics tools continue to penetrate the education market, this alliance of semi-autonomous human and AI agents will lead to better evidence-based/informed pedagogy ( Nelson and Campbell, 2017 ). Such a human-AI alliance can also be guided to autonomously self-regulate its own hypothesis-authoring and data-acquisition processes for purposes of measuring and advancing knowledge boundaries in human learning.

Real-Time Formative Pedagogical Feedback

This paper provides the evidence that deep learning and SHAP can be used not only to score essays automatically but also to offer explanations in real-time. More specifically, the processing time to derive the 314 explanations of the testing set’s essays has been benchmarked for several types of explainers. It was found that the faster DeepSHAP and GradientSHAP implementations, which took only a few seconds of processing, did not produce less accurate explanations than the much slower KernelSHAP. KernelSHAP took approximately 8 h of processing to derive the explanation model of a 2-layer MLP neural network predictive model and 16 h for the 6-layer predictive model.

This finding also holds for various configurations of KernelSHAP, where the number of features (16 vs. 32) shaping the explanation (where all other features are assigned zero contributions) did not produce a statistically significant difference in the reliability of the explanation models. On average, the models had a precision between 63.9 and 64.1% and a recall between 41.0 and 42.9%. This means that after perturbation of the predictive and explanation models, on average 64% of the predictions the explanation model identified as changing were accurate. On the other side, only about 42% of all predictions that changed were detected by the various 2-layer explainers. An explanation was considered as untrustworthy if the sum of its feature contributions, when added to the average prediction (base value), was not within 0.1 from the perturbated prediction. Similarly, the average precision and recall of 2-layer explainers for the 0.25-threshold were about 69% and 62%, respectively.

Impact of Deep Learning on Descriptive Accuracy of Explanations

By analyzing the performance of the various predictive models in Table 4 , no clear conclusion can be reached as to which model should be deemed as the most desirable. Despite the fact that the 6-layer models slightly outperform the other models in terms of accuracy (percentage of exact matches between the resolved [human] and predicted [machine] scores), they are not the best when it comes to the percentages of adjacent (within 1 and 2) matches. Nevertheless, if the selection of the “best” model is based on the quadratic weighted kappas, the decision remains a nebulous one to make. Moreover, ensuring that machine learning actually learned something meaningful remains paramount, especially in contexts where the performance of a majority classifier is close to the human and machine performance. For example, a majority classifier model would get 46.3% of predictions accurate ( Table 4 ), while trained predictive models at best produce accurate predictions between 51.9 and 55.1%.

Since the interpretability of a machine learning model should be prioritized over accuracy ( Ribeiro et al., 2016 ; Murdoch et al., 2019 ) for questions of transparency and trust, this paper investigated whether the impact of the depth of a MLP neural network might be more visible when assessing its interpretability, that is, the trustworthiness of its corresponding SHAP explanation model. The data in Tables 1 , 5 , 6 effectively support the hypothesis that as the depth of the neural network increases, the precision and recall of the corresponding explanation model improve. Besides, this observation is particularly interesting because the 4-layer (Grad) explainer, which has hardly more parameters than the 2-layer model, is also more accurate than the 2-layer model, suggesting that the 6-layer explainer is most likely superior to other explainers not only because of its greater number of parameters, but also because of its number of hidden layers. By increasing the number of hidden layers, it can be seen that the precision and recall of an explanation model can pass on average from approximately 64 to 73% and from 42 to 52%, respectively, for the 0.10-threshold; and for the 0.25-threshold, from 69 to 79% and from 62 to 75%, respectively.

These results imply that the descriptive accuracy of an explanation model is an evidence of effective machine learning, which may exceed the level of agreement between the human and machine graders. Moreover, given that the superiority of a trained predictive model over a majority classifier is not always obvious, the consistency of its associated explanation model demonstrates this better. Note that theoretically the SHAP explanation model of the majority classifier should assign a zero contribution to each writing feature since the average prediction of such a model is actually the most frequent rubric score given by the human raters; hence, the base value is the explanation.

An interesting fact emerges from Figure 3 , that is, all explainers (2-layer to 6-layer) are more or less similar. It appears that they do not contradict each other. More specifically, they all agree on the direction of the contributions of the most important features. In other words, they unanimously determine that a feature should increase or decrease the predicted score. However, they differ from each other on the magnitude of the feature contributions.

To conclude, this study highlights the need to train predictive models that consider the descriptive accuracy of explanations. The idea is that explanation models consider predictions to derive explanations; explanations should be considered when training predictive models. This would not only help train interpretable models the very first time but also potentially break the status quo that may exist among similar explainers to possibly produce more powerful models. In addition, this research calls for a mechanism (e.g., causal diagrams) to allow teachers to guide the training process of predictive models. Put another way, as LA practitioners debug predictive models, their insights should be encoded in a language that will be understood by the machine and that will guide the training process to avoid learning the same errors and to accelerate the training time.

Accountable AES

Now that the superiority of the 6-layer predictive and explanation models has been demonstrated, some aspects of the relevancy of explanations should be examined more deeply, knowing that having an explanation model consistent with its underlying predictive model does not guarantee relevant explanations. Table 7 discloses the set of four essays that optimize the coverage of most globally important features to evaluate the correctness of the best AES model. It is quite intriguing to note that two of the four essays are among the 16 essays that have a major disagreement (off by 2) between the resolved and predicted rubric scores (1 vs. 3 and 4 vs. 2). The AES tool clearly overrated Essay 228, while it underrated Essay 219. Naturally, these two essays offer an opportunity to understand what is wrong with the model and ultimately debug the model to improve its accuracy and interpretability.

In particular, Essay 228 raises suspicion on the positive contributions of features such as “Ortho_N,” “lemma_mattr,” “all_logical,” “det_pobj_deps_struct,” and “dobj_per_cl.” Moreover, notice how the remaining 262 less important features (not visible in the decision plot in Figure 5 ) have already inflated the rubric score beyond the base value, more than any other essay. Given the very short length and very low quality of the essay, whose meaning is seriously undermined by spelling and grammatical errors, it is of utmost importance to verify how some of these features are computed. For example, is the average number of orthographic neighbors (Ortho_N) per token computed for unmeaningful tokens such as “R” and “whe”? Similarly, are these tokens considered as types in the type-token ratio over lemmas (lemma_mattr)? Given the absence of a meaningful grammatical structure conveying a complete idea through well-articulated words, it becomes obvious that the quality of NLP (natural language processing) parsing may become a source of (measurement) bias impacting both the way some writing features are computed and the predicted rubric score. To remedy this, two solutions are proposed: (1) enhancing the dataset with the part-of-speech sequence or the structure of dependency relationships along with associated confidence levels, or (2) augmenting the essay dataset with essays enclosing various types of non-sensical content to improve the learning of these feature contributions.

Note that all four essays have a text length smaller than the average: 171 words. Notice also how the “hdd42_aw” and “hdd42_fw” play a significant role to decrease the predicted score of Essays 228 and 68. The reader should note that these metrics require a minimum of 42 tokens in order to compute a non-zero D index, a measure of lexical diversity as explained in Supplementary Table 1 . Figure 6B also shows how zero “hdd42_aw” values are heavily penalized. This is extra evidence that supports the strong role that the number of words plays in determining these rubric scores, especially for very short essays where it is one of the few observations that can be reliably recorded.

Two other issues with the best trained AES model were identified. First, in the eyes of the model, the lowest the average number of direct objects per clause (dobj_per_cl), as seen in Figure 6D , the best it is. This appears to contradict one of the requirements of the “Style” rubric, which looks for a variety of sentence structures. Remember that direct objects imply the presence of transitive verbs (action verbs) and that the balanced usage of linking verbs and action verbs as well as of transitive and intransitive verbs is key to meet the requirement of variety of sentence structures. Moreover, note that the writing feature is about counting the number of direct objects per clause, not by sentence. Only one direct object is therefore possible per clause. On the other side, a sentence may contain several clauses, which determines if the sentence is a simple, compound, or a complex sentence. This also means that a sentence may have multiple direct objects and that a high ratio of direct objects per clause is indicative of sentence complexity. Too much complexity is also undesirable. Hence, it is fair to conclude that the higher range of feature values has reasonable feature contributions (SHAP values), while the lower range does not capture well the requirements of the rubric. The dependence plot should rather display a positive peak somewhere in the middle. Notice how the poor quality of Essay 228’s single sentence prevented the proper detection of the single direct object, “broke my finger,” and the so-called absence of direct objects was one of the reasons to wrongfully improve the predicted rubric score.

The model’s second issue discussed here is the presence of sentence fragments, a type of grammatical errors. Essentially, a sentence fragment is a clause that misses one of three critical components: a subject, a verb, or a complete idea. Figure 6E shows the contribution model of grammatical errors, all types combined, while Figure 6F shows specifically the contribution model of sentence fragments. It is interesting to see how SHAP further penalizes larger numbers of grammatical errors and that it takes into account the length of the essay (red dots represent essays with larger numbers of words; blue dots represent essays with smaller numbers of words). For example, except for essays with no identified grammatical errors, longer essays are less penalized than shorter ones. This is particularly obvious when there are 2−4 grammatical errors. The model increases the predicted rubric score only when there is no grammatical error. Moreover, the model tolerates longer essays with only one grammatical error, which sounds quite reasonable. On the other side, the model finds desirable high numbers of sentence fragments, a non-trivial type of grammatical errors. Even worse, the model decreases the rubric score of essays having no sentence fragment. Although grammatical issues are beyond the scope of the “Style” rubric, the model has probably included these features because of their impact on the quality of assessment of vocabulary usage and sentence fluency. The reader should observe how the very poor quality of an essay can even prevent the detection of such fundamental grammatical errors such as in the case of Essay 228, where the AES tool did not find any grammatical error or sentence fragment. Therefore, there should be a way for AES systems to detect a minimum level of text quality before attempting to score an essay. Note that the objective of this section was not to undertake thorough debugging of the model, but rather to underscore the effectiveness of SHAP in doing so.

Formative Feedback

Once an AES model is considered reasonably valid, SHAP can be a suitable formalism to empower the machine to provide formative feedback. For instance, the explanation of Essay 124, which has been assigned a rubric score of 3 by both human and machine markers, indicates that the top two factors contributing to decreasing the predicted rubric score are: (1) the essay length being smaller than average, and (2) the average number of verb lemma types occurring at least once in the next sentence (adjacent_overlap_verb_sent). Figures 6A,H give the overall picture in which the realism of the contributions of these two features can be analyzed. More specifically, Essay 124 is one of very few essays ( Figure 6H ) that makes redundant usage of the same verbs across adjacent sentences. Moreover, the essay displays poor sentence fluency where everything is only expressed in two sentences. To understand more accurately the impact of “adjacent_overlap_verb_sent” on the prediction, a few spelling errors have been corrected and the text has been divided in four sentences instead of two. Revision 1 in Table 8 exhibits the corrections made to the original essay. The decision plot’s dashed line in Figure 3D represents the original explanation of Essay 124, while Figure 7A demonstrates the new explanation of the revised essay. It can be seen that the “adjacent_overlap_verb_sent” feature is still the second most important feature in the new explanation of Essay 124, with a feature value of 0.429, still considered as very poor according to the dependence plot in Figure 6H .

www.frontiersin.org

Table 8. Revisions of Essay 124: improvement of sentence splitting, correction of some spelling errors, and elimination of redundant usage of same verbs (bold for emphasis in Essay 124’s original version; corrections in bold for Revisions 1 and 2).

www.frontiersin.org

Figure 7. Explanations of the various versions of Essay 124 and evaluation of feature effect for a range of feature values. (A) Explanation of Essay 124’s first revision. (B) Forecasting the effect of changing the ‘adjacent_overlap_verb_sent’ feature on the rubric score. (C) Explanation of Essay 124’s second revision. (D) Comparison of the explanations of all Essay 124’s versions.

To show how SHAP could be leveraged to offer remedial formative feedback, the revised version of Essay 124 will be explained again for eight different values of “adjacent_overlap_verb_sent” (0, 0.143, 0.286, 0.429, 0.571, 0.714, 0.857, 1.0), while keeping the values of all other features constant. The set of these eight essays are explained by a newly trained SHAP explainer (Gradient), producing new SHAP values for each feature and each “revised” essay. Notice how the new model, called the feedback model, allows to foresee by how much a novice writer can hope to improve his/her score according to the “Style” rubric. If the student employs different verbs at every sentence, the feedback model estimates that the rubric score could be improved from 3.47 up to 3.65 ( Figure 7B ). Notice that the dashed line represents Revision 1, while other lines simulate one of the seven other altered essays. Moreover, it is important to note how changing the value of a single feature may influence the contributions that other features may have on the predicted score. Again, all explanations look similar in terms of direction, but certain features differ in terms of the magnitude of their contributions. However, the reader should observe how the targeted feature varies not only in terms of magnitude, but also of direction, allowing the student to ponder the relevancy of executing the recommended writing strategy.

Thus, upon receiving this feedback, assume that a student sets the goal to improve the effectiveness of his/her verb choice by eliminating any redundant verb, producing Revision 2 in Table 8 . The student submits his essay again to the AES system, which finally gives a new rubric score of 3.98, a significant improvement from the previous 3.47, allowing the student to get a 4 instead of a 3. Figure 7C exhibits the decision plot of Revision 2. To better observe how the various revisions of the student’s essay changed over time, their respective explanations have been plotted in the same decision plot ( Figure 7D ). Notice this time that the ordering of the features has changed to list the features of common importance to all of the essay’s versions. The feature ordering in Figures 7A−C complies with the same ordering as in Figure 3D , the decision plot of the original essay. These figures underscore the importance of tracking the interaction between the various features so that the model understands well the impact that changing one feature has on the others. TreeSHAP, an implementation for tree-based models, offers this capability and its potential on improving the quality of feedback provided to novice writers will be tested in a future version of this AES system.

This paper serves as a proof of concept of the applicability of XAI techniques in automated essay scoring, providing learning analytics practitioners and educators with a methodology on how to “hire” AI markers and make them accountable to their human counterparts. In addition to debug predictive models, SHAP explanation models can serve as some formalism of a broader learning analytics platform, where aspects of prescriptive analytics (provision of remedial formative feedback) can be added on top of the more pervasive predictive analytics.

However, the main weakness of the approach put forward in this paper consists in omitting many types of spatio-temporal data. In other words, it ignores precious information inherent to the writing process, which may prove essential to guess the intent of the student, especially in contexts of poor sentence structures and high grammatical inaccuracy. Hence, this paper calls for adapting current NLP technologies to educational purposes, where the quality of writing may be suboptimal, which is contrary to many utopian scenarios where NLP is used for content analysis, opinion mining, topic modeling, or fact extraction trained on corpora of high-quality texts. By capturing the writing process preceding a submission of an essay to an AES tool, other kinds of explanation models can also be trained to offer feedback not only from a linguistic perspective but also from a behavioral one (e.g., composing vs. revising); that is, the AES system could inform novice writers about suboptimal and optimal writing strategies (e.g., planning a revision phase after bursts of writing).

In addition, associating sections of text with suboptimal writing features, those whose contributions lower the predicted score, would be much more informative. This spatial information would not only allow to point out what is wrong and but also where it is wrong, answering more efficiently the question why an essay is wrong. This problem could be simply approached through a multiple-inputs and mixed-data feature-based (MLP) neural network architecture fed by both linguistic indices and textual data ( n -grams), where the SHAP explanation model would assign feature contributions to both types of features and any potential interaction between them. A more complex approach could address the problem through special types of recurrent neural networks such as Ordered-Neurons LSTMs (long short-term memory), which are well adapted to the parsing of natural language, and where the natural sequence of text is not only captured but also its hierarchy of constituents ( Shen et al., 2018 ). After all, this paper highlights the fact that the potential of deep learning can reach beyond the training of powerful predictive models and be better visible in the higher trustworthiness of explanation models. This paper also calls for optimizing the training of predictive models by considering the descriptive accuracy of explanations and the human expert’s qualitative knowledge (e.g., indicating the direction of feature contributions) during the training process.

Data Availability Statement

The datasets and code of this study can be found in these Open Science Framework’s online repositories: https://osf.io/fxvru/ .

Author Contributions

VK architected the concept of an ethics-bound, semi-autonomous, and trust-enabled human-AI fusion that measures and advances knowledge boundaries in human learning, which essentially defines the key traits of learning analytics. DB was responsible for its implementation in the area of explainable automated essay scoring and for the training and validation of the predictive and explanation models. Together they offer an XAI-based proof of concept of a prescriptive model that can offer real-time formative remedial feedback to novice writers. Both authors contributed to the article and approved its publication.

Research reported in this article was supported by the Academic Research Fund (ARF) publication grant of Athabasca University under award number (24087).

Conflict of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Supplementary Material

The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/feduc.2020.572367/full#supplementary-material

  • ^ https://www.kaggle.com/c/asap-aes
  • ^ https://www.linguisticanalysistools.org/

Abbass, H. A. (2019). Social integration of artificial intelligence: functions, automation allocation logic and human-autonomy trust. Cogn. Comput. 11, 159–171. doi: 10.1007/s12559-018-9619-0

CrossRef Full Text | Google Scholar

Adadi, A., and Berrada, M. (2018). Peeking inside the black-box: a survey on explainable artificial intelligence (XAI). IEEE Access 6, 52138–52160. doi: 10.1109/ACCESS.2018.2870052

Amorim, E., Cançado, M., and Veloso, A. (2018). “Automated essay scoring in the presence of biased ratings,” in Proceedings of the 2018 Conference of the North American Chapter of the Association for Computational Linguistics: Human Language Technologies , New Orleans, LA, 229–237.

Google Scholar

Arrieta, A. B., Díaz-Rodríguez, N., Ser, J., Del Bennetot, A., Tabik, S., Barbado, A., et al. (2020). Explainable Artificial Intelligence (XAI): concepts, taxonomies, opportunities and challenges toward responsible AI. Inform. Fusion 58, 82–115. doi: 10.1016/j.inffus.2019.12.012

Balota, D. A., Yap, M. J., Hutchison, K. A., Cortese, M. J., Kessler, B., Loftis, B., et al. (2007). The English lexicon project. Behav. Res. Methods 39, 445–459. doi: 10.3758/BF03193014

PubMed Abstract | CrossRef Full Text | Google Scholar

Boulanger, D., and Kumar, V. (2018). “Deep learning in automated essay scoring,” in Proceedings of the International Conference of Intelligent Tutoring Systems , eds R. Nkambou, R. Azevedo, and J. Vassileva (Cham: Springer International Publishing), 294–299. doi: 10.1007/978-3-319-91464-0_30

Boulanger, D., and Kumar, V. (2019). “Shedding light on the automated essay scoring process,” in Proceedings of the International Conference on Educational Data Mining , 512–515.

Boulanger, D., and Kumar, V. (2020). “SHAPed automated essay scoring: explaining writing features’ contributions to English writing organization,” in Intelligent Tutoring Systems , eds V. Kumar and C. Troussas (Cham: Springer International Publishing), 68–78. doi: 10.1007/978-3-030-49663-0_10

Chen, H., Lundberg, S., and Lee, S.-I. (2019). Explaining models by propagating Shapley values of local components. arXiv [Preprint]. Available online at: https://arxiv.org/abs/1911.11888 (accessed September 22, 2020).

Crossley, S. A., Bradfield, F., and Bustamante, A. (2019). Using human judgments to examine the validity of automated grammar, syntax, and mechanical errors in writing. J. Writ. Res. 11, 251–270. doi: 10.17239/jowr-2019.11.02.01

Crossley, S. A., Kyle, K., and McNamara, D. S. (2016). The tool for the automatic analysis of text cohesion (TAACO): automatic assessment of local, global, and text cohesion. Behav. Res. Methods 48, 1227–1237. doi: 10.3758/s13428-015-0651-7

Crossley, S. A., Kyle, K., and McNamara, D. S. (2017). Sentiment analysis and social cognition engine (SEANCE): an automatic tool for sentiment, social cognition, and social-order analysis. Behav. Res. Methods 49, 803–821. doi: 10.3758/s13428-016-0743-z

Dronen, N., Foltz, P. W., and Habermehl, K. (2015). “Effective sampling for large-scale automated writing evaluation systems,” in Proceedings of the Second (2015) ACM Conference on Learning @ Scale , 3–10.

Goldin, I., Narciss, S., Foltz, P., and Bauer, M. (2017). New directions in formative feedback in interactive learning environments. Int. J. Artif. Intellig. Educ. 27, 385–392. doi: 10.1007/s40593-016-0135-7

Hao, Q., and Tsikerdekis, M. (2019). “How automated feedback is delivered matters: formative feedback and knowledge transfer,” in Proceedings of the 2019 IEEE Frontiers in Education Conference (FIE) , Covington, KY, 1–6.

Hellman, S., Rosenstein, M., Gorman, A., Murray, W., Becker, L., Baikadi, A., et al. (2019). “Scaling up writing in the curriculum: batch mode active learning for automated essay scoring,” in Proceedings of the Sixth (2019) ACM Conference on Learning @ Scale , (New York, NY: Association for Computing Machinery).

Hussein, M. A., Hassan, H., and Nassef, M. (2019). Automated language essay scoring systems: a literature review. PeerJ Comput. Sci. 5:e208. doi: 10.7717/peerj-cs.208

Kumar, V., and Boulanger, D. (2020). Automated essay scoring and the deep learning black box: how are rubric scores determined? Int. J. Artif. Intellig. Educ. doi: 10.1007/s40593-020-00211-5

Kumar, V., Fraser, S. N., and Boulanger, D. (2017). Discovering the predictive power of five baseline writing competences. J. Writ. Anal. 1, 176–226.

Kyle, K. (2016). Measuring Syntactic Development In L2 Writing: Fine Grained Indices Of Syntactic Complexity And Usage-Based Indices Of Syntactic Sophistication. Dissertation, Georgia State University, Atlanta, GA.

Kyle, K., Crossley, S., and Berger, C. (2018). The tool for the automatic analysis of lexical sophistication (TAALES): version 2.0. Behav. Res. Methods 50, 1030–1046. doi: 10.3758/s13428-017-0924-4

Lundberg, S. M., Erion, G. G., and Lee, S.-I. (2018). Consistent individualized feature attribution for tree ensembles. arXiv [Preprint]. Available online at: https://arxiv.org/abs/1802.03888 (accessed September 22, 2020).

Lundberg, S. M., and Lee, S.-I. (2017). “A unified approach to interpreting model predictions,” in Advances in Neural Information Processing Systems , eds I. Guyon, U. V. Luxburg, S. Bengio, H. Wallach, R. Fergus, S. Vishwanathan, et al. (Red Hook, NY: Curran Associates, Inc), 4765–4774.

Madnani, N., and Cahill, A. (2018). “Automated scoring: beyond natural language processing,” in Proceedings of the 27th International Conference on Computational Linguistics , (Santa Fe: Association for Computational Linguistics), 1099–1109.

Madnani, N., Loukina, A., von Davier, A., Burstein, J., and Cahill, A. (2017). “Building better open-source tools to support fairness in automated scoring,” in Proceedings of the First (ACL) Workshop on Ethics in Natural Language Processing , (Valencia: Association for Computational Linguistics), 41–52.

McCarthy, P. M., and Jarvis, S. (2010). MTLD, vocd-D, and HD-D: a validation study of sophisticated approaches to lexical diversity assessment. Behav. Res. Methods 42, 381–392. doi: 10.3758/brm.42.2.381

Mizumoto, T., Ouchi, H., Isobe, Y., Reisert, P., Nagata, R., Sekine, S., et al. (2019). “Analytic score prediction and justification identification in automated short answer scoring,” in Proceedings of the Fourteenth Workshop on Innovative Use of NLP for Building Educational Applications , Florence, 316–325.

Molnar, C. (2020). Interpretable Machine Learning . Abu Dhabi: Lulu

Murdoch, W. J., Singh, C., Kumbier, K., Abbasi-Asl, R., and Yu, B. (2019). Definitions, methods, and applications in interpretable machine learning. Proc. Natl. Acad. Sci. U.S.A. 116, 22071–22080. doi: 10.1073/pnas.1900654116

Nelson, J., and Campbell, C. (2017). Evidence-informed practice in education: meanings and applications. Educ. Res. 59, 127–135. doi: 10.1080/00131881.2017.1314115

Rahimi, Z., Litman, D., Correnti, R., Wang, E., and Matsumura, L. C. (2017). Assessing students’ use of evidence and organization in response-to-text writing: using natural language processing for rubric-based automated scoring. Int. J. Artif. Intellig. Educ. 27, 694–728. doi: 10.1007/s40593-017-0143-2

Reinertsen, N. (2018). Why can’t it mark this one? A qualitative analysis of student writing rejected by an automated essay scoring system. English Austral. 53:52.

Ribeiro, M. T., Singh, S., and Guestrin, C. (2016). “Why should i trust you?”: explaining the predictions of any classifier. CoRR, abs/1602.0. arXiv [Preprint]. Available online at: http://arxiv.org/abs/1602.04938 (accessed September 22, 2020).

Rupp, A. A. (2018). Designing, evaluating, and deploying automated scoring systems with validity in mind: methodological design decisions. Appl. Meas. Educ. 31, 191–214. doi: 10.1080/08957347.2018.1464448

Rupp, A. A., Casabianca, J. M., Krüger, M., Keller, S., and Köller, O. (2019). Automated essay scoring at scale: a case study in Switzerland and Germany. ETS Res. Rep. Ser. 2019, 1–23. doi: 10.1002/ets2.12249

Shen, Y., Tan, S., Sordoni, A., and Courville, A. C. (2018). Ordered Neurons: Integrating Tree Structures into Recurrent Neural Networks. CoRR, abs/1810.0. arXiv [Preprint]. Available online at: http://arxiv.org/abs/1810.09536 (accessed September 22, 2020).

Shermis, M. D. (2014). State-of-the-art automated essay scoring: competition, results, and future directions from a United States demonstration. Assess. Writ. 20, 53–76. doi: 10.1016/j.asw.2013.04.001

Taghipour, K. (2017). Robust Trait-Specific Essay Scoring using Neural Networks and Density Estimators. Dissertation, National University of Singapore, Singapore.

West-Smith, P., Butler, S., and Mayfield, E. (2018). “Trustworthy automated essay scoring without explicit construct validity,” in Proceedings of the 2018 AAAI Spring Symposium Series , (New York, NY: ACM).

Woods, B., Adamson, D., Miel, S., and Mayfield, E. (2017). “Formative essay feedback using predictive scoring models,” in Proceedings of the 23rd ACM SIGKDD International Conference on Knowledge Discovery and Data Mining , (New York, NY: ACM), 2071–2080.

Keywords : explainable artificial intelligence, SHAP, automated essay scoring, deep learning, trust, learning analytics, feedback, rubric

Citation: Kumar V and Boulanger D (2020) Explainable Automated Essay Scoring: Deep Learning Really Has Pedagogical Value. Front. Educ. 5:572367. doi: 10.3389/feduc.2020.572367

Received: 14 June 2020; Accepted: 09 September 2020; Published: 06 October 2020.

Reviewed by:

Copyright © 2020 Kumar and Boulanger. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY) . The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: David Boulanger, [email protected]

This article is part of the Research Topic

Learning Analytics for Supporting Individualization: Data-informed Adaptation of Learning

Deep Learning in Automated Essay Scoring

  • Conference paper
  • First Online: 17 May 2018
  • Cite this conference paper

automated essay scoring dataset

  • David Boulanger 16 &
  • Vivekanandan Kumar 16  

Part of the book series: Lecture Notes in Computer Science ((LNPSE,volume 10858))

Included in the following conference series:

  • International Conference on Intelligent Tutoring Systems

3139 Accesses

2 Citations

This paper explores the application of deep learning in automated essay scoring (AES). It uses the essay dataset #8 from the Automated Student Assessment Prize competition, hosted by the Kaggle platform, and a state-of-the-art Suite of Automatic Linguistic Analysis Tools (SALAT) to extract 1,463 writing features. A non-linear regressor deep neural network is trained to predict holistic scores on a scale of 10–60. This study shows that deep learning holds the promise to improve significantly the accuracy of AES systems, but that the current dataset and most essay datasets fall short of providing them with enough expertise (hand-graded essays) to exploit that potential. After the tuning of different sets of hyperparameters, the results show that the levels of agreement, as measured by the quadratic weighted kappa metric, obtained on the training, validation, and testing sets are 0.84, 0.63, and 0.58, respectively, while an ensemble (bagging) produced a kappa value of 0.80 on the testing set. Finally, this paper upholds that more than 1,000 hand-graded essays per writing construct would be necessary to adequately train the predictive student models on automated essay scoring, provided that all score categories are equally or fairly represented in the sample dataset.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
  • Available as EPUB and PDF
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Similar content being viewed by others

automated essay scoring dataset

A review of deep-neural automated essay scoring models

automated essay scoring dataset

Automatically Grading Brazilian Student Essays

automated essay scoring dataset

Automated Essay Scoring and the Deep Learning Black Box: How Are Rubric Scores Determined?

Dataset and code are available at: https://1drv.ms/u/s!ApcQo2VlCqiPmXGh2_zjgDLnPzKp

Crossley, S.A., Kyle, K., McNamara, D.S.: The tool for the automatic analysis of text cohesion (TAACO): automatic assessment of local, global, and text cohesion. Behav. Res. Methods 48 (4), 1227–1237 (2016)

Article   Google Scholar  

Crossley, S.A., Kyle, K., McNamara, D.S.: Sentiment analysis and social cognition engine (SEANCE): an automatic tool for sentiment, social cognition, and social order analysis. Behav. Res. Methods 49 (3), 803–821 (2017)

Guestrin, C., Fox, E.: Machine Learning: Regression. Coursera (2017). https://www.coursera.org/learn/ml-regression . Accessed 22 Mar 2018

Kumar, V., Fraser, S.N., Boulanger, D.: Discovering the predictive power of five baseline writing competences. J. Writ. Anal. 1 (1), 176–226 (2017)

Google Scholar  

Kyle, K., Crossley, S.A.: Automatically assessing lexical sophistication: indices, tools, findings, and application. TESOL Q. 49 (4), 757–786 (2015)

Kyle, K.: Suite of Automatic Linguistic Analysis Tools (SALAT) (2016a). http://www.kristopherkyle.com/ . Accessed 25 Apr 2018

Kyle, K.: Measuring syntactic development in L2 writing: fine grained indices of syntactic complexity and usage-based indices of syntactic sophistication. Doctoral Dissertation (2016b). http://scholarworks.gsu.edu/alesl_diss/35

Ng, A.: Improving Deep Neural Networks: Hyperparameter tuning, Regularization and Optimization. Coursera (2017). https://www.coursera.org/learn/deep-neural-network . Accessed 22 Mar 2018

Rosebrock, A.: Deep Learning for Computer Vision with Python, 1st edn. PyImageSearch (2017). https://www.pyimagesearch.com/deep-learning-computer-vision-python-book/ . Accessed 22 Mar 2018

Shermis, M.D.: State-of-the-art automated essay scoring: competition, results, and future directions from a United States demonstration. Assess. Writ 20 (1), 53–76 (2014)

Zupanc, K., Bosnić, Z.: Automated essay evaluation with semantic analysis. Knowl.-Based Syst. 120 , 118–132 (2017)

Download references

Author information

Authors and affiliations.

Athabasca University, Edmonton, AB, T5J 3S8, Canada

David Boulanger & Vivekanandan Kumar

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to David Boulanger .

Editor information

Editors and affiliations.

Université du Québec, Montreal, Québec, Canada

Roger Nkambou

NCSU, Raleigh, North Carolina, USA

Roger Azevedo

University of Saskatchewan, Saskatoon, Saskatchewan, Canada

Julita Vassileva

Rights and permissions

Reprints and permissions

Copyright information

© 2018 Springer International Publishing AG, part of Springer Nature

About this paper

Cite this paper.

Boulanger, D., Kumar, V. (2018). Deep Learning in Automated Essay Scoring. In: Nkambou, R., Azevedo, R., Vassileva, J. (eds) Intelligent Tutoring Systems. ITS 2018. Lecture Notes in Computer Science(), vol 10858. Springer, Cham. https://doi.org/10.1007/978-3-319-91464-0_30

Download citation

DOI : https://doi.org/10.1007/978-3-319-91464-0_30

Published : 17 May 2018

Publisher Name : Springer, Cham

Print ISBN : 978-3-319-91463-3

Online ISBN : 978-3-319-91464-0

eBook Packages : Computer Science Computer Science (R0)

Share this paper

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Publish with us

Policies and ethics

  • Find a journal
  • Track your research

Neural Automated Essay Scoring and Coherence Modeling for Adversarially Crafted Input

We demonstrate that current state-of-the-art approaches to Automated Essay Scoring (AES) are not well-suited to capturing adversarially crafted input of grammatical but incoherent sequences of sentences. We develop a neural model of local coherence that can effectively learn connectedness features between sentences, and propose a framework for integrating and jointly training the local coherence model with a state-of-the-art AES model. We evaluate our approach against a number of baselines and experimentally demonstrate its effectiveness on both the AES task and the task of flagging adversarial input, further contributing to the development of an approach that strengthens the validity of neural essay scoring models.

1 Introduction

Automated Essay Scoring (AES) focuses on automatically analyzing the quality of writing and assigning a score to the text. Typically, AES models exploit a wide range of manually-tuned shallow and deep linguistic features  Shermis and Hammer ( 2012 ) ; Burstein et al. ( 2003 ) ; Rudner et al. ( 2006 ) ; Williamson et al. ( 2012 ) ; Andersen et al. ( 2013 ) . Recent advances in deep learning have shown that neural approaches to AES achieve state-of-the-art results   Alikaniotis et al. ( 2016 ) ; Taghipour and Ng ( 2016 ) with the additional advantage of utilizing features that are automatically learned from the data. In order to facilitate interpretability of neural models, a number of visualization techniques have been proposed to identify textual (superficial) features that contribute to model performance   Alikaniotis et al. ( 2016 ) .

To the best of our knowledge, however, no prior work has investigated the robustness of neural AES systems to adversarially crafted input that is designed to trick the model into assigning desired missclassifications; for instance, a high score to a low quality text. Examining and addressing such validity issues is critical and imperative for AES deployment. Previous work has primarily focused on assessing the robustness of ‘‘standard’’ machine learning approaches that rely on manual feature engineering; for example, Powers et al. ( 2002 ) ; Yannakoudakis et al. ( 2011 ) have shown that such AES systems, unless explicitly designed to handle adversarial input, can be susceptible to subversion by writers who understand something of the systems’ workings and can exploit this to maximize their score.

In this paper, we make the following contributions:

We examine the robustness of state-of-the-art neural AES models to adversarially crafted input, 1 1 1 We use the terms ‘adversarially crafted input’ and ‘adversarial input’ to refer to text that is designed with the intention to trick the system. and specifically focus on input related to local coherence ; that is, grammatical but incoherent sequences of sentences. 2 2 2 Coherence can be assessed locally in terms of transitions between adjacent sentences. In addition to the superiority in performance of neural approaches against ‘‘standard’’ machine learning models Alikaniotis et al. ( 2016 ) ; Taghipour and Ng ( 2016 ) , such a setup allows us to investigate their potential superiority / capacity in handling adversarial input without being explicitly designed to do so.

We demonstrate that state-of-the-art neural AES is not well-suited to capturing adversarial input of grammatical but incoherent sequences of sentences, and develop a neural model of local coherence that can effectively learn connectedness features between sentences.

A local coherence model is typically evaluated based on its ability to rank coherently ordered sequences of sentences higher than their incoherent / permuted counterparts (e.g., Barzilay and Lapata ( 2008 ) ). We focus on a stricter evaluation setting in which the model is tested on its ability to rank coherent sequences of sentences higher than any incoherent / permuted set of sentences, and not just its own permuted counterparts. This supports a more rigorous evaluation that facilitates development of more robust models.

We propose a framework for integrating and jointly training the local coherence model with a state-of-the-art AES model. We evaluate our approach against a number of baselines and experimentally demonstrate its effectiveness on both the AES task and the task of flagging adversarial input, further contributing to the development of an approach that strengthens AES validity.

At the outset, our goal is to develop a framework that strengthens the validity of state-of-the-art neural AES approaches with respect to adversarial input related to local aspects of coherence. For our experiments, we use the Automated Student Assessment Prize (ASAP) dataset, 3 3 3 https://www.kaggle.com/c/asap-aes/ which contains essays written by students ranging from Grade 7 to Grade 10 in response to a number of different prompts (see Section 4 ).

2 Related Work

AES Evaluation against Adversarial Input One of the earliest attempts at evaluating AES models against adversarial input was by Powers et al. ( 2002 ) who asked writing experts -- that had been briefed on how the e-Rater scoring system works -- to write essays to trick e-Rater Burstein et al. ( 1998 ) . The participants managed to fool the system into assigning higher-than-deserved grades, most notably by simply repeating a few well-written paragraphs several times. Yannakoudakis et al. ( 2011 ) and Yannakoudakis and Briscoe ( 2012 ) created and used an adversarial dataset of well-written texts and their random sentence permutations, which they released in the public domain, together with the grades assigned by a human expert to each piece of text. Unfortunately, however, the dataset is quite small, consisting of 12 12 12 texts in total. Higgins and Heilman ( 2014 ) proposed a framework for evaluating the susceptibility of AES systems to gaming behavior. Neural AES Models Alikaniotis et al. ( 2016 ) developed a deep bidirectional Long Short-Term Memory (LSTM)  Hochreiter and Schmidhuber ( 1997 ) network, augmented with score-specific word embeddings that capture both contextual and usage information for words. Their approach outperformed traditional feature-engineered AES models on the ASAP dataset. Taghipour and Ng ( 2016 ) investigated various recurrent and convolutional architectures on the same dataset and found that an LSTM layer followed by a Mean over Time operation achieves state-of-the-art results.   Dong and Zhang ( 2016 ) showed that a two-layer Convolutional Neural Network (CNN) outperformed other baselines (e.g., Bayesian Linear Ridge Regression) on both in-domain and domain-adaptation experiments on the ASAP dataset. Neural Coherence Models A number of approaches have investigated neural models of coherence on news data.  Li and Hovy ( 2014 ) used a window approach where a sliding kernel of weights was applied over neighboring sentence representations to extract local coherence features. The sentence representations were constructed with recursive and recurrent neural methods. Their approach outperformed previous methods on the task of selecting maximally coherent sentence orderings from sets of candidate permutations   Barzilay and Lapata ( 2008 ) .   Lin et al. ( 2015 ) developed a hierarchical Recurrent Neural Network (RNN) for document modeling. Among others, they looked at capturing coherence between sentences using a sentence-level language model, and evaluated their approach on the sentence ordering task. Tien Nguyen and Joty ( 2017 ) built a CNN over entity grid representations, and trained the network in a pairwise ranking fashion. Their model outperformed other graph-based and distributed sentence models. We note that our goal is not to identify the ‘‘best’’ model of local coherence on randomly permuted grammatical sentences in the domain of AES, but rather to propose a framework that strengthens the validity of AES approaches with respect to adversarial input related to local aspects of coherence.

Refer to caption

3.1 Local Coherence (LC) Model

Our local coherence model is inspired by the model of  Li and Hovy ( 2014 ) which uses a window approach to evaluate coherence. 4 4 4 We note that Li and Jurafsky ( 2017 ) also present an extended version of the work by Li and Hovy ( 2014 ) , evaluated on different domains. Figure  1 presents a visual representation of the network architecture, which is described below in detail. Sentence Representation This part of the model composes the sentence representations that can be utilized to learn connectedness features between sentences. Each word in the text is initialized with a k 𝑘 k -dimensional vector w 𝑤 w from a pre-trained word embedding space. Unlike Li and Hovy ( 2014 ) , we use an LSTM layer 5 5 5 LSTMs have been shown to produce state-of-the-art results in AES   Alikaniotis et al. ( 2016 ); Taghipour and Ng ( 2016 ) . to capture sentence compositionality by mapping words in a sentence s = { w 1 , w 2 , … , w n } 𝑠 subscript 𝑤 1 subscript 𝑤 2 … subscript 𝑤 𝑛 s=\{w_{1},w_{2},...,w_{n}\} at each time step t 𝑡 t ( w t subscript 𝑤 𝑡 w_{t} , where t ≤ n 𝑡 𝑛 t\leq n ) onto a fixed-size vector h t w ​ r ​ d ∈ ℝ d l ​ s ​ t ​ m superscript subscript ℎ 𝑡 𝑤 𝑟 𝑑 superscript ℝ subscript 𝑑 𝑙 𝑠 𝑡 𝑚 h_{t}^{wrd}\in\mathbb{R}^{d_{lstm}} (where d l ​ s ​ t ​ m subscript 𝑑 𝑙 𝑠 𝑡 𝑚 {d_{lstm}} is a hyperparameter). The sentence representation h s ​ n ​ t superscript ℎ 𝑠 𝑛 𝑡 h^{snt} is then the representation of the last word in the sentence:

(1)

Clique Representation Each window of sentences in a text represents a clique q = { s 1 , … , s m } 𝑞 subscript 𝑠 1 … subscript 𝑠 𝑚 q=\{s_{1},...,s_{m}\} , where m 𝑚 m is a hyperparameter indicating the window size. A clique is assigned a score of 1 1 1 if it is coherent (i.e., the sentences are not shuffled) and 0 0 if it is incoherent (i.e., the sentences are shuffled). The clique embedding is created by concatenating the representations of the sentences it contains according to Equation  1 . A convolutional operation -- using a filter W c ​ l ​ q ∈ ℝ m × d l ​ s ​ t ​ m × d c ​ n ​ n superscript 𝑊 𝑐 𝑙 𝑞 superscript ℝ 𝑚 subscript 𝑑 𝑙 𝑠 𝑡 𝑚 subscript 𝑑 𝑐 𝑛 𝑛 W^{clq}\in\mathbb{R}^{m\times d_{lstm}\times d_{cnn}} , where d c ​ n ​ n subscript 𝑑 𝑐 𝑛 𝑛 d_{cnn} denotes the convolutional output size -- is then applied to the clique embedding, followed by a non-linearity in order to extract the clique representation h c ​ l ​ q ∈ ℝ d c ​ n ​ n superscript ℎ 𝑐 𝑙 𝑞 superscript ℝ subscript 𝑑 𝑐 𝑛 𝑛 h^{clq}\in\mathbb{R}^{d_{cnn}} :

(2)

𝑁 𝑚 1 j\in\{1,...,N-m+1\} , N 𝑁 N is the number of sentences in the text, and ∗ * is the linear convolutional operation. Scoring The cliques’ predicted scores are calculated via a linear operation followed by a sigmoid function to project the predictions to a [ 0 , 1 ] 0 1 [0,1] probability space:

(3)

where V ∈ ℝ d c ​ n ​ n 𝑉 superscript ℝ subscript 𝑑 𝑐 𝑛 𝑛 V\in\mathbb{R}^{d_{cnn}} is a learned weight. The network optimizes its parameters to minimize the negative log-likelihood of the cliques’ gold scores y c ​ l ​ q superscript 𝑦 𝑐 𝑙 𝑞 y^{clq} , given the network’s predicted scores:

(4)

𝑁 𝑚 1 T=N-m+1 (number of cliques in text). The final prediction of the text’s coherence score is calculated as the average of all of its clique scores:

(5)

This is in contrast to  Li and Hovy ( 2014 ) , who multiply all the estimated clique scores to generate the overall document score. This means that if only one clique is misclassified as incoherent and assigned a score of 0 0 , the whole document is regarded as incoherent. We aim to soften this assumption and use the average instead to allow for a more fine-grained modeling of degrees of coherence. 6 6 6 Our experiments showed that using the multiplicative approach gives poor results, as presented in Section 6 .

We train the LC model on synthetic data automatically generated by creating random permutations of highly-scored ASAP essays (Section 4 ).

Refer to caption

3.2 LSTM AES Model

We utilize the LSTM AES model of Taghipour and Ng ( 2016 ) shown in Figure  2 (LSTM T&N ), which is trained, and yields state-of-the-art results on the ASAP dataset. The model is a one-layer LSTM that encodes the sequence of words in an essay, followed by a Mean over Time operation that averages the word representations generated from the LSTM layer. 7 7 7 We note that the authors achieve slightly higher results when averaging ensemble results of their LSTM model together with CNN models. We use their main LSTM model which, for the purposes of our experiments, does not affect our conclusions.

3.3 Combined Models

We propose a framework for integrating the LSTM T&N model with the Local Coherence (LC) one. Our goal is to have a robust AES system that is able to correctly flag adversarial input while maintaining a high performance on essay scoring.

3.3.1 Baseline: Vector Concatenation (VecConcat)

The baseline model simply concatenates the output representations of the pre-prediction layers of the trained LSTM T&N and LC networks, and feeds the resulting vector to a machine learning algorithm (e.g., Support Vector Machines, SVMs) to predict the final overall score. In the LSTM T&N model, the output representation (hereafter referred to as the essay representation ) is the vector produced from the Mean Over Time operation; in the LC model, we use the generated clique representations (Figure  1 ) aggregated with a max operation; 8 8 8 We note that max aggregation outperformed other aggregation functions. (hereafter referred to as the clique representation ). Although the LC model is trained on permuted ASAP essays (Section 4 ) and the LSTM T&N model on the original ASAP data, essay and clique representations are generated for both the ASAP and the synthetic essays containing reordered sentences.

3.3.2 Joint Learning

Instead of training the LSTM T&N and LC models separately and then concatenating their output representations, we propose a framework where both models are trained jointly, and where the final network has then the capacity to predict AES scores and flag adversarial input (Figure 3 ).

Specifically, the LSTM T&N and LC networks predict an essay and coherence score respectively (as described earlier), but now they both share the word embedding layer. The training set is the aggregate of both the ASAP and permuted data to allow the final network to learn from both simultaneously. Concretely, during training, for the ASAP essays, we assume that both the gold essay and coherence scores are the same and equal to the gold ASAP scores. This is not too strict an assumption, as overall scores of writing competence tend to correlate highly with overall coherence. For the synthetic essays, we set the ‘‘gold’’ coherence scores to zero, and the ‘‘gold’’ essay scores to those of their original non-permuted counterparts in the ASAP dataset. The intuition is as follows: firstly, setting the ‘‘gold’’ essay scores of synthetic essays to zero would bias the model into over-predicting zeros; secondly, our approach reinforces the LSTM T&N ’s inability to detect adversarial input, and forces the overall network to rely on the LC branch to identify such input. 9 9 9 We note that, during training, the scores are mapped to a range between 0 and 1 (similarly to Taghipour and Ng ( 2016 ) ), and then scaled back to their original range during evaluation.

The two sub-networks are trained together and the error gradients are back-propagated to the word embeddings. To detect whether an essay is adversarial, we further augment the system with an adversarial text detection component that simply captures adversarial input based on the difference between the predicted essay and coherence scores. Specifically, we use our development set to learn a threshold for this difference, and flag an essay as adversarial if the difference is larger than the threshold. We experimentally demonstrate that this approach enables the model to perform well on both original ASAP and synthetic essays. During model evaluation, the texts flagged as adversarial by the model are assigned a score of zero, while the rest are assigned the predicted essay score ( y ^ e ​ s ​ y superscript ^ 𝑦 𝑒 𝑠 𝑦 \hat{y}^{esy} in Figure 3 ).

Refer to caption

Prompt #ASAP essays Score Range Synthetic Dataset
threshold #ASAP essays total size
1 1,783 2--12 10 472 5,192
2 1,800 1--6 5 82 902
3 1,726 0--3 3 407 4,477
4 1,772 0--3 3 244 2,684
5 1,805 0--4 4 258 2,838
6 1,800 0--4 4 367 4,037
7 1,569 0--30 23 179 1,969
8 723 0--60 45 72 792

4 Data and Evaluation

To create adversarial input, we select high scoring essays per prompt (given a pre-defined score threshold, Table 1 ) 10 10 10 We note that this threshold is different than the one mentioned in Section 3.3.2 . that are assumed coherent, and create 10 10 10 permutations per essay by randomly shuffling its sentences. In the joint learning setup, we augment the original ASAP dataset with a subset of the synthetic essays. Specifically, we randomly select 4 4 4 permutations per essay to include in the training set, 11 11 11 This is primarily done to keep the data balanced: initial experiments showed that training with all 10 10 10 permutations per essay harms AES performance, but has negligible effect on adversarial input detection. but include all 10 10 10 permutations in the test set. Table 1 presents the details of the datasets. We test performance on the ASAP dataset using Quadratic Weighted Kappa (QWK), which was the official evaluation metric in the ASAP competition, while we test performance on the synthetic dataset using pairwise ranking accuracy (PRA) between an original non-permuted essay and its permuted counterparts. PRA is typically used as an evaluation metric on coherence assessment tasks on other domains  Barzilay and Lapata ( 2008 ) , and is based on the fraction of correct pairwise rankings in the test data (i.e., a coherent essay should be ranked higher than its permuted counterpart). Herein, we extend this metric and furthermore evaluate the models by comparing each original essay to all adversarial / permuted essays in the test data, and not just its own permuted counterparts -- we refer to this metric as total pairwise ranking accuracy (TPRA).

5 Model Parameters and Baselines

Coherence models We train and test the LC model described in Section 3.1 on the synthetic dataset and evaluate it using PRA and TPRA. During pre-processing, words are lowercased and initialized with pre-trained word embeddings Zou et al. ( 2013 ) . Words that occur only once in the training set are mapped to a special UNK embedding. All network weights are initialized to values drawn randomly from a uniform distribution with scale = 0.05 absent 0.05 =0.05 , and biases are initialized to zeros. We apply a learning rate of 0.001 0.001 0.001 and RMSProp   Tieleman and Hinton ( 2012 ) for optimization. A size of 100 100 100 is chosen for the hidden layers ( d l ​ s ​ t ​ m subscript 𝑑 𝑙 𝑠 𝑡 𝑚 d_{lstm} and d c ​ n ​ n subscript 𝑑 𝑐 𝑛 𝑛 d_{cnn} ), and the convolutional window size ( m 𝑚 m ) is set to 3 3 3 . Dropout  Srivastava et al. ( 2014 ) is applied for regularization to the output of the convolutional operation with probability 0.3 0.3 0.3 . The network is trained for 60 60 60 epochs and performance is monitored on the development sets -- we select the model that yields the highest PRA value. 12 12 12 Our implementation is available at https://github.com/Youmna-H/Coherence_AES

We use as a baseline the LC model that is based on the multiplication of the clique scores (similarly to Li and Hovy ( 2014 ) ), and compare the results (LC mul ) to our averaged approach. As another baseline, we use the entity grid (EGrid) Barzilay and Lapata ( 2008 ) that models transitions between sentences based on sequences of entity mentions labeled with their grammatical role. EGrid has been shown to give competitive results on similar coherence tasks in other domains. Using the Brown Coherence Toolkit Eisner and Charniak ( 2011 ) , 13 13 13 https://bitbucket.org/melsner/browncoherence we construct the entity transition probabilities with length = 3 3 3 and salience = 2 2 2 . The transition probabilities are then used as features that are fed as input to an SVM classifier with an RBF kernel and penalty parameter C = 1.5 𝐶 1.5 C=1.5 to predict a coherence score. LSTM T&N model We replicate and evaluate the LSTM model of Taghipour and Ng ( 2016 ) 14 14 14 https://github.com/nusnlp/nea on ASAP and our synthetic data. Combined models After training the LC and LSTM T&N models, we concatenate their output vectors to build the Baseline: Vector Concatenation (VecConcat) model as described in Section 3.3.1 , and train a Kernel Ridge Regression model. 15 15 15 We use scikit-learn with the following parameters: alpha= 0.1 0.1 0.1 , coef0= 1 1 1 , degree= 3 3 3 , gamma= 0.1 0.1 0.1 , kernel=‘rbf’.

The Joint Learning network is trained on both the ASAP and synthetic dataset as described in Section 3.3.2 . Adversarial input is detected based on an estimated threshold on the difference between the predicted essay and coherence scores (Figure 3 ). The threshold value is empirically calculated on the development sets, and set to be the average difference between the predicted essay and coherence scores in the synthetic data:

where M 𝑀 M is the number of synthetic essays in the development set.

We furthermore evaluate a baseline where the joint model is trained without sharing the word embedding layer between the two sub-models, and report the effect on performance (Joint Learning no_layer_sharing ). Finally, we evaluate a baseline where for the joint model we set the ‘‘gold’’ essay scores of synthetic data to zero (Joint Learning zero_score ), as opposed to our proposed approach of setting them to be the same as the score of their original non-permuted counterpart in the ASAP dataset.

The state-of-the-art LSTM T&N model, as shown in Table  2 , gives the highest performance on the ASAP data, but is not robust to adversarial input and therefore unable to capture aspects of local coherence, with performance on synthetic data that is less than 0.5 0.5 0.5 . On the other hand, both our LC model and the EGrid significantly outperform LSTM T&N on synthetic data. While EGrid is slightly better in terms of TPRA compared to LC ( 0.706 0.706 0.706 vs. 0.689 0.689 0.689 ), LC is substantially better on PRA ( 0.946 0.946 0.946 vs. 0.718 0.718 0.718 ). This could be attributed to the fact that LC is optimised using PRA on the development set. The LC mul variation has a performance similar to LC in terms of PRA, but is significantly worse in terms of TPRA, which further supports the use of our proposed LC model.

Our Joint Learning model manages to exploit the best of both the LSTM T&N and LC approaches: performance on synthetic data is significantly better compared to LSTM T&N (and in particular gives the highest TPRA value on synthetic data compared to all models), while manages to maintain the high performance of LSTM T&N on ASAP data (performance slighly drops from 0.739 0.739 0.739 to 0.724 0.724 0.724 though not significantly). When the Joint Learning model is compared against the VecConcat baseline, we can again confirm its superiority on both datasets, giving significant differences on synthetic data.

Model ASAP Synthetic
QWK PRA TPRA
EGrid 0.718 0.706
LC 0.946 0.689
LC 0.948 0.620
LSTM 0.739 0.430 0.473
VecConcat 0.719 0.614 0.567
Joint Learning 0.724 0.770 0.777

7 Further Analysis

We furthermore evaluate the performance of the the Joint Learning model when trained using different parameters (Table  3 ). When assigning ‘‘gold’’ essay scores of zero to adversarial essays (Joint Learning zero_score ), AES performance on the ASAP data drops to 0.449 0.449 0.449 QWK, and the results are statistically significant. 16 16 16 Note that we do not report performance of this model on synthetic data. In this case, the thresholding technique cannot be applied as both sub-models are trained with the same “gold” scores and thus have very similar predictions on synthetic data. This is partly explained by the fact that the model, given the training data gold scores, is biased towards predicting zeros. The result, however, further supports our hypothesis that forcing the Joint Learning model to rely on the coherence branch for adversarial input detection further improves performance. Importantly, we need something more than just training a state-of-the-art AES model (in our case, LSTM T&N ) on both original and synthetic data.

Model ASAP Synthetic
QWK PRA TPRA
Joint Learning 0.724 0.770 0.777
Joint Learning 0.720 0.741 0.753
Joint Learning 0.449

Refer to caption

We also compare Joint Learning to Joint Learning no_layer_sharing in which the the two sub-models are trained separately without sharing the first layer of word representations. While the difference in performance on the ASAP test data is small, the differences are much larger on synthetic data, and are significant in terms of TPRA. By examining the false positives of both systems (i.e., the coherent essays that are misclassified as adversarial), we find that when the embeddings are not shared, the system is biased towards flagging long essays as adversarial, while interestingly, this bias is not present when the embeddings are shared. For instance, the average number of words in the false positive cases of Joint Learning no_layer_sharing on the ASAP data is 426 426 426 , and the average number of sentences is 26 26 26 ; on the other hand, with the Joint Learning model, these numbers are 340 340 340 and 19 19 19 respectively. 17 17 17 Adversarial texts in the synthetic dataset have an average number of 306 306 306 words and an average number of 18 18 18 sentences. A possible explanation for this is that training the words with more contextual information (in our case, via embeddings sharing), is advantageous for longer documents with a large number of sentences.

Ideally, no essays in the ASAP data should be flagged as adversarial as they were not designed to trick the system. We calculate the number of ASAP texts incorrectly detected as adversarial, and find that the average error in the Joint Learning model is quite small ( 0.382 % percent 0.382 0.382\% ). This increases with Joint Learning no_layer_sharing ( 1 % percent 1 1\% ), although still remains relatively small.

We further investigate the essay and coherence scores predicted by our best model, Joint Learning, for the permuted and original ASAP essays in the synthetic dataset (for which we assume that the selected, highly scored ASAP essays are coherent, Section 4 ), and present results for 3 3 3 randomly selected prompts in Figure  4 . The graphs show a large difference between predicted essay and coherence scores on permuted / adversarial data ((a), (b) and (c)), where the system predicts high essay scores for permuted texts (as a result of our training strategy), but low coherence scores (as predicted by the LC model). For highly scored ASAP essays ((d), (e) and (f)), the system predictions are less varied and positively contributes to the performance of our proposed approach.

8 Conclusion

We evaluated the robustness of state-of-the-art neural AES approaches on adversarial input of grammatical but incoherent sequences of sentences, and demonstrated that they are not well-suited to capturing such cases. We created a synthetic dataset of such adversarial examples and trained a neural local coherence model that is able to discriminate between such cases and their coherent counterparts. We furthermore proposed a framework for jointly training the coherence model with a state-of-the-art neural AES model, and introduced an effective strategy for assigning ‘‘gold’’ scores to adversarial input during training. When compared against a number of baselines, our joint model achieves better performance on randomly permuted sentences, while maintains a high performance on the AES task. Among others, our results demonstrate that it is not enough to simply (re-)train neural AES models with adversarially crafted input, nor is it sufficient to rely on ‘‘simple’’ approaches that concatenate output representations from different neural models. Finally, our framework strengthens the validity of neural AES approaches with respect to adversarial input designed to trick the system.

Acknowledgements

We gratefully acknowledge the support of NVIDIA Corporation with the donation of the Titan X Pascal GPU used for this research. We are also grateful to Cambridge Assessment for their support of the ALTA Institute. Special thanks to Christopher Bryant and Marek Rei for their valuable feedback.

  • Alikaniotis et al. (2016) Dimitrios Alikaniotis, Helen Yannakoudakis, and Marek Rei. 2016. Automatic text scoring using neural networks. In Proceedings of the 54th Annual Meeting of the Association for Computational Linguistics (Volume 1: Long Papers) . Association for Computational Linguistics, pages 715--725.
  • Andersen et al. (2013) Øistein E Andersen, Helen Yannakoudakis, Fiona Barker, and Tim Parish. 2013. Developing and testing a self-assessment and tutoring system. In Proceedings of the Eighth Workshop on Innovative Use of NLP for Building Educational Applications, BEA . Association for Computational Linguistics, pages 32--41.
  • Barzilay and Lapata (2008) Regina Barzilay and Mirella Lapata. 2008. Modeling local coherence: An entity-based approach. Computational Linguistics 34(1):1--34.
  • Burstein et al. (2003) Jill Burstein, Martin Chodorow, and Claudia Leacock. 2003. Criterion: Online essay evaluation: An application for automated evaluation of student essays. In Proceedings of the fifteenth annual conference on innovative applications of artificial intelligence . American Association for Artificial Intelligence, pages 3--10.
  • Burstein et al. (1998) Jill Burstein, Karen Kukich, Susanne Wolff, Chi Lu, Martin Chodorow, Lisa Braden-Harder, and Mary Dee Harris. 1998. Automated scoring using a hybrid feature identification technique. In Proceedings of the 36th Annual Meeting of the Association for Computational Linguistics and 17th International Conference on Computational Linguistics - Volume 1 . pages 206--210.
  • Dong and Zhang (2016) Fei Dong and Yue Zhang. 2016. Automatic features for essay scoring -- an empirical study. In Proceedings of the 2016 Conference on Empirical Methods in Natural Language Processing . pages 1072--1077.
  • Eisner and Charniak (2011) Micha Eisner and Eugene Charniak. 2011. Extending the entity grid with entity-specific features. In Proceedings of the 49th Annual Meeting of the Association for Computational Linguistics: Human Language Technologies: Short Papers - Volume 2 . Association for Computational Linguistics, pages 125--129.
  • Higgins and Heilman (2014) Derrick Higgins and Michael Heilman. 2014. Managing what we can measure: Quantifying the susceptibility of automated scoring systems to gaming behavior. Educational Measurement: Issues and Practice 33:36–46.
  • Hochreiter and Schmidhuber (1997) Sepp Hochreiter and Jürgen Schmidhuber. 1997. Long short-term memory. Neural computation 9(8):1735--1780.
  • Li and Hovy (2014) Jiwei Li and Eduard Hovy. 2014. A model of coherence based on distributed sentence representation. In Proceedings of the 2014 Conference on Empirical Methods in Natural Language Processing (EMNLP) . pages 2039--2048.
  • Li and Jurafsky (2017) Jiwei Li and Dan Jurafsky. 2017. Neural net models for open-domain discourse coherence. In Proceedings of the 2017 Conference on Empirical Methods in Natural Language Processing . pages 198--209.
  • Lin et al. (2015) Rui Lin, Shujie Liu, Muyun Yang, Mu Li, Ming Zhou, and Sheng Li. 2015. Hierarchical recurrent neural network for document modeling. In Proceedings of the 2015 Conference on Empirical Methods in Natural Language Processing . pages 899--907.
  • Powers et al. (2002) Donald E. Powers, Jill Burstein, Martin Chodorow, Mary E. Fowles, and Karen Kukich. 2002. Stumping e-rater: challenging the validity of automated essay scoring. Computers in Human Behavior 18(2):103--134.
  • Rudner et al. (2006) LM Rudner, Veronica Garcia, and Catherine Welch. 2006. An evaluation of IntelliMetric essay scoring system. The Journal of Technology, Learning, and Assessment 4(4):1 -- 22.
  • Shermis and Hammer (2012) M Shermis and B Hammer. 2012. Contrasting state-of-the-art automated scoring of essays: analysis. In Annual National Council on Measurement in Education Meeting . pages 1--54.
  • Srivastava et al. (2014) Nitish Srivastava, Geoffrey E Hinton, Alex Krizhevsky, Ilya Sutskever, and Ruslan Salakhutdinov. 2014. Dropout: a simple way to prevent neural networks from overfitting. Journal of Machine Learning Research 15(1):1929--1958.
  • Taghipour and Ng (2016) Kaveh Taghipour and Hwee Tou Ng. 2016. A neural approach to automated essay scoring. In Proceedings of the 2016 Conference on Empirical Methods in Natural Language Processing . Association for Computational Linguistics, pages 1882--1891.
  • Tieleman and Hinton (2012) Tijmen Tieleman and Geoffrey Hinton. 2012. Lecture 6.5 - rmsprop. Technical report .
  • Tien Nguyen and Joty (2017) Dat Tien Nguyen and Shafiq Joty. 2017. A neural local coherence model. In Proceedings of the 55th Annual Meeting of the Association for Computational Linguistics (Volume 1: Long Papers) . Association for Computational Linguistics, pages 1320--1330.
  • Williamson et al. (2012) DM Williamson, Xiaoming Xi, and FJ Breyer. 2012. A framework for evaluation and use of automated scoring. Educational Measurement: Issues and Practice 31(1):2--13.
  • Yannakoudakis and Briscoe (2012) Helen Yannakoudakis and Ted Briscoe. 2012. Modeling coherence in esol learner texts. In Proceedings of the Seventh Workshop on Building Educational Applications Using NLP . Association for Computational Linguistics, pages 33--43.
  • Yannakoudakis et al. (2011) Helen Yannakoudakis, Ted Briscoe, and Ben Medlock. 2011. A new dataset and method for automatically grading esol texts. In Proceedings of the 49th Annual Meeting of the Association for Computational Linguistics: Human Language Technologies-Volume 1 . Association for Computational Linguistics, pages 180--189.
  • Zou et al. (2013) Will Y Zou, Richard Socher, Daniel M Cer, and Christopher D Manning. 2013. Bilingual word embeddings for phrase-based machine translation. In EMNLP . pages 1393--1398.

ar5iv homepage

Automated Essay Grading

Two Bears Playing Chess

  • Give Me More Feedback I: Annotating Argument Persuasiveness and Related Attributes in Student Essays . Winston Carlile, Nishant Gurrapadi, Zixuan Ke and Vincent Ng Proceedings of the 56th Annual Meeting of the Association for Computational Linguistics (Volume 1: Long Papers), 2018. - Data used in this paper is here .
  • Give Me More Feedback II: Annotating Thesis Strength and Related Attributes in Student Essays . Zixuan Ke, Hrishikesh Inamdar, Hui Lin and Vincent Ng. Proceedings of the 57th Annual Meeting of the Association for Computational Linguistics (Volume 1: Long Papers), 2019. - Data used in this paper is here .

Subscribe to the PwC Newsletter

Join the community, edit social preview.

automated essay scoring dataset

Add a new code entry for this paper

Remove a code repository from this paper.

automated essay scoring dataset

Mark the official implementation from paper authors

Add a new evaluation result row.

TASK DATASET MODEL METRIC NAME METRIC VALUE GLOBAL RANK REMOVE
  • AUTOMATED ESSAY SCORING

Remove a task

Add a method.

  • ABSOLUTE POSITION ENCODINGS
  • ATTENTION DROPOUT
  • DENSE CONNECTIONS
  • LABEL SMOOTHING
  • LAYER NORMALIZATION
  • LINEAR LAYER
  • LINEAR WARMUP WITH LINEAR DECAY
  • MULTI-HEAD ATTENTION
  • POSITION-WISE FEED-FORWARD LAYER
  • RESIDUAL CONNECTION
  • SCALED DOT-PRODUCT ATTENTION
  • SENTENCEPIECE
  • TRANSFORMER
  • WEIGHT DECAY

Remove a method

  • ABSOLUTE POSITION ENCODINGS -
  • ATTENTION DROPOUT -
  • DENSE CONNECTIONS -
  • LABEL SMOOTHING -
  • LAYER NORMALIZATION -
  • LINEAR LAYER -
  • LINEAR WARMUP WITH LINEAR DECAY -
  • MULTI-HEAD ATTENTION -
  • POSITION-WISE FEED-FORWARD LAYER -
  • RESIDUAL CONNECTION -
  • SCALED DOT-PRODUCT ATTENTION -
  • SENTENCEPIECE -
  • TRANSFORMER -
  • WEIGHT DECAY -
  • WORDPIECE -

Edit Datasets

Language models and automated essay scoring.

18 Sep 2019  ·  Pedro Uria Rodriguez , Amir Jafari , Christopher M. Ormerod · Edit social preview

In this paper, we present a new comparative study on automatic essay scoring (AES). The current state-of-the-art natural language processing (NLP) neural network architectures are used in this work to achieve above human-level accuracy on the publicly available Kaggle AES dataset. We compare two powerful language models, BERT and XLNet, and describe all the layers and network architectures in these models. We elucidate the network architectures of BERT and XLNet using clear notation and diagrams and explain the advantages of transformer architectures over traditional recurrent neural network architectures. Linear algebra notation is used to clarify the functions of transformers and attention mechanisms. We compare the results with more traditional methods, such as bag of words (BOW) and long short term memory (LSTM) networks.

Code Edit Add Remove Mark official

Tasks edit add remove, datasets edit, results from the paper edit, methods edit add remove.

Navigation Menu

Search code, repositories, users, issues, pull requests..., provide feedback.

We read every piece of feedback, and take your input very seriously.

Saved searches

Use saved searches to filter your results more quickly.

To see all available qualifiers, see our documentation .

  • Notifications You must be signed in to change notification settings

Automatically grades the essays written by users

satin-04/Automatic-Essay-Scoring

Folders and files.

NameName
2 Commits

Repository files navigation

Automatic-essay-scoring (aes).

Automated Essay Scoring (AES) is a tool for evaluating and scoring of essays written in response to specific prompts. It can be defined as the process of scoring written essays using computer programs. The process of automating the assessment process could be useful for both educators and learners since it encourages the iterative improvements of students' writings.

Automated grading if proven effective will not only reduce the time for assessment but comparing it with human scores will also make the score realistic. The project aims to develop an automated essay assessment system by use of machine learning techniques and Neural networks by classifying a corpus of textual entities into a small number of discrete categories, corresponding to possible grades.

The dataset we are using is ‘The Hewlett Foundation: Automated Essay Scoring Dataset’ by ASAP. You can find in the below link or download from the Dataset folder.

https://www.kaggle.com/c/asap-aes/data

Architecture Diagram

arch

Proposed Model

We create a list of words from each sentence and from each essay. This list is fed into the Word2Vec model. This model makes sense of the available words by assigning numerical vector values to each word. Features are generated by passing the essays through Word2Vec model. The Word2Vec model acts as an Embedding Layer in a neural network. Features from this model are passed through our LSTM layers. We implement 2 LSTM layers. The first layer accepts all features from the Embedding Layer (Word2Vec) as input and passes 300 features as output to the second LSTM layer. The second layer accepts 300 features as input and 64 features as output. Next we add a Dropout layer with value 0.5. Finally a fully connected Dense Layer with output 1 which represents the score of Essay. The model was compiled with loss function Mean Squared Error and Optimizer Root Mean Square. The model was trained for 150 epochs with batch size of 64.

But first the model achieved is divided into 4 modules as follows:

1. Data Preprocessing

We began by doing some standard preprocessing steps like filling in null values and selecting valid features from the entire dataset after a thorough study.Next we plotted a graph to get a measure of the skewness of our data and applied normalisation techniques to reduce this skewness.The next step involved cleaning the essays to make our training process easier for getting a better accuracy.To achieve this we removed all the unnecessary symbols ,stop words and punctuations from our essays. To increase our accuracy even more we even planned to add some extra features like the number of sentences , number of words,number of characters, average word length etc. Moreover , we even worked on techniques like getting the noun ,verb ,adjective and adverb counts using parts of speech tagging as well as getting the total misspellings in an essay by comparison with a corpus.We applied various machine learning algorithms on this data as explained in the next section.

Processed dataset can be found in the file, Processed_data.csv

2. Machine Learning

For making our data ready to apply algorithms,we require one more step.Machine learning algorithms can not be applied on sentences or words,they can only be used upon numeric data.Our dataset has a field which has essays that need to be converted into a numeric form first in order to train it.To do this we use something known as a CountVectorizer. Now the CountVectorizer works by tokenizing a collection of text documents and returning an encoded vector with a length of the entire vocabulary along with an integer count for the number of times each word appeared in the document.After this step our data is finally ready for predictive modelling.

Initially we applied machine learning algorithms like linear regression, SVR and Random Forest on the dataset without addition of features that were mentioned in the preprocessing section before. Our results were not really satisfactory as our mean squared error was quite high for all the above algorithms. After this initial evaluation, we added the extra features,applied CountVectorizer again on this modified dataset and applied the same three algorithms.There was a great improvement in the performance of all three algorithms especially Random forest for which the mean squared error reduced drastically.

Python notebook for the implementation of this module can be found in the file, Essay_Scoring_1.ipynb

3. Applying Neural Networks

Preprocessing steps for neural networks are different from preprocessing steps for machine learning algorithms. Our training data is fed into the Embedding Layer which is Word2Vec. Word2Vec is a shallow, two-layer neural network which is trained to reconstruct linguistic contexts of words. It takes as its input a large corpus of words and produces a vector space, typically of several hundred dimensions, with each unique word in the corpus being assigned a corresponding vector in the space. Word vectors are positioned in the vector space such that words that share common contexts in the corpus are located in close proximity to one another in the space. Word2Vec is a particularly computationally-efficient predictive model for learning word embeddings from raw text. Features from Word2Vec are fed into LSTM. LSTM can learn which data in a sequence is important to keep or throw away. This largely helps in calculating scores from essays. Finally the Dense layer with output 1 predicts the score of each essay.

NN

4. Creation of web App

After training our model, the next step was to make our project available to users.For this purpose we planned to develop a web application for our model. To accomplish this we used the Flask framework to deploy our model. Flask is a popular Python web framework, meaning it is a thirdparty Python library used for developing web applications.Using Flask we were able to make an API which receives essay details through GUI and computes the predicted score value based on our model.The results can be shown by making a POST request . It receives JSON inputs, uses the trained model to make a prediction and returns that prediction in JSON format which can be accessed through the API endpoint.

The essential of the webpage can be found in the folder, webapp . Screenshots of the webpage can be found in the following links:

webpage1 jpg

In this project, we introduced a deep neural network model capable of representing both local and contextual usage of information by essay scoring. This model yields score-specific word embeddings used later by a recurrent neural network in order to form essay representations. We have shown that this kind of architecture is able to surpass similar state-of-the-art systems. We also introduced a novel way of exploring the basis of the network’s internal scoring criteria and showed that such models are interpretable and can be further explored to provide useful feedback to the author.

It was satisfying that our neural network model using 300-dimensional LSTM as initialization to the embedding layer was our most successful model. We believe that a more extensive hyperparameter search with our LSTM based models could outperform this result. There are many ideas moving forward. Trying out the models in Ensemble mode is also an extension we wish to try out in the near future.

  • Jupyter Notebook 54.1%
  • Python 0.4%

automated essay scoring dataset

Automated Essay Scoring — Kaggle Competition End to End Project Implementation

Part 2: data preprocessing and trainign with lstm.

Mayur Pethani

Mayur Pethani

Analytics Vidhya

Kindly go through Part 1 , Part 2 and Part 3 for complete understanding and project execution with given Github .

  • Training LSTM Model.ipynb for training and saving the model.

Importing the Data

  • Gensim, NLTK, Django libraries have been added.
  • Constants have been added in terms of DATASET_DIR, GLOVE_DIR, SAVE_DIR paths.
  • Loading data using pandas library from training_set_rel3.tsv.
  • Removing unnecessary columns like domain_score and raters_domain.
  • Defining Minimum and Maximum score which we will be using at the time of predicting the actual score.

Preprocessing the Data

We will preprocess all essays and convert them to feature vectors so that they can be fed into the RNN.

These are all helper functions used to clean the essays.

  • There are 4 functions defined:
  • getAvgFeatureVecs : This function accepts 3 parameters: essays, model, num_features. It internall calls makeFeatureVec function to convert essays into FeatureVector.
  • makeFeatureVec: This function accepts 3 parameters: words, model, num_features. Using Word2Vec index2word function and np.divide it gives ultimately average feature vectors for the passed model.
  • essay_to_sentense : This function accepts 2 parameters: essay_v, remove_stopwords. it internally calls essay_to_wordlist and converts essays to sentences.
  • essay_to_wordlist : This function accepts 2 parameters: essay_v, remove_stopwords. It removes the stopwords and returns words.
  • Whenever you are working with NLP Machine Learning and Deep Learning tasks the above mentioned steps are almost neccessary because machine understands numbers or we can say that computation is very easy when we use numbers here we refer to vectors.
  • We are trying to convert essay or corpus to first sentenses and then to words which can also be called are tokens and then convert them to vectors.
I would strongly suggest to go through few NLP terminology and concepts like Tokenizer, Stemming, Limitization, Stopwords, different methods to convert words to Vectors like BOW, TF-IDF, n-gram. These are NLP data pre-processing techniques before feeding to any machine learning algorithms or deep learning algorithms.

Defining the model

Here we define a 2-Layer LSTM Model.

Note that instead of using sigmoid activation in the output layer we will use Relu since we are not normalising training labels.

  • /models folder contains 6 different models which you should try and check accuracy.
  • As part of training, you just need to replace above code with those model files code

Training Phase

Now we train the model on the dataset.

We will use 5-Fold Cross Validation and measure the Quadratic Weighted Kappa for each fold. We will then calculate Average Kappa for all the folds.

  • We are first training the essays using Word2Vec model which is available in gensim library. Later on, we are saving into word2vecmodel.bin file which we will be using at the time of predicting the score.
  • Now we are using the functions which we have previously defined to convert essay to vector representation.
  • We are also passing this vectors into LSTM model and saving the model in final_lstm.h5 file.
  • As part of result, we are calculating cohen kapp score with quadratic weights for 5 times using KFold cross validation and then we are taking the average. As you can see the result will be
  • A Neural Approach to Automated Essay Scoring
  • Automatic Text Scoring Using Neural Networks

The above research papers clearly explains the importance of using cohen’s kappa score as well as different Models they have tried in their research and which model has given the best result.

IMPORTANT: In practical project implementation also this is very important that you must try with different models to get maximum accuracy and then that model will be saved and used as a part of production.

Next, we will be going through Web Application code where we will see how saved model is actually predicting the score in Part 3 .

References:

  • Special thanks goes to Ronit Mankad’s Github Repo which I had forked and got end to end understanding about this project.
  • References taken from Wikipedia and other websites.

If you really like this article series, kindly clap, follow me and enjoy the extreme power of artificial intelligence just like below minions.

Mayur Pethani

Written by Mayur Pethani

Data Scientist

Text to speech

ACM Digital Library home

  • Advanced Search

An automated essay scoring systems: a systematic literature review

School of Computer Science and Artificial Intelligence, SR University, Warangal, TS, India

Research Scholar, JNTU, Hyderabad, India

Department of Information Technology, JNTUH College of Engineering, Nachupally, Kondagattu, Jagtial, TS, India

  • 13 citation

New Citation Alert added!

This alert has been successfully added and will be sent to:

You will be notified whenever a record that you have chosen has been cited.

To manage your alert preferences, click on the button below.

New Citation Alert!

Please log in to your account

  • Publisher Site

Artificial Intelligence Review

ACM Digital Library

Assessment in the Education system plays a significant role in judging student performance. The present evaluation system is through human assessment. As the number of teachers' student ratio is gradually increasing, the manual evaluation process becomes complicated. The drawback of manual evaluation is that it is time-consuming, lacks reliability, and many more. This connection online examination system evolved as an alternative tool for pen and paper-based methods. Present Computer-based evaluation system works only for multiple-choice questions, but there is no proper evaluation system for grading essays and short answers. Many researchers are working on automated essay grading and short answer scoring for the last few decades, but assessing an essay by considering all parameters like the relevance of the content to the prompt, development of ideas, Cohesion, and Coherence is a big challenge till now. Few researchers focused on Content-based evaluation, while many of them addressed style-based assessment. This paper provides a systematic literature review on automated essay scoring systems. We studied the Artificial Intelligence and Machine Learning techniques used to evaluate automatic essay scoring and analyzed the limitations of the current studies and research trends. We observed that the essay evaluation is not done based on the relevance of the content and coherence.

Recommendations

Essay scoring tool by employing roberta architecture.

The automated essay scoring (AES) has significant importance in machine grading of student essays particularly in standardized exams like the Graduate Record Examination (GRE). However, some issues in AES have remained unsolved over the past several ...

Fully Automated Short Answer Scoring of the Trial Tests for Common Entrance Examinations for Japanese University

Studies on automated short-answer scoring (SAS) have been conducted to apply natural language processing to education. Short-answer scoring is a task to grade the responses from linguistic information. Most answer sheets for short-answer questions ...

Automated Essay Scoring via Example-Based Learning

Automated essay scoring (AES) is the task of assigning grades to essays. It can be applied for quality assessment as well as pricing on User Generated Content. Previous works mainly consider using the prompt information for scoring. However, some ...

Login options

Check if you have access through your login credentials or your institution to get full access on this article.

Full Access

  • Information
  • Contributors

Published in

© The Author(s), under exclusive licence to Springer Nature B.V. 2021

In-Cooperation

Kluwer Academic Publishers

United States

Publication History

  • Published: 1 March 2022

Author Tags

  • Short answer scoring
  • Essay grading
  • Natural language processing
  • Deep learning
  • research-article

Funding Sources

Other metrics.

  • Bibliometrics
  • Citations 13

Article Metrics

  • 13 Total Citations View Citations
  • 0 Total Downloads
  • Downloads (Last 12 months) 0
  • Downloads (Last 6 weeks) 0

Digital Edition

View this article in digital edition.

Share this Publication link

https://dl.acm.org/doi/abs/10.1007/s10462-021-10068-2

Share on Social Media

  • 0 References

Export Citations

  • Please download or close your previous search result export first before starting a new bulk export. Preview is not available. By clicking download, a status dialog will open to start the export process. The process may take a few minutes but once it finishes a file will be downloadable from your browser. You may continue to browse the DL while the export process is in progress. Download
  • Download citation
  • Copy citation

We are preparing your search results for download ...

We will inform you here when the file is ready.

Your file of search results citations is now ready.

Your search export query has expired. Please try again.

figshare

Automated Essay Grading Pre-Processed Dataset

The dataset is for the work “An Empirical Analysis of Machine Learning Models for Automated Essay Grading” - The dataset is a pre-processed version of the used publically available Hewlett Foundation’s Automated Student Assessment Prize (ASAP) dataset for experimental evaluation. Users can freely sign-up on Kaggle and download the raw dataset. This dataset has been extensively used in literature for evaluating automatic grading techniques (Dong and Zhang, 2016)(Cummins et al., 2016). We process the raw data and provide the processed data. Details in our paper titled: An Empirical Analysis of Machine Learning Models for Automated Essay Grading.

Ashoka University

Usage metrics.

  • Analog electronics and interfaces

CC BY 4.0

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Open access
  • Published: 03 June 2024

Applying large language models for automated essay scoring for non-native Japanese

  • Wenchao Li 1 &
  • Haitao Liu 2  

Humanities and Social Sciences Communications volume  11 , Article number:  723 ( 2024 ) Cite this article

413 Accesses

14 Altmetric

Metrics details

  • Language and linguistics

Recent advancements in artificial intelligence (AI) have led to an increased use of large language models (LLMs) for language assessment tasks such as automated essay scoring (AES), automated listening tests, and automated oral proficiency assessments. The application of LLMs for AES in the context of non-native Japanese, however, remains limited. This study explores the potential of LLM-based AES by comparing the efficiency of different models, i.e. two conventional machine training technology-based methods (Jess and JWriter), two LLMs (GPT and BERT), and one Japanese local LLM (Open-Calm large model). To conduct the evaluation, a dataset consisting of 1400 story-writing scripts authored by learners with 12 different first languages was used. Statistical analysis revealed that GPT-4 outperforms Jess and JWriter, BERT, and the Japanese language-specific trained Open-Calm large model in terms of annotation accuracy and predicting learning levels. Furthermore, by comparing 18 different models that utilize various prompts, the study emphasized the significance of prompts in achieving accurate and reliable evaluations using LLMs.

Similar content being viewed by others

automated essay scoring dataset

Scoring method of English composition integrating deep learning in higher vocational colleges

automated essay scoring dataset

ChatGPT-3.5 as writing assistance in students’ essays

automated essay scoring dataset

Detecting contract cheating through linguistic fingerprint

Conventional machine learning technology in aes.

AES has experienced significant growth with the advancement of machine learning technologies in recent decades. In the earlier stages of AES development, conventional machine learning-based approaches were commonly used. These approaches involved the following procedures: a) feeding the machine with a dataset. In this step, a dataset of essays is provided to the machine learning system. The dataset serves as the basis for training the model and establishing patterns and correlations between linguistic features and human ratings. b) the machine learning model is trained using linguistic features that best represent human ratings and can effectively discriminate learners’ writing proficiency. These features include lexical richness (Lu, 2012 ; Kyle and Crossley, 2015 ; Kyle et al. 2021 ), syntactic complexity (Lu, 2010 ; Liu, 2008 ), text cohesion (Crossley and McNamara, 2016 ), and among others. Conventional machine learning approaches in AES require human intervention, such as manual correction and annotation of essays. This human involvement was necessary to create a labeled dataset for training the model. Several AES systems have been developed using conventional machine learning technologies. These include the Intelligent Essay Assessor (Landauer et al. 2003 ), the e-rater engine by Educational Testing Service (Attali and Burstein, 2006 ; Burstein, 2003 ), MyAccess with the InterlliMetric scoring engine by Vantage Learning (Elliot, 2003 ), and the Bayesian Essay Test Scoring system (Rudner and Liang, 2002 ). These systems have played a significant role in automating the essay scoring process and providing quick and consistent feedback to learners. However, as touched upon earlier, conventional machine learning approaches rely on predetermined linguistic features and often require manual intervention, making them less flexible and potentially limiting their generalizability to different contexts.

In the context of the Japanese language, conventional machine learning-incorporated AES tools include Jess (Ishioka and Kameda, 2006 ) and JWriter (Lee and Hasebe, 2017 ). Jess assesses essays by deducting points from the perfect score, utilizing the Mainichi Daily News newspaper as a database. The evaluation criteria employed by Jess encompass various aspects, such as rhetorical elements (e.g., reading comprehension, vocabulary diversity, percentage of complex words, and percentage of passive sentences), organizational structures (e.g., forward and reverse connection structures), and content analysis (e.g., latent semantic indexing). JWriter employs linear regression analysis to assign weights to various measurement indices, such as average sentence length and total number of characters. These weights are then combined to derive the overall score. A pilot study involving the Jess model was conducted on 1320 essays at different proficiency levels, including primary, intermediate, and advanced. However, the results indicated that the Jess model failed to significantly distinguish between these essay levels. Out of the 16 measures used, four measures, namely median sentence length, median clause length, median number of phrases, and maximum number of phrases, did not show statistically significant differences between the levels. Additionally, two measures exhibited between-level differences but lacked linear progression: the number of attributives declined words and the Kanji/kana ratio. On the other hand, the remaining measures, including maximum sentence length, maximum clause length, number of attributive conjugated words, maximum number of consecutive infinitive forms, maximum number of conjunctive-particle clauses, k characteristic value, percentage of big words, and percentage of passive sentences, demonstrated statistically significant between-level differences and displayed linear progression.

Both Jess and JWriter exhibit notable limitations, including the manual selection of feature parameters and weights, which can introduce biases into the scoring process. The reliance on human annotators to label non-native language essays also introduces potential noise and variability in the scoring. Furthermore, an important concern is the possibility of system manipulation and cheating by learners who are aware of the regression equation utilized by the models (Hirao et al. 2020 ). These limitations emphasize the need for further advancements in AES systems to address these challenges.

Deep learning technology in AES

Deep learning has emerged as one of the approaches for improving the accuracy and effectiveness of AES. Deep learning-based AES methods utilize artificial neural networks that mimic the human brain’s functioning through layered algorithms and computational units. Unlike conventional machine learning, deep learning autonomously learns from the environment and past errors without human intervention. This enables deep learning models to establish nonlinear correlations, resulting in higher accuracy. Recent advancements in deep learning have led to the development of transformers, which are particularly effective in learning text representations. Noteworthy examples include bidirectional encoder representations from transformers (BERT) (Devlin et al. 2019 ) and the generative pretrained transformer (GPT) (OpenAI).

BERT is a linguistic representation model that utilizes a transformer architecture and is trained on two tasks: masked linguistic modeling and next-sentence prediction (Hirao et al. 2020 ; Vaswani et al. 2017 ). In the context of AES, BERT follows specific procedures, as illustrated in Fig. 1 : (a) the tokenized prompts and essays are taken as input; (b) special tokens, such as [CLS] and [SEP], are added to mark the beginning and separation of prompts and essays; (c) the transformer encoder processes the prompt and essay sequences, resulting in hidden layer sequences; (d) the hidden layers corresponding to the [CLS] tokens (T[CLS]) represent distributed representations of the prompts and essays; and (e) a multilayer perceptron uses these distributed representations as input to obtain the final score (Hirao et al. 2020 ).

figure 1

AES system with BERT (Hirao et al. 2020 ).

The training of BERT using a substantial amount of sentence data through the Masked Language Model (MLM) allows it to capture contextual information within the hidden layers. Consequently, BERT is expected to be capable of identifying artificial essays as invalid and assigning them lower scores (Mizumoto and Eguchi, 2023 ). In the context of AES for nonnative Japanese learners, Hirao et al. ( 2020 ) combined the long short-term memory (LSTM) model proposed by Hochreiter and Schmidhuber ( 1997 ) with BERT to develop a tailored automated Essay Scoring System. The findings of their study revealed that the BERT model outperformed both the conventional machine learning approach utilizing character-type features such as “kanji” and “hiragana”, as well as the standalone LSTM model. Takeuchi et al. ( 2021 ) presented an approach to Japanese AES that eliminates the requirement for pre-scored essays by relying solely on reference texts or a model answer for the essay task. They investigated multiple similarity evaluation methods, including frequency of morphemes, idf values calculated on Wikipedia, LSI, LDA, word-embedding vectors, and document vectors produced by BERT. The experimental findings revealed that the method utilizing the frequency of morphemes with idf values exhibited the strongest correlation with human-annotated scores across different essay tasks. The utilization of BERT in AES encounters several limitations. Firstly, essays often exceed the model’s maximum length limit. Second, only score labels are available for training, which restricts access to additional information.

Mizumoto and Eguchi ( 2023 ) were pioneers in employing the GPT model for AES in non-native English writing. Their study focused on evaluating the accuracy and reliability of AES using the GPT-3 text-davinci-003 model, analyzing a dataset of 12,100 essays from the corpus of nonnative written English (TOEFL11). The findings indicated that AES utilizing the GPT-3 model exhibited a certain degree of accuracy and reliability. They suggest that GPT-3-based AES systems hold the potential to provide support for human ratings. However, applying GPT model to AES presents a unique natural language processing (NLP) task that involves considerations such as nonnative language proficiency, the influence of the learner’s first language on the output in the target language, and identifying linguistic features that best indicate writing quality in a specific language. These linguistic features may differ morphologically or syntactically from those present in the learners’ first language, as observed in (1)–(3).

我-送了-他-一本-书

Wǒ-sòngle-tā-yī běn-shū

1 sg .-give. past- him-one .cl- book

“I gave him a book.”

Agglutinative

彼-に-本-を-あげ-まし-た

Kare-ni-hon-o-age-mashi-ta

3 sg .- dat -hon- acc- give.honorification. past

Inflectional

give, give-s, gave, given, giving

Additionally, the morphological agglutination and subject-object-verb (SOV) order in Japanese, along with its idiomatic expressions, pose additional challenges for applying language models in AES tasks (4).

足-が 棒-に なり-ました

Ashi-ga bo-ni nar-mashita

leg- nom stick- dat become- past

“My leg became like a stick (I am extremely tired).”

The example sentence provided demonstrates the morpho-syntactic structure of Japanese and the presence of an idiomatic expression. In this sentence, the verb “なる” (naru), meaning “to become”, appears at the end of the sentence. The verb stem “なり” (nari) is attached with morphemes indicating honorification (“ます” - mashu) and tense (“た” - ta), showcasing agglutination. While the sentence can be literally translated as “my leg became like a stick”, it carries an idiomatic interpretation that implies “I am extremely tired”.

To overcome this issue, CyberAgent Inc. ( 2023 ) has developed the Open-Calm series of language models specifically designed for Japanese. Open-Calm consists of pre-trained models available in various sizes, such as Small, Medium, Large, and 7b. Figure 2 depicts the fundamental structure of the Open-Calm model. A key feature of this architecture is the incorporation of the Lora Adapter and GPT-NeoX frameworks, which can enhance its language processing capabilities.

figure 2

GPT-NeoX Model Architecture (Okgetheng and Takeuchi 2024 ).

In a recent study conducted by Okgetheng and Takeuchi ( 2024 ), they assessed the efficacy of Open-Calm language models in grading Japanese essays. The research utilized a dataset of approximately 300 essays, which were annotated by native Japanese educators. The findings of the study demonstrate the considerable potential of Open-Calm language models in automated Japanese essay scoring. Specifically, among the Open-Calm family, the Open-Calm Large model (referred to as OCLL) exhibited the highest performance. However, it is important to note that, as of the current date, the Open-Calm Large model does not offer public access to its server. Consequently, users are required to independently deploy and operate the environment for OCLL. In order to utilize OCLL, users must have a PC equipped with an NVIDIA GeForce RTX 3060 (8 or 12 GB VRAM).

In summary, while the potential of LLMs in automated scoring of nonnative Japanese essays has been demonstrated in two studies—BERT-driven AES (Hirao et al. 2020 ) and OCLL-based AES (Okgetheng and Takeuchi, 2024 )—the number of research efforts in this area remains limited.

Another significant challenge in applying LLMs to AES lies in prompt engineering and ensuring its reliability and effectiveness (Brown et al. 2020 ; Rae et al. 2021 ; Zhang et al. 2021 ). Various prompting strategies have been proposed, such as the zero-shot chain of thought (CoT) approach (Kojima et al. 2022 ), which involves manually crafting diverse and effective examples. However, manual efforts can lead to mistakes. To address this, Zhang et al. ( 2021 ) introduced an automatic CoT prompting method called Auto-CoT, which demonstrates matching or superior performance compared to the CoT paradigm. Another prompt framework is trees of thoughts, enabling a model to self-evaluate its progress at intermediate stages of problem-solving through deliberate reasoning (Yao et al. 2023 ).

Beyond linguistic studies, there has been a noticeable increase in the number of foreign workers in Japan and Japanese learners worldwide (Ministry of Health, Labor, and Welfare of Japan, 2022 ; Japan Foundation, 2021 ). However, existing assessment methods, such as the Japanese Language Proficiency Test (JLPT), J-CAT, and TTBJ Footnote 1 , primarily focus on reading, listening, vocabulary, and grammar skills, neglecting the evaluation of writing proficiency. As the number of workers and language learners continues to grow, there is a rising demand for an efficient AES system that can reduce costs and time for raters and be utilized for employment, examinations, and self-study purposes.

This study aims to explore the potential of LLM-based AES by comparing the effectiveness of five models: two LLMs (GPT Footnote 2 and BERT), one Japanese local LLM (OCLL), and two conventional machine learning-based methods (linguistic feature-based scoring tools - Jess and JWriter).

The research questions addressed in this study are as follows:

To what extent do the LLM-driven AES and linguistic feature-based AES, when used as automated tools to support human rating, accurately reflect test takers’ actual performance?

What influence does the prompt have on the accuracy and performance of LLM-based AES methods?

The subsequent sections of the manuscript cover the methodology, including the assessment measures for nonnative Japanese writing proficiency, criteria for prompts, and the dataset. The evaluation section focuses on the analysis of annotations and rating scores generated by LLM-driven and linguistic feature-based AES methods.

Methodology

The dataset utilized in this study was obtained from the International Corpus of Japanese as a Second Language (I-JAS) Footnote 3 . This corpus consisted of 1000 participants who represented 12 different first languages. For the study, the participants were given a story-writing task on a personal computer. They were required to write two stories based on the 4-panel illustrations titled “Picnic” and “The key” (see Appendix A). Background information for the participants was provided by the corpus, including their Japanese language proficiency levels assessed through two online tests: J-CAT and SPOT. These tests evaluated their reading, listening, vocabulary, and grammar abilities. The learners’ proficiency levels were categorized into six levels aligned with the Common European Framework of Reference for Languages (CEFR) and the Reference Framework for Japanese Language Education (RFJLE): A1, A2, B1, B2, C1, and C2. According to Lee et al. ( 2015 ), there is a high level of agreement (r = 0.86) between the J-CAT and SPOT assessments, indicating that the proficiency certifications provided by J-CAT are consistent with those of SPOT. However, it is important to note that the scores of J-CAT and SPOT do not have a one-to-one correspondence. In this study, the J-CAT scores were used as a benchmark to differentiate learners of different proficiency levels. A total of 1400 essays were utilized, representing the beginner (aligned with A1), A2, B1, B2, C1, and C2 levels based on the J-CAT scores. Table 1 provides information about the learners’ proficiency levels and their corresponding J-CAT and SPOT scores.

A dataset comprising a total of 1400 essays from the story writing tasks was collected. Among these, 714 essays were utilized to evaluate the reliability of the LLM-based AES method, while the remaining 686 essays were designated as development data to assess the LLM-based AES’s capability to distinguish participants with varying proficiency levels. The GPT 4 API was used in this study. A detailed explanation of the prompt-assessment criteria is provided in Section Prompt . All essays were sent to the model for measurement and scoring.

Measures of writing proficiency for nonnative Japanese

Japanese exhibits a morphologically agglutinative structure where morphemes are attached to the word stem to convey grammatical functions such as tense, aspect, voice, and honorifics, e.g. (5).

食べ-させ-られ-まし-た-か

tabe-sase-rare-mashi-ta-ka

[eat (stem)-causative-passive voice-honorification-tense. past-question marker]

Japanese employs nine case particles to indicate grammatical functions: the nominative case particle が (ga), the accusative case particle を (o), the genitive case particle の (no), the dative case particle に (ni), the locative/instrumental case particle で (de), the ablative case particle から (kara), the directional case particle へ (e), and the comitative case particle と (to). The agglutinative nature of the language, combined with the case particle system, provides an efficient means of distinguishing between active and passive voice, either through morphemes or case particles, e.g. 食べる taberu “eat concusive . ” (active voice); 食べられる taberareru “eat concusive . ” (passive voice). In the active voice, “パン を 食べる” (pan o taberu) translates to “to eat bread”. On the other hand, in the passive voice, it becomes “パン が 食べられた” (pan ga taberareta), which means “(the) bread was eaten”. Additionally, it is important to note that different conjugations of the same lemma are considered as one type in order to ensure a comprehensive assessment of the language features. For example, e.g., 食べる taberu “eat concusive . ”; 食べている tabeteiru “eat progress .”; 食べた tabeta “eat past . ” as one type.

To incorporate these features, previous research (Suzuki, 1999 ; Watanabe et al. 1988 ; Ishioka, 2001 ; Ishioka and Kameda, 2006 ; Hirao et al. 2020 ) has identified complexity, fluency, and accuracy as crucial factors for evaluating writing quality. These criteria are assessed through various aspects, including lexical richness (lexical density, diversity, and sophistication), syntactic complexity, and cohesion (Kyle et al. 2021 ; Mizumoto and Eguchi, 2023 ; Ure, 1971 ; Halliday, 1985 ; Barkaoui and Hadidi, 2020 ; Zenker and Kyle, 2021 ; Kim et al. 2018 ; Lu, 2017 ; Ortega, 2015 ). Therefore, this study proposes five scoring categories: lexical richness, syntactic complexity, cohesion, content elaboration, and grammatical accuracy. A total of 16 measures were employed to capture these categories. The calculation process and specific details of these measures can be found in Table 2 .

T-unit, first introduced by Hunt ( 1966 ), is a measure used for evaluating speech and composition. It serves as an indicator of syntactic development and represents the shortest units into which a piece of discourse can be divided without leaving any sentence fragments. In the context of Japanese language assessment, Sakoda and Hosoi ( 2020 ) utilized T-unit as the basic unit to assess the accuracy and complexity of Japanese learners’ speaking and storytelling. The calculation of T-units in Japanese follows the following principles:

A single main clause constitutes 1 T-unit, regardless of the presence or absence of dependent clauses, e.g. (6).

ケンとマリはピクニックに行きました (main clause): 1 T-unit.

If a sentence contains a main clause along with subclauses, each subclause is considered part of the same T-unit, e.g. (7).

天気が良かった の で (subclause)、ケンとマリはピクニックに行きました (main clause): 1 T-unit.

In the case of coordinate clauses, where multiple clauses are connected, each coordinated clause is counted separately. Thus, a sentence with coordinate clauses may have 2 T-units or more, e.g. (8).

ケンは地図で場所を探して (coordinate clause)、マリはサンドイッチを作りました (coordinate clause): 2 T-units.

Lexical diversity refers to the range of words used within a text (Engber, 1995 ; Kyle et al. 2021 ) and is considered a useful measure of the breadth of vocabulary in L n production (Jarvis, 2013a , 2013b ).

The type/token ratio (TTR) is widely recognized as a straightforward measure for calculating lexical diversity and has been employed in numerous studies. These studies have demonstrated a strong correlation between TTR and other methods of measuring lexical diversity (e.g., Bentz et al. 2016 ; Čech and Miroslav, 2018 ; Çöltekin and Taraka, 2018 ). TTR is computed by considering both the number of unique words (types) and the total number of words (tokens) in a given text. Given that the length of learners’ writing texts can vary, this study employs the moving average type-token ratio (MATTR) to mitigate the influence of text length. MATTR is calculated using a 50-word moving window. Initially, a TTR is determined for words 1–50 in an essay, followed by words 2–51, 3–52, and so on until the end of the essay is reached (Díez-Ortega and Kyle, 2023 ). The final MATTR scores were obtained by averaging the TTR scores for all 50-word windows. The following formula was employed to derive MATTR:

\({\rm{MATTR}}({\rm{W}})=\frac{{\sum }_{{\rm{i}}=1}^{{\rm{N}}-{\rm{W}}+1}{{\rm{F}}}_{{\rm{i}}}}{{\rm{W}}({\rm{N}}-{\rm{W}}+1)}\)

Here, N refers to the number of tokens in the corpus. W is the randomly selected token size (W < N). \({F}_{i}\) is the number of types in each window. The \({\rm{MATTR}}({\rm{W}})\) is the mean of a series of type-token ratios (TTRs) based on the word form for all windows. It is expected that individuals with higher language proficiency will produce texts with greater lexical diversity, as indicated by higher MATTR scores.

Lexical density was captured by the ratio of the number of lexical words to the total number of words (Lu, 2012 ). Lexical sophistication refers to the utilization of advanced vocabulary, often evaluated through word frequency indices (Crossley et al. 2013 ; Haberman, 2008 ; Kyle and Crossley, 2015 ; Laufer and Nation, 1995 ; Lu, 2012 ; Read, 2000 ). In line of writing, lexical sophistication can be interpreted as vocabulary breadth, which entails the appropriate usage of vocabulary items across various lexicon-grammatical contexts and registers (Garner et al. 2019 ; Kim et al. 2018 ; Kyle et al. 2018 ). In Japanese specifically, words are considered lexically sophisticated if they are not included in the “Japanese Education Vocabulary List Ver 1.0”. Footnote 4 Consequently, lexical sophistication was calculated by determining the number of sophisticated word types relative to the total number of words per essay. Furthermore, it has been suggested that, in Japanese writing, sentences should ideally have a length of no more than 40 to 50 characters, as this promotes readability. Therefore, the median and maximum sentence length can be considered as useful indices for assessment (Ishioka and Kameda, 2006 ).

Syntactic complexity was assessed based on several measures, including the mean length of clauses, verb phrases per T-unit, clauses per T-unit, dependent clauses per T-unit, complex nominals per clause, adverbial clauses per clause, coordinate phrases per clause, and mean dependency distance (MDD). The MDD reflects the distance between the governor and dependent positions in a sentence. A larger dependency distance indicates a higher cognitive load and greater complexity in syntactic processing (Liu, 2008 ; Liu et al. 2017 ). The MDD has been established as an efficient metric for measuring syntactic complexity (Jiang, Quyang, and Liu, 2019 ; Li and Yan, 2021 ). To calculate the MDD, the position numbers of the governor and dependent are subtracted, assuming that words in a sentence are assigned in a linear order, such as W1 … Wi … Wn. In any dependency relationship between words Wa and Wb, Wa is the governor and Wb is the dependent. The MDD of the entire sentence was obtained by taking the absolute value of governor – dependent:

MDD = \(\frac{1}{n}{\sum }_{i=1}^{n}|{\rm{D}}{{\rm{D}}}_{i}|\)

In this formula, \(n\) represents the number of words in the sentence, and \({DD}i\) is the dependency distance of the \({i}^{{th}}\) dependency relationship of a sentence. Building on this, the annotation of sentence ‘Mary-ga-John-ni-keshigomu-o-watashita was [Mary- top -John- dat -eraser- acc -give- past] ’. The sentence’s MDD would be 2. Table 3 provides the CSV file as a prompt for GPT 4.

Cohesion (semantic similarity) and content elaboration aim to capture the ideas presented in test taker’s essays. Cohesion was assessed using three measures: Synonym overlap/paragraph (topic), Synonym overlap/paragraph (keywords), and word2vec cosine similarity. Content elaboration and development were measured as the number of metadiscourse markers (type)/number of words. To capture content closely, this study proposed a novel-distance based representation, by encoding the cosine distance between the essay (by learner) and essay task’s (topic and keyword) i -vectors. The learner’s essay is decoded into a word sequence, and aligned to the essay task’ topic and keyword for log-likelihood measurement. The cosine distance reveals the content elaboration score in the leaners’ essay. The mathematical equation of cosine similarity between target-reference vectors is shown in (11), assuming there are i essays and ( L i , …. L n ) and ( N i , …. N n ) are the vectors representing the learner and task’s topic and keyword respectively. The content elaboration distance between L i and N i was calculated as follows:

\(\cos \left(\theta \right)=\frac{{\rm{L}}\,\cdot\, {\rm{N}}}{\left|{\rm{L}}\right|{\rm{|N|}}}=\frac{\mathop{\sum }\nolimits_{i=1}^{n}{L}_{i}{N}_{i}}{\sqrt{\mathop{\sum }\nolimits_{i=1}^{n}{L}_{i}^{2}}\sqrt{\mathop{\sum }\nolimits_{i=1}^{n}{N}_{i}^{2}}}\)

A high similarity value indicates a low difference between the two recognition outcomes, which in turn suggests a high level of proficiency in content elaboration.

To evaluate the effectiveness of the proposed measures in distinguishing different proficiency levels among nonnative Japanese speakers’ writing, we conducted a multi-faceted Rasch measurement analysis (Linacre, 1994 ). This approach applies measurement models to thoroughly analyze various factors that can influence test outcomes, including test takers’ proficiency, item difficulty, and rater severity, among others. The underlying principles and functionality of multi-faceted Rasch measurement are illustrated in (12).

\(\log \left(\frac{{P}_{{nijk}}}{{P}_{{nij}(k-1)}}\right)={B}_{n}-{D}_{i}-{C}_{j}-{F}_{k}\)

(12) defines the logarithmic transformation of the probability ratio ( P nijk /P nij(k-1) )) as a function of multiple parameters. Here, n represents the test taker, i denotes a writing proficiency measure, j corresponds to the human rater, and k represents the proficiency score. The parameter B n signifies the proficiency level of test taker n (where n ranges from 1 to N). D j represents the difficulty parameter of test item i (where i ranges from 1 to L), while C j represents the severity of rater j (where j ranges from 1 to J). Additionally, F k represents the step difficulty for a test taker to move from score ‘k-1’ to k . P nijk refers to the probability of rater j assigning score k to test taker n for test item i . P nij(k-1) represents the likelihood of test taker n being assigned score ‘k-1’ by rater j for test item i . Each facet within the test is treated as an independent parameter and estimated within the same reference framework. To evaluate the consistency of scores obtained through both human and computer analysis, we utilized the Infit mean-square statistic. This statistic is a chi-square measure divided by the degrees of freedom and is weighted with information. It demonstrates higher sensitivity to unexpected patterns in responses to items near a person’s proficiency level (Linacre, 2002 ). Fit statistics are assessed based on predefined thresholds for acceptable fit. For the Infit MNSQ, which has a mean of 1.00, different thresholds have been suggested. Some propose stricter thresholds ranging from 0.7 to 1.3 (Bond et al. 2021 ), while others suggest more lenient thresholds ranging from 0.5 to 1.5 (Eckes, 2009 ). In this study, we adopted the criterion of 0.70–1.30 for the Infit MNSQ.

Moving forward, we can now proceed to assess the effectiveness of the 16 proposed measures based on five criteria for accurately distinguishing various levels of writing proficiency among non-native Japanese speakers. To conduct this evaluation, we utilized the development dataset from the I-JAS corpus, as described in Section Dataset . Table 4 provides a measurement report that presents the performance details of the 14 metrics under consideration. The measure separation was found to be 4.02, indicating a clear differentiation among the measures. The reliability index for the measure separation was 0.891, suggesting consistency in the measurement. Similarly, the person separation reliability index was 0.802, indicating the accuracy of the assessment in distinguishing between individuals. All 16 measures demonstrated Infit mean squares within a reasonable range, ranging from 0.76 to 1.28. The Synonym overlap/paragraph (topic) measure exhibited a relatively high outfit mean square of 1.46, although the Infit mean square falls within an acceptable range. The standard error for the measures ranged from 0.13 to 0.28, indicating the precision of the estimates.

Table 5 further illustrated the weights assigned to different linguistic measures for score prediction, with higher weights indicating stronger correlations between those measures and higher scores. Specifically, the following measures exhibited higher weights compared to others: moving average type token ratio per essay has a weight of 0.0391. Mean dependency distance had a weight of 0.0388. Mean length of clause, calculated by dividing the number of words by the number of clauses, had a weight of 0.0374. Complex nominals per T-unit, calculated by dividing the number of complex nominals by the number of T-units, had a weight of 0.0379. Coordinate phrases rate, calculated by dividing the number of coordinate phrases by the number of clauses, had a weight of 0.0325. Grammatical error rate, representing the number of errors per essay, had a weight of 0.0322.

Criteria (output indicator)

The criteria used to evaluate the writing ability in this study were based on CEFR, which follows a six-point scale ranging from A1 to C2. To assess the quality of Japanese writing, the scoring criteria from Table 6 were utilized. These criteria were derived from the IELTS writing standards and served as assessment guidelines and prompts for the written output.

A prompt is a question or detailed instruction that is provided to the model to obtain a proper response. After several pilot experiments, we decided to provide the measures (Section Measures of writing proficiency for nonnative Japanese ) as the input prompt and use the criteria (Section Criteria (output indicator) ) as the output indicator. Regarding the prompt language, considering that the LLM was tasked with rating Japanese essays, would prompt in Japanese works better Footnote 5 ? We conducted experiments comparing the performance of GPT-4 using both English and Japanese prompts. Additionally, we utilized the Japanese local model OCLL with Japanese prompts. Multiple trials were conducted using the same sample. Regardless of the prompt language used, we consistently obtained the same grading results with GPT-4, which assigned a grade of B1 to the writing sample. This suggested that GPT-4 is reliable and capable of producing consistent ratings regardless of the prompt language. On the other hand, when we used Japanese prompts with the Japanese local model “OCLL”, we encountered inconsistent grading results. Out of 10 attempts with OCLL, only 6 yielded consistent grading results (B1), while the remaining 4 showed different outcomes, including A1 and B2 grades. These findings indicated that the language of the prompt was not the determining factor for reliable AES. Instead, the size of the training data and the model parameters played crucial roles in achieving consistent and reliable AES results for the language model.

The following is the utilized prompt, which details all measures and requires the LLM to score the essays using holistic and trait scores.

Please evaluate Japanese essays written by Japanese learners and assign a score to each essay on a six-point scale, ranging from A1, A2, B1, B2, C1 to C2. Additionally, please provide trait scores and display the calculation process for each trait score. The scoring should be based on the following criteria:

Moving average type-token ratio.

Number of lexical words (token) divided by the total number of words per essay.

Number of sophisticated word types divided by the total number of words per essay.

Mean length of clause.

Verb phrases per T-unit.

Clauses per T-unit.

Dependent clauses per T-unit.

Complex nominals per clause.

Adverbial clauses per clause.

Coordinate phrases per clause.

Mean dependency distance.

Synonym overlap paragraph (topic and keywords).

Word2vec cosine similarity.

Connectives per essay.

Conjunctions per essay.

Number of metadiscourse markers (types) divided by the total number of words.

Number of errors per essay.

Japanese essay text

出かける前に二人が地図を見ている間に、サンドイッチを入れたバスケットに犬が入ってしまいました。それに気づかずに二人は楽しそうに出かけて行きました。やがて突然犬がバスケットから飛び出し、二人は驚きました。バスケット の 中を見ると、食べ物はすべて犬に食べられていて、二人は困ってしまいました。(ID_JJJ01_SW1)

The score of the example above was B1. Figure 3 provides an example of holistic and trait scores provided by GPT-4 (with a prompt indicating all measures) via Bing Footnote 6 .

figure 3

Example of GPT-4 AES and feedback (with a prompt indicating all measures).

Statistical analysis

The aim of this study is to investigate the potential use of LLM for nonnative Japanese AES. It seeks to compare the scoring outcomes obtained from feature-based AES tools, which rely on conventional machine learning technology (i.e. Jess, JWriter), with those generated by AI-driven AES tools utilizing deep learning technology (BERT, GPT, OCLL). To assess the reliability of a computer-assisted annotation tool, the study initially established human-human agreement as the benchmark measure. Subsequently, the performance of the LLM-based method was evaluated by comparing it to human-human agreement.

To assess annotation agreement, the study employed standard measures such as precision, recall, and F-score (Brants 2000 ; Lu 2010 ), along with the quadratically weighted kappa (QWK) to evaluate the consistency and agreement in the annotation process. Assume A and B represent human annotators. When comparing the annotations of the two annotators, the following results are obtained. The evaluation of precision, recall, and F-score metrics was illustrated in equations (13) to (15).

\({\rm{Recall}}(A,B)=\frac{{\rm{Number}}\,{\rm{of}}\,{\rm{identical}}\,{\rm{nodes}}\,{\rm{in}}\,A\,{\rm{and}}\,B}{{\rm{Number}}\,{\rm{of}}\,{\rm{nodes}}\,{\rm{in}}\,A}\)

\({\rm{Precision}}(A,\,B)=\frac{{\rm{Number}}\,{\rm{of}}\,{\rm{identical}}\,{\rm{nodes}}\,{\rm{in}}\,A\,{\rm{and}}\,B}{{\rm{Number}}\,{\rm{of}}\,{\rm{nodes}}\,{\rm{in}}\,B}\)

The F-score is the harmonic mean of recall and precision:

\({\rm{F}}-{\rm{score}}=\frac{2* ({\rm{Precision}}* {\rm{Recall}})}{{\rm{Precision}}+{\rm{Recall}}}\)

The highest possible value of an F-score is 1.0, indicating perfect precision and recall, and the lowest possible value is 0, if either precision or recall are zero.

In accordance with Taghipour and Ng ( 2016 ), the calculation of QWK involves two steps:

Step 1: Construct a weight matrix W as follows:

\({W}_{{ij}}=\frac{{(i-j)}^{2}}{{(N-1)}^{2}}\)

i represents the annotation made by the tool, while j represents the annotation made by a human rater. N denotes the total number of possible annotations. Matrix O is subsequently computed, where O_( i, j ) represents the count of data annotated by the tool ( i ) and the human annotator ( j ). On the other hand, E refers to the expected count matrix, which undergoes normalization to ensure that the sum of elements in E matches the sum of elements in O.

Step 2: With matrices O and E, the QWK is obtained as follows:

K = 1- \(\frac{\sum i,j{W}_{i,j}\,{O}_{i,j}}{\sum i,j{W}_{i,j}\,{E}_{i,j}}\)

The value of the quadratic weighted kappa increases as the level of agreement improves. Further, to assess the accuracy of LLM scoring, the proportional reductive mean square error (PRMSE) was employed. The PRMSE approach takes into account the variability observed in human ratings to estimate the rater error, which is then subtracted from the variance of the human labels. This calculation provides an overall measure of agreement between the automated scores and true scores (Haberman et al. 2015 ; Loukina et al. 2020 ; Taghipour and Ng, 2016 ). The computation of PRMSE involves the following steps:

Step 1: Calculate the mean squared errors (MSEs) for the scoring outcomes of the computer-assisted tool (MSE tool) and the human scoring outcomes (MSE human).

Step 2: Determine the PRMSE by comparing the MSE of the computer-assisted tool (MSE tool) with the MSE from human raters (MSE human), using the following formula:

\({\rm{PRMSE}}=1-\frac{({\rm{MSE}}\,{\rm{tool}})\,}{({\rm{MSE}}\,{\rm{human}})\,}=1-\,\frac{{\sum }_{i}^{n}=1{({{\rm{y}}}_{i}-{\hat{{\rm{y}}}}_{{\rm{i}}})}^{2}}{{\sum }_{i}^{n}=1{({{\rm{y}}}_{i}-\hat{{\rm{y}}})}^{2}}\)

In the numerator, ŷi represents the scoring outcome predicted by a specific LLM-driven AES system for a given sample. The term y i − ŷ i represents the difference between this predicted outcome and the mean value of all LLM-driven AES systems’ scoring outcomes. It quantifies the deviation of the specific LLM-driven AES system’s prediction from the average prediction of all LLM-driven AES systems. In the denominator, y i − ŷ represents the difference between the scoring outcome provided by a specific human rater for a given sample and the mean value of all human raters’ scoring outcomes. It measures the discrepancy between the specific human rater’s score and the average score given by all human raters. The PRMSE is then calculated by subtracting the ratio of the MSE tool to the MSE human from 1. PRMSE falls within the range of 0 to 1, with larger values indicating reduced errors in LLM’s scoring compared to those of human raters. In other words, a higher PRMSE implies that LLM’s scoring demonstrates greater accuracy in predicting the true scores (Loukina et al. 2020 ). The interpretation of kappa values, ranging from 0 to 1, is based on the work of Landis and Koch ( 1977 ). Specifically, the following categories are assigned to different ranges of kappa values: −1 indicates complete inconsistency, 0 indicates random agreement, 0.0 ~ 0.20 indicates extremely low level of agreement (slight), 0.21 ~ 0.40 indicates moderate level of agreement (fair), 0.41 ~ 0.60 indicates medium level of agreement (moderate), 0.61 ~ 0.80 indicates high level of agreement (substantial), 0.81 ~ 1 indicates almost perfect level of agreement. All statistical analyses were executed using Python script.

Results and discussion

Annotation reliability of the llm.

This section focuses on assessing the reliability of the LLM’s annotation and scoring capabilities. To evaluate the reliability, several tests were conducted simultaneously, aiming to achieve the following objectives:

Assess the LLM’s ability to differentiate between test takers with varying levels of oral proficiency.

Determine the level of agreement between the annotations and scoring performed by the LLM and those done by human raters.

The evaluation of the results encompassed several metrics, including: precision, recall, F-Score, quadratically-weighted kappa, proportional reduction of mean squared error, Pearson correlation, and multi-faceted Rasch measurement.

Inter-annotator agreement (human–human annotator agreement)

We started with an agreement test of the two human annotators. Two trained annotators were recruited to determine the writing task data measures. A total of 714 scripts, as the test data, was utilized. Each analysis lasted 300–360 min. Inter-annotator agreement was evaluated using the standard measures of precision, recall, and F-score and QWK. Table 7 presents the inter-annotator agreement for the various indicators. As shown, the inter-annotator agreement was fairly high, with F-scores ranging from 1.0 for sentence and word number to 0.666 for grammatical errors.

The findings from the QWK analysis provided further confirmation of the inter-annotator agreement. The QWK values covered a range from 0.950 ( p  = 0.000) for sentence and word number to 0.695 for synonym overlap number (keyword) and grammatical errors ( p  = 0.001).

Agreement of annotation outcomes between human and LLM

To evaluate the consistency between human annotators and LLM annotators (BERT, GPT, OCLL) across the indices, the same test was conducted. The results of the inter-annotator agreement (F-score) between LLM and human annotation are provided in Appendix B-D. The F-scores ranged from 0.706 for Grammatical error # for OCLL-human to a perfect 1.000 for GPT-human, for sentences, clauses, T-units, and words. These findings were further supported by the QWK analysis, which showed agreement levels ranging from 0.807 ( p  = 0.001) for metadiscourse markers for OCLL-human to 0.962 for words ( p  = 0.000) for GPT-human. The findings demonstrated that the LLM annotation achieved a significant level of accuracy in identifying measurement units and counts.

Reliability of LLM-driven AES’s scoring and discriminating proficiency levels

This section examines the reliability of the LLM-driven AES scoring through a comparison of the scoring outcomes produced by human raters and the LLM ( Reliability of LLM-driven AES scoring ). It also assesses the effectiveness of the LLM-based AES system in differentiating participants with varying proficiency levels ( Reliability of LLM-driven AES discriminating proficiency levels ).

Reliability of LLM-driven AES scoring

Table 8 summarizes the QWK coefficient analysis between the scores computed by the human raters and the GPT-4 for the individual essays from I-JAS Footnote 7 . As shown, the QWK of all measures ranged from k  = 0.819 for lexical density (number of lexical words (tokens)/number of words per essay) to k  = 0.644 for word2vec cosine similarity. Table 9 further presents the Pearson correlations between the 16 writing proficiency measures scored by human raters and GPT 4 for the individual essays. The correlations ranged from 0.672 for syntactic complexity to 0.734 for grammatical accuracy. The correlations between the writing proficiency scores assigned by human raters and the BERT-based AES system were found to range from 0.661 for syntactic complexity to 0.713 for grammatical accuracy. The correlations between the writing proficiency scores given by human raters and the OCLL-based AES system ranged from 0.654 for cohesion to 0.721 for grammatical accuracy. These findings indicated an alignment between the assessments made by human raters and both the BERT-based and OCLL-based AES systems in terms of various aspects of writing proficiency.

Reliability of LLM-driven AES discriminating proficiency levels

After validating the reliability of the LLM’s annotation and scoring, the subsequent objective was to evaluate its ability to distinguish between various proficiency levels. For this analysis, a dataset of 686 individual essays was utilized. Table 10 presents a sample of the results, summarizing the means, standard deviations, and the outcomes of the one-way ANOVAs based on the measures assessed by the GPT-4 model. A post hoc multiple comparison test, specifically the Bonferroni test, was conducted to identify any potential differences between pairs of levels.

As the results reveal, seven measures presented linear upward or downward progress across the three proficiency levels. These were marked in bold in Table 10 and comprise one measure of lexical richness, i.e. MATTR (lexical diversity); four measures of syntactic complexity, i.e. MDD (mean dependency distance), MLC (mean length of clause), CNT (complex nominals per T-unit), CPC (coordinate phrases rate); one cohesion measure, i.e. word2vec cosine similarity and GER (grammatical error rate). Regarding the ability of the sixteen measures to distinguish adjacent proficiency levels, the Bonferroni tests indicated that statistically significant differences exist between the primary level and the intermediate level for MLC and GER. One measure of lexical richness, namely LD, along with three measures of syntactic complexity (VPT, CT, DCT, ACC), two measures of cohesion (SOPT, SOPK), and one measure of content elaboration (IMM), exhibited statistically significant differences between proficiency levels. However, these differences did not demonstrate a linear progression between adjacent proficiency levels. No significant difference was observed in lexical sophistication between proficiency levels.

To summarize, our study aimed to evaluate the reliability and differentiation capabilities of the LLM-driven AES method. For the first objective, we assessed the LLM’s ability to differentiate between test takers with varying levels of oral proficiency using precision, recall, F-Score, and quadratically-weighted kappa. Regarding the second objective, we compared the scoring outcomes generated by human raters and the LLM to determine the level of agreement. We employed quadratically-weighted kappa and Pearson correlations to compare the 16 writing proficiency measures for the individual essays. The results confirmed the feasibility of using the LLM for annotation and scoring in AES for nonnative Japanese. As a result, Research Question 1 has been addressed.

Comparison of BERT-, GPT-, OCLL-based AES, and linguistic-feature-based computation methods

This section aims to compare the effectiveness of five AES methods for nonnative Japanese writing, i.e. LLM-driven approaches utilizing BERT, GPT, and OCLL, linguistic feature-based approaches using Jess and JWriter. The comparison was conducted by comparing the ratings obtained from each approach with human ratings. All ratings were derived from the dataset introduced in Dataset . To facilitate the comparison, the agreement between the automated methods and human ratings was assessed using QWK and PRMSE. The performance of each approach was summarized in Table 11 .

The QWK coefficient values indicate that LLMs (GPT, BERT, OCLL) and human rating outcomes demonstrated higher agreement compared to feature-based AES methods (Jess and JWriter) in assessing writing proficiency criteria, including lexical richness, syntactic complexity, content, and grammatical accuracy. Among the LLMs, the GPT-4 driven AES and human rating outcomes showed the highest agreement in all criteria, except for syntactic complexity. The PRMSE values suggest that the GPT-based method outperformed linguistic feature-based methods and other LLM-based approaches. Moreover, an interesting finding emerged during the study: the agreement coefficient between GPT-4 and human scoring was even higher than the agreement between different human raters themselves. This discovery highlights the advantage of GPT-based AES over human rating. Ratings involve a series of processes, including reading the learners’ writing, evaluating the content and language, and assigning scores. Within this chain of processes, various biases can be introduced, stemming from factors such as rater biases, test design, and rating scales. These biases can impact the consistency and objectivity of human ratings. GPT-based AES may benefit from its ability to apply consistent and objective evaluation criteria. By prompting the GPT model with detailed writing scoring rubrics and linguistic features, potential biases in human ratings can be mitigated. The model follows a predefined set of guidelines and does not possess the same subjective biases that human raters may exhibit. This standardization in the evaluation process contributes to the higher agreement observed between GPT-4 and human scoring. Section Prompt strategy of the study delves further into the role of prompts in the application of LLMs to AES. It explores how the choice and implementation of prompts can impact the performance and reliability of LLM-based AES methods. Furthermore, it is important to acknowledge the strengths of the local model, i.e. the Japanese local model OCLL, which excels in processing certain idiomatic expressions. Nevertheless, our analysis indicated that GPT-4 surpasses local models in AES. This superior performance can be attributed to the larger parameter size of GPT-4, estimated to be between 500 billion and 1 trillion, which exceeds the sizes of both BERT and the local model OCLL.

Prompt strategy

In the context of prompt strategy, Mizumoto and Eguchi ( 2023 ) conducted a study where they applied the GPT-3 model to automatically score English essays in the TOEFL test. They found that the accuracy of the GPT model alone was moderate to fair. However, when they incorporated linguistic measures such as cohesion, syntactic complexity, and lexical features alongside the GPT model, the accuracy significantly improved. This highlights the importance of prompt engineering and providing the model with specific instructions to enhance its performance. In this study, a similar approach was taken to optimize the performance of LLMs. GPT-4, which outperformed BERT and OCLL, was selected as the candidate model. Model 1 was used as the baseline, representing GPT-4 without any additional prompting. Model 2, on the other hand, involved GPT-4 prompted with 16 measures that included scoring criteria, efficient linguistic features for writing assessment, and detailed measurement units and calculation formulas. The remaining models (Models 3 to 18) utilized GPT-4 prompted with individual measures. The performance of these 18 different models was assessed using the output indicators described in Section Criteria (output indicator) . By comparing the performances of these models, the study aimed to understand the impact of prompt engineering on the accuracy and effectiveness of GPT-4 in AES tasks.

  

Model 1: GPT-4

  

  

Model 2: GPT-4 + 17 measures

  

  

Model 3: GPT-4 + MATTR

Model 4: GPT-4 + LD

Model 5: GPT-4 + LS

Model 6: GPT-4 + MLC

Model 7: GPT-4 + VPT

Model 8: GPT-4 + CT

Model 9: GPT-4 + DCT

Model 10: GPT-4 + CNT

Model 11: GPT-4 + ACC

Model 12: GPT-4 + CPC

Model 13: GPT-4 + MDD

Model 14: GPT-4 + SOPT

Model 15: GPT-4 + SOPK

Model 16: GPT-4 + word2vec

 

Model 17: GPT-4 + IMM

Model 18: GPT-4 + GER

 

Based on the PRMSE scores presented in Fig. 4 , it was observed that Model 1, representing GPT-4 without any additional prompting, achieved a fair level of performance. However, Model 2, which utilized GPT-4 prompted with all measures, outperformed all other models in terms of PRMSE score, achieving a score of 0.681. These results indicate that the inclusion of specific measures and prompts significantly enhanced the performance of GPT-4 in AES. Among the measures, syntactic complexity was found to play a particularly significant role in improving the accuracy of GPT-4 in assessing writing quality. Following that, lexical diversity emerged as another important factor contributing to the model’s effectiveness. The study suggests that a well-prompted GPT-4 can serve as a valuable tool to support human assessors in evaluating writing quality. By utilizing GPT-4 as an automated scoring tool, the evaluation biases associated with human raters can be minimized. This has the potential to empower teachers by allowing them to focus on designing writing tasks and guiding writing strategies, while leveraging the capabilities of GPT-4 for efficient and reliable scoring.

figure 4

PRMSE scores of the 18 AES models.

This study aimed to investigate two main research questions: the feasibility of utilizing LLMs for AES and the impact of prompt engineering on the application of LLMs in AES.

To address the first objective, the study compared the effectiveness of five different models: GPT, BERT, the Japanese local LLM (OCLL), and two conventional machine learning-based AES tools (Jess and JWriter). The PRMSE values indicated that the GPT-4-based method outperformed other LLMs (BERT, OCLL) and linguistic feature-based computational methods (Jess and JWriter) across various writing proficiency criteria. Furthermore, the agreement coefficient between GPT-4 and human scoring surpassed the agreement among human raters themselves, highlighting the potential of using the GPT-4 tool to enhance AES by reducing biases and subjectivity, saving time, labor, and cost, and providing valuable feedback for self-study. Regarding the second goal, the role of prompt design was investigated by comparing 18 models, including a baseline model, a model prompted with all measures, and 16 models prompted with one measure at a time. GPT-4, which outperformed BERT and OCLL, was selected as the candidate model. The PRMSE scores of the models showed that GPT-4 prompted with all measures achieved the best performance, surpassing the baseline and other models.

In conclusion, this study has demonstrated the potential of LLMs in supporting human rating in assessments. By incorporating automation, we can save time and resources while reducing biases and subjectivity inherent in human rating processes. Automated language assessments offer the advantage of accessibility, providing equal opportunities and economic feasibility for individuals who lack access to traditional assessment centers or necessary resources. LLM-based language assessments provide valuable feedback and support to learners, aiding in the enhancement of their language proficiency and the achievement of their goals. This personalized feedback can cater to individual learner needs, facilitating a more tailored and effective language-learning experience.

There are three important areas that merit further exploration. First, prompt engineering requires attention to ensure optimal performance of LLM-based AES across different language types. This study revealed that GPT-4, when prompted with all measures, outperformed models prompted with fewer measures. Therefore, investigating and refining prompt strategies can enhance the effectiveness of LLMs in automated language assessments. Second, it is crucial to explore the application of LLMs in second-language assessment and learning for oral proficiency, as well as their potential in under-resourced languages. Recent advancements in self-supervised machine learning techniques have significantly improved automatic speech recognition (ASR) systems, opening up new possibilities for creating reliable ASR systems, particularly for under-resourced languages with limited data. However, challenges persist in the field of ASR. First, ASR assumes correct word pronunciation for automatic pronunciation evaluation, which proves challenging for learners in the early stages of language acquisition due to diverse accents influenced by their native languages. Accurately segmenting short words becomes problematic in such cases. Second, developing precise audio-text transcriptions for languages with non-native accented speech poses a formidable task. Last, assessing oral proficiency levels involves capturing various linguistic features, including fluency, pronunciation, accuracy, and complexity, which are not easily captured by current NLP technology.

Data availability

The dataset utilized was obtained from the International Corpus of Japanese as a Second Language (I-JAS). The data URLs: [ https://www2.ninjal.ac.jp/jll/lsaj/ihome2.html ].

J-CAT and TTBJ are two computerized adaptive tests used to assess Japanese language proficiency.

SPOT is a specific component of the TTBJ test.

J-CAT: https://www.j-cat2.org/html/ja/pages/interpret.html

SPOT: https://ttbj.cegloc.tsukuba.ac.jp/p1.html#SPOT .

The study utilized a prompt-based GPT-4 model, developed by OpenAI, which has an impressive architecture with 1.8 trillion parameters across 120 layers. GPT-4 was trained on a vast dataset of 13 trillion tokens, using two stages: initial training on internet text datasets to predict the next token, and subsequent fine-tuning through reinforcement learning from human feedback.

https://www2.ninjal.ac.jp/jll/lsaj/ihome2-en.html .

http://jhlee.sakura.ne.jp/JEV/ by Japanese Learning Dictionary Support Group 2015.

We express our sincere gratitude to the reviewer for bringing this matter to our attention.

On February 7, 2023, Microsoft began rolling out a major overhaul to Bing that included a new chatbot feature based on OpenAI’s GPT-4 (Bing.com).

Appendix E-F present the analysis results of the QWK coefficient between the scores computed by the human raters and the BERT, OCLL models.

Attali Y, Burstein J (2006) Automated essay scoring with e-rater® V.2. J. Technol., Learn. Assess., 4

Barkaoui K, Hadidi A (2020) Assessing Change in English Second Language Writing Performance (1st ed.). Routledge, New York. https://doi.org/10.4324/9781003092346

Bentz C, Tatyana R, Koplenig A, Tanja S (2016) A comparison between morphological complexity. measures: Typological data vs. language corpora. In Proceedings of the workshop on computational linguistics for linguistic complexity (CL4LC), 142–153. Osaka, Japan: The COLING 2016 Organizing Committee

Bond TG, Yan Z, Heene M (2021) Applying the Rasch model: Fundamental measurement in the human sciences (4th ed). Routledge

Brants T (2000) Inter-annotator agreement for a German newspaper corpus. Proceedings of the Second International Conference on Language Resources and Evaluation (LREC’00), Athens, Greece, 31 May-2 June, European Language Resources Association

Brown TB, Mann B, Ryder N, et al. (2020) Language models are few-shot learners. Advances in Neural Information Processing Systems, Online, 6–12 December, Curran Associates, Inc., Red Hook, NY

Burstein J (2003) The E-rater scoring engine: Automated essay scoring with natural language processing. In Shermis MD and Burstein JC (ed) Automated Essay Scoring: A Cross-Disciplinary Perspective. Lawrence Erlbaum Associates, Mahwah, NJ

Čech R, Miroslav K (2018) Morphological richness of text. In Masako F, Václav C (ed) Taming the corpus: From inflection and lexis to interpretation, 63–77. Cham, Switzerland: Springer Nature

Çöltekin Ç, Taraka, R (2018) Exploiting Universal Dependencies treebanks for measuring morphosyntactic complexity. In Aleksandrs B, Christian B (ed), Proceedings of first workshop on measuring language complexity, 1–7. Torun, Poland

Crossley SA, Cobb T, McNamara DS (2013) Comparing count-based and band-based indices of word frequency: Implications for active vocabulary research and pedagogical applications. System 41:965–981. https://doi.org/10.1016/j.system.2013.08.002

Article   Google Scholar  

Crossley SA, McNamara DS (2016) Say more and be more coherent: How text elaboration and cohesion can increase writing quality. J. Writ. Res. 7:351–370

CyberAgent Inc (2023) Open-Calm series of Japanese language models. Retrieved from: https://www.cyberagent.co.jp/news/detail/id=28817

Devlin J, Chang MW, Lee K, Toutanova K (2019) BERT: Pre-training of deep bidirectional transformers for language understanding. Proceedings of the 2019 Conference of the North American Chapter of the Association for Computational Linguistics, Minneapolis, Minnesota, 2–7 June, pp. 4171–4186. Association for Computational Linguistics

Diez-Ortega M, Kyle K (2023) Measuring the development of lexical richness of L2 Spanish: a longitudinal learner corpus study. Studies in Second Language Acquisition 1-31

Eckes T (2009) On common ground? How raters perceive scoring criteria in oral proficiency testing. In Brown A, Hill K (ed) Language testing and evaluation 13: Tasks and criteria in performance assessment (pp. 43–73). Peter Lang Publishing

Elliot S (2003) IntelliMetric: from here to validity. In: Shermis MD, Burstein JC (ed) Automated Essay Scoring: A Cross-Disciplinary Perspective. Lawrence Erlbaum Associates, Mahwah, NJ

Google Scholar  

Engber CA (1995) The relationship of lexical proficiency to the quality of ESL compositions. J. Second Lang. Writ. 4:139–155

Garner J, Crossley SA, Kyle K (2019) N-gram measures and L2 writing proficiency. System 80:176–187. https://doi.org/10.1016/j.system.2018.12.001

Haberman SJ (2008) When can subscores have value? J. Educat. Behav. Stat., 33:204–229

Haberman SJ, Yao L, Sinharay S (2015) Prediction of true test scores from observed item scores and ancillary data. Brit. J. Math. Stat. Psychol. 68:363–385

Halliday MAK (1985) Spoken and Written Language. Deakin University Press, Melbourne, Australia

Hirao R, Arai M, Shimanaka H et al. (2020) Automated essay scoring system for nonnative Japanese learners. Proceedings of the 12th Conference on Language Resources and Evaluation (LREC 2020), pp. 1250–1257. European Language Resources Association

Hunt KW (1966) Recent Measures in Syntactic Development. Elementary English, 43(7), 732–739. http://www.jstor.org/stable/41386067

Ishioka T (2001) About e-rater, a computer-based automatic scoring system for essays [Konpyūta ni yoru essei no jidō saiten shisutemu e − rater ni tsuite]. University Entrance Examination. Forum [Daigaku nyūshi fōramu] 24:71–76

Hochreiter S, Schmidhuber J (1997) Long short- term memory. Neural Comput. 9(8):1735–1780

Article   CAS   PubMed   Google Scholar  

Ishioka T, Kameda M (2006) Automated Japanese essay scoring system based on articles written by experts. Proceedings of the 21st International Conference on Computational Linguistics and 44th Annual Meeting of the Association for Computational Linguistics, Sydney, Australia, 17–18 July 2006, pp. 233-240. Association for Computational Linguistics, USA

Japan Foundation (2021) Retrieved from: https://www.jpf.gp.jp/j/project/japanese/survey/result/dl/survey2021/all.pdf

Jarvis S (2013a) Defining and measuring lexical diversity. In Jarvis S, Daller M (ed) Vocabulary knowledge: Human ratings and automated measures (Vol. 47, pp. 13–44). John Benjamins. https://doi.org/10.1075/sibil.47.03ch1

Jarvis S (2013b) Capturing the diversity in lexical diversity. Lang. Learn. 63:87–106. https://doi.org/10.1111/j.1467-9922.2012.00739.x

Jiang J, Quyang J, Liu H (2019) Interlanguage: A perspective of quantitative linguistic typology. Lang. Sci. 74:85–97

Kim M, Crossley SA, Kyle K (2018) Lexical sophistication as a multidimensional phenomenon: Relations to second language lexical proficiency, development, and writing quality. Mod. Lang. J. 102(1):120–141. https://doi.org/10.1111/modl.12447

Kojima T, Gu S, Reid M et al. (2022) Large language models are zero-shot reasoners. Advances in Neural Information Processing Systems, New Orleans, LA, 29 November-1 December, Curran Associates, Inc., Red Hook, NY

Kyle K, Crossley SA (2015) Automatically assessing lexical sophistication: Indices, tools, findings, and application. TESOL Q 49:757–786

Kyle K, Crossley SA, Berger CM (2018) The tool for the automatic analysis of lexical sophistication (TAALES): Version 2.0. Behav. Res. Methods 50:1030–1046. https://doi.org/10.3758/s13428-017-0924-4

Article   PubMed   Google Scholar  

Kyle K, Crossley SA, Jarvis S (2021) Assessing the validity of lexical diversity using direct judgements. Lang. Assess. Q. 18:154–170. https://doi.org/10.1080/15434303.2020.1844205

Landauer TK, Laham D, Foltz PW (2003) Automated essay scoring and annotation of essays with the Intelligent Essay Assessor. In Shermis MD, Burstein JC (ed), Automated Essay Scoring: A Cross-Disciplinary Perspective. Lawrence Erlbaum Associates, Mahwah, NJ

Landis JR, Koch GG (1977) The measurement of observer agreement for categorical data. Biometrics 159–174

Laufer B, Nation P (1995) Vocabulary size and use: Lexical richness in L2 written production. Appl. Linguist. 16:307–322. https://doi.org/10.1093/applin/16.3.307

Lee J, Hasebe Y (2017) jWriter Learner Text Evaluator, URL: https://jreadability.net/jwriter/

Lee J, Kobayashi N, Sakai T, Sakota K (2015) A Comparison of SPOT and J-CAT Based on Test Analysis [Tesuto bunseki ni motozuku ‘SPOT’ to ‘J-CAT’ no hikaku]. Research on the Acquisition of Second Language Japanese [Dainigengo to shite no nihongo no shūtoku kenkyū] (18) 53–69

Li W, Yan J (2021) Probability distribution of dependency distance based on a Treebank of. Japanese EFL Learners’ Interlanguage. J. Quant. Linguist. 28(2):172–186. https://doi.org/10.1080/09296174.2020.1754611

Article   MathSciNet   Google Scholar  

Linacre JM (2002) Optimizing rating scale category effectiveness. J. Appl. Meas. 3(1):85–106

PubMed   Google Scholar  

Linacre JM (1994) Constructing measurement with a Many-Facet Rasch Model. In Wilson M (ed) Objective measurement: Theory into practice, Volume 2 (pp. 129–144). Norwood, NJ: Ablex

Liu H (2008) Dependency distance as a metric of language comprehension difficulty. J. Cognitive Sci. 9:159–191

Liu H, Xu C, Liang J (2017) Dependency distance: A new perspective on syntactic patterns in natural languages. Phys. Life Rev. 21. https://doi.org/10.1016/j.plrev.2017.03.002

Loukina A, Madnani N, Cahill A, et al. (2020) Using PRMSE to evaluate automated scoring systems in the presence of label noise. Proceedings of the Fifteenth Workshop on Innovative Use of NLP for Building Educational Applications, Seattle, WA, USA → Online, 10 July, pp. 18–29. Association for Computational Linguistics

Lu X (2010) Automatic analysis of syntactic complexity in second language writing. Int. J. Corpus Linguist. 15:474–496

Lu X (2012) The relationship of lexical richness to the quality of ESL learners’ oral narratives. Mod. Lang. J. 96:190–208

Lu X (2017) Automated measurement of syntactic complexity in corpus-based L2 writing research and implications for writing assessment. Lang. Test. 34:493–511

Lu X, Hu R (2022) Sense-aware lexical sophistication indices and their relationship to second language writing quality. Behav. Res. Method. 54:1444–1460. https://doi.org/10.3758/s13428-021-01675-6

Ministry of Health, Labor, and Welfare of Japan (2022) Retrieved from: https://www.mhlw.go.jp/stf/newpage_30367.html

Mizumoto A, Eguchi M (2023) Exploring the potential of using an AI language model for automated essay scoring. Res. Methods Appl. Linguist. 3:100050

Okgetheng B, Takeuchi K (2024) Estimating Japanese Essay Grading Scores with Large Language Models. Proceedings of 30th Annual Conference of the Language Processing Society in Japan, March 2024

Ortega L (2015) Second language learning explained? SLA across 10 contemporary theories. In VanPatten B, Williams J (ed) Theories in Second Language Acquisition: An Introduction

Rae JW, Borgeaud S, Cai T, et al. (2021) Scaling Language Models: Methods, Analysis & Insights from Training Gopher. ArXiv, abs/2112.11446

Read J (2000) Assessing vocabulary. Cambridge University Press. https://doi.org/10.1017/CBO9780511732942

Rudner LM, Liang T (2002) Automated Essay Scoring Using Bayes’ Theorem. J. Technol., Learning and Assessment, 1 (2)

Sakoda K, Hosoi Y (2020) Accuracy and complexity of Japanese Language usage by SLA learners in different learning environments based on the analysis of I-JAS, a learners’ corpus of Japanese as L2. Math. Linguist. 32(7):403–418. https://doi.org/10.24701/mathling.32.7_403

Suzuki N (1999) Summary of survey results regarding comprehensive essay questions. Final report of “Joint Research on Comprehensive Examinations for the Aim of Evaluating Applicability to Each Specialized Field of Universities” for 1996-2000 [shōronbun sōgō mondai ni kansuru chōsa kekka no gaiyō. Heisei 8 - Heisei 12-nendo daigaku no kaku senmon bun’ya e no tekisei no hyōka o mokuteki to suru sōgō shiken no arikata ni kansuru kyōdō kenkyū’ saishū hōkoku-sho]. University Entrance Examination Section Center Research and Development Department [Daigaku nyūshi sentā kenkyū kaihatsubu], 21–32

Taghipour K, Ng HT (2016) A neural approach to automated essay scoring. Proceedings of the 2016 Conference on Empirical Methods in Natural Language Processing, Austin, Texas, 1–5 November, pp. 1882–1891. Association for Computational Linguistics

Takeuchi K, Ohno M, Motojin K, Taguchi M, Inada Y, Iizuka M, Abo T, Ueda H (2021) Development of essay scoring methods based on reference texts with construction of research-available Japanese essay data. In IPSJ J 62(9):1586–1604

Ure J (1971) Lexical density: A computational technique and some findings. In Coultard M (ed) Talking about Text. English Language Research, University of Birmingham, Birmingham, England

Vaswani A, Shazeer N, Parmar N, et al. (2017) Attention is all you need. In Advances in Neural Information Processing Systems, Long Beach, CA, 4–7 December, pp. 5998–6008, Curran Associates, Inc., Red Hook, NY

Watanabe H, Taira Y, Inoue Y (1988) Analysis of essay evaluation data [Shōronbun hyōka dēta no kaiseki]. Bulletin of the Faculty of Education, University of Tokyo [Tōkyōdaigaku kyōiku gakubu kiyō], Vol. 28, 143–164

Yao S, Yu D, Zhao J, et al. (2023) Tree of thoughts: Deliberate problem solving with large language models. Advances in Neural Information Processing Systems, 36

Zenker F, Kyle K (2021) Investigating minimum text lengths for lexical diversity indices. Assess. Writ. 47:100505. https://doi.org/10.1016/j.asw.2020.100505

Zhang Y, Warstadt A, Li X, et al. (2021) When do you need billions of words of pretraining data? Proceedings of the 59th Annual Meeting of the Association for Computational Linguistics and the 11th International Joint Conference on Natural Language Processing, Online, pp. 1112-1125. Association for Computational Linguistics. https://doi.org/10.18653/v1/2021.acl-long.90

Download references

This research was funded by National Foundation of Social Sciences (22BYY186) to Wenchao Li.

Author information

Authors and affiliations.

Department of Japanese Studies, Zhejiang University, Hangzhou, China

Department of Linguistics and Applied Linguistics, Zhejiang University, Hangzhou, China

You can also search for this author in PubMed   Google Scholar

Contributions

Wenchao Li is in charge of conceptualization, validation, formal analysis, investigation, data curation, visualization and writing the draft. Haitao Liu is in charge of supervision.

Corresponding author

Correspondence to Wenchao Li .

Ethics declarations

Competing interests.

The authors declare no competing interests.

Ethical approval

Ethical approval was not required as the study did not involve human participants.

Informed consent

This article does not contain any studies with human participants performed by any of the authors.

Additional information

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

Supplementary information

Supplemental material file #1, rights and permissions.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/ .

Reprints and permissions

About this article

Cite this article.

Li, W., Liu, H. Applying large language models for automated essay scoring for non-native Japanese. Humanit Soc Sci Commun 11 , 723 (2024). https://doi.org/10.1057/s41599-024-03209-9

Download citation

Received : 02 February 2024

Accepted : 16 May 2024

Published : 03 June 2024

DOI : https://doi.org/10.1057/s41599-024-03209-9

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

automated essay scoring dataset

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings

Preview improvements coming to the PMC website in October 2024. Learn More or Try it out now .

  • Advanced Search
  • Journal List
  • Springer Nature - PMC COVID-19 Collection

Logo of phenaturepg

An automated essay scoring systems: a systematic literature review

Dadi ramesh.

1 School of Computer Science and Artificial Intelligence, SR University, Warangal, TS India

2 Research Scholar, JNTU, Hyderabad, India

Suresh Kumar Sanampudi

3 Department of Information Technology, JNTUH College of Engineering, Nachupally, Kondagattu, Jagtial, TS India

Associated Data

Assessment in the Education system plays a significant role in judging student performance. The present evaluation system is through human assessment. As the number of teachers' student ratio is gradually increasing, the manual evaluation process becomes complicated. The drawback of manual evaluation is that it is time-consuming, lacks reliability, and many more. This connection online examination system evolved as an alternative tool for pen and paper-based methods. Present Computer-based evaluation system works only for multiple-choice questions, but there is no proper evaluation system for grading essays and short answers. Many researchers are working on automated essay grading and short answer scoring for the last few decades, but assessing an essay by considering all parameters like the relevance of the content to the prompt, development of ideas, Cohesion, and Coherence is a big challenge till now. Few researchers focused on Content-based evaluation, while many of them addressed style-based assessment. This paper provides a systematic literature review on automated essay scoring systems. We studied the Artificial Intelligence and Machine Learning techniques used to evaluate automatic essay scoring and analyzed the limitations of the current studies and research trends. We observed that the essay evaluation is not done based on the relevance of the content and coherence.

Supplementary Information

The online version contains supplementary material available at 10.1007/s10462-021-10068-2.

Introduction

Due to COVID 19 outbreak, an online educational system has become inevitable. In the present scenario, almost all the educational institutions ranging from schools to colleges adapt the online education system. The assessment plays a significant role in measuring the learning ability of the student. Most automated evaluation is available for multiple-choice questions, but assessing short and essay answers remain a challenge. The education system is changing its shift to online-mode, like conducting computer-based exams and automatic evaluation. It is a crucial application related to the education domain, which uses natural language processing (NLP) and Machine Learning techniques. The evaluation of essays is impossible with simple programming languages and simple techniques like pattern matching and language processing. Here the problem is for a single question, we will get more responses from students with a different explanation. So, we need to evaluate all the answers concerning the question.

Automated essay scoring (AES) is a computer-based assessment system that automatically scores or grades the student responses by considering appropriate features. The AES research started in 1966 with the Project Essay Grader (PEG) by Ajay et al. ( 1973 ). PEG evaluates the writing characteristics such as grammar, diction, construction, etc., to grade the essay. A modified version of the PEG by Shermis et al. ( 2001 ) was released, which focuses on grammar checking with a correlation between human evaluators and the system. Foltz et al. ( 1999 ) introduced an Intelligent Essay Assessor (IEA) by evaluating content using latent semantic analysis to produce an overall score. Powers et al. ( 2002 ) proposed E-rater and Intellimetric by Rudner et al. ( 2006 ) and Bayesian Essay Test Scoring System (BESTY) by Rudner and Liang ( 2002 ), these systems use natural language processing (NLP) techniques that focus on style and content to obtain the score of an essay. The vast majority of the essay scoring systems in the 1990s followed traditional approaches like pattern matching and a statistical-based approach. Since the last decade, the essay grading systems started using regression-based and natural language processing techniques. AES systems like Dong et al. ( 2017 ) and others developed from 2014 used deep learning techniques, inducing syntactic and semantic features resulting in better results than earlier systems.

Ohio, Utah, and most US states are using AES systems in school education, like Utah compose tool, Ohio standardized test (an updated version of PEG), evaluating millions of student's responses every year. These systems work for both formative, summative assessments and give feedback to students on the essay. Utah provided basic essay evaluation rubrics (six characteristics of essay writing): Development of ideas, organization, style, word choice, sentence fluency, conventions. Educational Testing Service (ETS) has been conducting significant research on AES for more than a decade and designed an algorithm to evaluate essays on different domains and providing an opportunity for test-takers to improve their writing skills. In addition, they are current research content-based evaluation.

The evaluation of essay and short answer scoring should consider the relevance of the content to the prompt, development of ideas, Cohesion, Coherence, and domain knowledge. Proper assessment of the parameters mentioned above defines the accuracy of the evaluation system. But all these parameters cannot play an equal role in essay scoring and short answer scoring. In a short answer evaluation, domain knowledge is required, like the meaning of "cell" in physics and biology is different. And while evaluating essays, the implementation of ideas with respect to prompt is required. The system should also assess the completeness of the responses and provide feedback.

Several studies examined AES systems, from the initial to the latest AES systems. In which the following studies on AES systems are Blood ( 2011 ) provided a literature review from PEG 1984–2010. Which has covered only generalized parts of AES systems like ethical aspects, the performance of the systems. Still, they have not covered the implementation part, and it’s not a comparative study and has not discussed the actual challenges of AES systems.

Burrows et al. ( 2015 ) Reviewed AES systems on six dimensions like dataset, NLP techniques, model building, grading models, evaluation, and effectiveness of the model. They have not covered feature extraction techniques and challenges in features extractions. Covered only Machine Learning models but not in detail. This system not covered the comparative analysis of AES systems like feature extraction, model building, and level of relevance, cohesion, and coherence not covered in this review.

Ke et al. ( 2019 ) provided a state of the art of AES system but covered very few papers and not listed all challenges, and no comparative study of the AES model. On the other hand, Hussein et al. in ( 2019 ) studied two categories of AES systems, four papers from handcrafted features for AES systems, and four papers from the neural networks approach, discussed few challenges, and did not cover feature extraction techniques, the performance of AES models in detail.

Klebanov et al. ( 2020 ). Reviewed 50 years of AES systems, listed and categorized all essential features that need to be extracted from essays. But not provided a comparative analysis of all work and not discussed the challenges.

This paper aims to provide a systematic literature review (SLR) on automated essay grading systems. An SLR is an Evidence-based systematic review to summarize the existing research. It critically evaluates and integrates all relevant studies' findings and addresses the research domain's specific research questions. Our research methodology uses guidelines given by Kitchenham et al. ( 2009 ) for conducting the review process; provide a well-defined approach to identify gaps in current research and to suggest further investigation.

We addressed our research method, research questions, and the selection process in Sect.  2 , and the results of the research questions have discussed in Sect.  3 . And the synthesis of all the research questions addressed in Sect.  4 . Conclusion and possible future work discussed in Sect.  5 .

Research method

We framed the research questions with PICOC criteria.

Population (P) Student essays and answers evaluation systems.

Intervention (I) evaluation techniques, data sets, features extraction methods.

Comparison (C) Comparison of various approaches and results.

Outcomes (O) Estimate the accuracy of AES systems,

Context (C) NA.

Research questions

To collect and provide research evidence from the available studies in the domain of automated essay grading, we framed the following research questions (RQ):

RQ1 what are the datasets available for research on automated essay grading?

The answer to the question can provide a list of the available datasets, their domain, and access to the datasets. It also provides a number of essays and corresponding prompts.

RQ2 what are the features extracted for the assessment of essays?

The answer to the question can provide an insight into various features so far extracted, and the libraries used to extract those features.

RQ3, which are the evaluation metrics available for measuring the accuracy of algorithms?

The answer will provide different evaluation metrics for accurate measurement of each Machine Learning approach and commonly used measurement technique.

RQ4 What are the Machine Learning techniques used for automatic essay grading, and how are they implemented?

It can provide insights into various Machine Learning techniques like regression models, classification models, and neural networks for implementing essay grading systems. The response to the question can give us different assessment approaches for automated essay grading systems.

RQ5 What are the challenges/limitations in the current research?

The answer to the question provides limitations of existing research approaches like cohesion, coherence, completeness, and feedback.

Search process

We conducted an automated search on well-known computer science repositories like ACL, ACM, IEEE Explore, Springer, and Science Direct for an SLR. We referred to papers published from 2010 to 2020 as much of the work during these years focused on advanced technologies like deep learning and natural language processing for automated essay grading systems. Also, the availability of free data sets like Kaggle (2012), Cambridge Learner Corpus-First Certificate in English exam (CLC-FCE) by Yannakoudakis et al. ( 2011 ) led to research this domain.

Search Strings : We used search strings like “Automated essay grading” OR “Automated essay scoring” OR “short answer scoring systems” OR “essay scoring systems” OR “automatic essay evaluation” and searched on metadata.

Selection criteria

After collecting all relevant documents from the repositories, we prepared selection criteria for inclusion and exclusion of documents. With the inclusion and exclusion criteria, it becomes more feasible for the research to be accurate and specific.

Inclusion criteria 1 Our approach is to work with datasets comprise of essays written in English. We excluded the essays written in other languages.

Inclusion criteria 2  We included the papers implemented on the AI approach and excluded the traditional methods for the review.

Inclusion criteria 3 The study is on essay scoring systems, so we exclusively included the research carried out on only text data sets rather than other datasets like image or speech.

Exclusion criteria  We removed the papers in the form of review papers, survey papers, and state of the art papers.

Quality assessment

In addition to the inclusion and exclusion criteria, we assessed each paper by quality assessment questions to ensure the article's quality. We included the documents that have clearly explained the approach they used, the result analysis and validation.

The quality checklist questions are framed based on the guidelines from Kitchenham et al. ( 2009 ). Each quality assessment question was graded as either 1 or 0. The final score of the study range from 0 to 3. A cut off score for excluding a study from the review is 2 points. Since the papers scored 2 or 3 points are included in the final evaluation. We framed the following quality assessment questions for the final study.

Quality Assessment 1: Internal validity.

Quality Assessment 2: External validity.

Quality Assessment 3: Bias.

The two reviewers review each paper to select the final list of documents. We used the Quadratic Weighted Kappa score to measure the final agreement between the two reviewers. The average resulted from the kappa score is 0.6942, a substantial agreement between the reviewers. The result of evolution criteria shown in Table ​ Table1. 1 . After Quality Assessment, the final list of papers for review is shown in Table ​ Table2. 2 . The complete selection process is shown in Fig. ​ Fig.1. 1 . The total number of selected papers in year wise as shown in Fig. ​ Fig.2. 2 .

Quality assessment analysis

Number of papersQuality assessment score
503
122
591
230

Final list of papers

Data basePaper count
ACL28
ACM5
IEEE Explore19
Springer5
Other5
Total62

An external file that holds a picture, illustration, etc.
Object name is 10462_2021_10068_Fig1_HTML.jpg

Selection process

An external file that holds a picture, illustration, etc.
Object name is 10462_2021_10068_Fig2_HTML.jpg

Year wise publications

What are the datasets available for research on automated essay grading?

To work with problem statement especially in Machine Learning and deep learning domain, we require considerable amount of data to train the models. To answer this question, we listed all the data sets used for training and testing for automated essay grading systems. The Cambridge Learner Corpus-First Certificate in English exam (CLC-FCE) Yannakoudakis et al. ( 2011 ) developed corpora that contain 1244 essays and ten prompts. This corpus evaluates whether a student can write the relevant English sentences without any grammatical and spelling mistakes. This type of corpus helps to test the models built for GRE and TOFEL type of exams. It gives scores between 1 and 40.

Bailey and Meurers ( 2008 ), Created a dataset (CREE reading comprehension) for language learners and automated short answer scoring systems. The corpus consists of 566 responses from intermediate students. Mohler and Mihalcea ( 2009 ). Created a dataset for the computer science domain consists of 630 responses for data structure assignment questions. The scores are range from 0 to 5 given by two human raters.

Dzikovska et al. ( 2012 ) created a Student Response Analysis (SRA) corpus. It consists of two sub-groups: the BEETLE corpus consists of 56 questions and approximately 3000 responses from students in the electrical and electronics domain. The second one is the SCIENTSBANK(SemEval-2013) (Dzikovska et al. 2013a ; b ) corpus consists of 10,000 responses on 197 prompts on various science domains. The student responses ladled with "correct, partially correct incomplete, Contradictory, Irrelevant, Non-domain."

In the Kaggle (2012) competition, released total 3 types of corpuses on an Automated Student Assessment Prize (ASAP1) (“ https://www.kaggle.com/c/asap-sas/ ” ) essays and short answers. It has nearly 17,450 essays, out of which it provides up to 3000 essays for each prompt. It has eight prompts that test 7th to 10th grade US students. It gives scores between the [0–3] and [0–60] range. The limitations of these corpora are: (1) it has a different score range for other prompts. (2) It uses statistical features such as named entities extraction and lexical features of words to evaluate essays. ASAP +  + is one more dataset from Kaggle. It is with six prompts, and each prompt has more than 1000 responses total of 10,696 from 8th-grade students. Another corpus contains ten prompts from science, English domains and a total of 17,207 responses. Two human graders evaluated all these responses.

Correnti et al. ( 2013 ) created a Response-to-Text Assessment (RTA) dataset used to check student writing skills in all directions like style, mechanism, and organization. 4–8 grade students give the responses to RTA. Basu et al. ( 2013 ) created a power grading dataset with 700 responses for ten different prompts from US immigration exams. It contains all short answers for assessment.

The TOEFL11 corpus Blanchard et al. ( 2013 ) contains 1100 essays evenly distributed over eight prompts. It is used to test the English language skills of a candidate attending the TOFEL exam. It scores the language proficiency of a candidate as low, medium, and high.

International Corpus of Learner English (ICLE) Granger et al. ( 2009 ) built a corpus of 3663 essays covering different dimensions. It has 12 prompts with 1003 essays that test the organizational skill of essay writing, and13 prompts, each with 830 essays that examine the thesis clarity and prompt adherence.

Argument Annotated Essays (AAE) Stab and Gurevych ( 2014 ) developed a corpus that contains 102 essays with 101 prompts taken from the essayforum2 site. It tests the persuasive nature of the student essay. The SCIENTSBANK corpus used by Sakaguchi et al. ( 2015 ) available in git-hub, containing 9804 answers to 197 questions in 15 science domains. Table ​ Table3 3 illustrates all datasets related to AES systems.

ALL types Datasets used in Automatic scoring systems

Data SetLanguageTotal responsesNumber of prompts
Cambridge Learner Corpus-First Certificate in English exam (CLC-FCE)English1244
CREEEnglish566
CSEnglish630
SRAEnglish300056
SCIENTSBANK(SemEval-2013)English10,000197
ASAP-AESEnglish17,4508
ASAP-SASEnglish17,20710
ASAP + + English10,6966
power gradingEnglish700
TOEFL11English11008
International Corpus of Learner English (ICLE)English3663

Features play a major role in the neural network and other supervised Machine Learning approaches. The automatic essay grading systems scores student essays based on different types of features, which play a prominent role in training the models. Based on their syntax and semantics and they are categorized into three groups. 1. statistical-based features Contreras et al. ( 2018 ); Kumar et al. ( 2019 ); Mathias and Bhattacharyya ( 2018a ; b ) 2. Style-based (Syntax) features Cummins et al. ( 2016 ); Darwish and Mohamed ( 2020 ); Ke et al. ( 2019 ). 3. Content-based features Dong et al. ( 2017 ). A good set of features appropriate models evolved better AES systems. The vast majority of the researchers are using regression models if features are statistical-based. For Neural Networks models, researches are using both style-based and content-based features. The following table shows the list of various features used in existing AES Systems. Table ​ Table4 4 represents all set of features used for essay grading.

Types of features

Statistical featuresStyle based featuresContent based features
Essay length with respect to the number of wordsSentence structureCohesion between sentences in a document
Essay length with respect to sentencePOSOverlapping (prompt)
Average sentence lengthPunctuationRelevance of information
Average word lengthGrammaticalSemantic role of words
N-gramLogical operatorsCorrectness
VocabularyConsistency
Sentence expressing key concepts

We studied all the feature extracting NLP libraries as shown in Fig. ​ Fig.3. that 3 . that are used in the papers. The NLTK is an NLP tool used to retrieve statistical features like POS, word count, sentence count, etc. With NLTK, we can miss the essay's semantic features. To find semantic features Word2Vec Mikolov et al. ( 2013 ), GloVe Jeffrey Pennington et al. ( 2014 ) is the most used libraries to retrieve the semantic text from the essays. And in some systems, they directly trained the model with word embeddings to find the score. From Fig. ​ Fig.4 4 as observed that non-content-based feature extraction is higher than content-based.

An external file that holds a picture, illustration, etc.
Object name is 10462_2021_10068_Fig3_HTML.jpg

Usages of tools

An external file that holds a picture, illustration, etc.
Object name is 10462_2021_10068_Fig4_HTML.jpg

Number of papers on content based features

RQ3 which are the evaluation metrics available for measuring the accuracy of algorithms?

The majority of the AES systems are using three evaluation metrics. They are (1) quadrated weighted kappa (QWK) (2) Mean Absolute Error (MAE) (3) Pearson Correlation Coefficient (PCC) Shehab et al. ( 2016 ). The quadratic weighted kappa will find agreement between human evaluation score and system evaluation score and produces value ranging from 0 to 1. And the Mean Absolute Error is the actual difference between human-rated score to system-generated score. The mean square error (MSE) measures the average squares of the errors, i.e., the average squared difference between the human-rated and the system-generated scores. MSE will always give positive numbers only. Pearson's Correlation Coefficient (PCC) finds the correlation coefficient between two variables. It will provide three values (0, 1, − 1). "0" represents human-rated and system scores that are not related. "1" represents an increase in the two scores. "− 1" illustrates a negative relationship between the two scores.

RQ4 what are the Machine Learning techniques being used for automatic essay grading, and how are they implemented?

After scrutinizing all documents, we categorize the techniques used in automated essay grading systems into four baskets. 1. Regression techniques. 2. Classification model. 3. Neural networks. 4. Ontology-based approach.

All the existing AES systems developed in the last ten years employ supervised learning techniques. Researchers using supervised methods viewed the AES system as either regression or classification task. The goal of the regression task is to predict the score of an essay. The classification task is to classify the essays belonging to (low, medium, or highly) relevant to the question's topic. Since the last three years, most AES systems developed made use of the concept of the neural network.

Regression based models

Mohler and Mihalcea ( 2009 ). proposed text-to-text semantic similarity to assign a score to the student essays. There are two text similarity measures like Knowledge-based measures, corpus-based measures. There eight knowledge-based tests with all eight models. They found the similarity. The shortest path similarity determines based on the length, which shortest path between two contexts. Leacock & Chodorow find the similarity based on the shortest path's length between two concepts using node-counting. The Lesk similarity finds the overlap between the corresponding definitions, and Wu & Palmer algorithm finds similarities based on the depth of two given concepts in the wordnet taxonomy. Resnik, Lin, Jiang&Conrath, Hirst& St-Onge find the similarity based on different parameters like the concept, probability, normalization factor, lexical chains. In corpus-based likeness, there LSA BNC, LSA Wikipedia, and ESA Wikipedia, latent semantic analysis is trained on Wikipedia and has excellent domain knowledge. Among all similarity scores, correlation scores LSA Wikipedia scoring accuracy is more. But these similarity measure algorithms are not using NLP concepts. These models are before 2010 and basic concept models to continue the research automated essay grading with updated algorithms on neural networks with content-based features.

Adamson et al. ( 2014 ) proposed an automatic essay grading system which is a statistical-based approach in this they retrieved features like POS, Character count, Word count, Sentence count, Miss spelled words, n-gram representation of words to prepare essay vector. They formed a matrix with these all vectors in that they applied LSA to give a score to each essay. It is a statistical approach that doesn’t consider the semantics of the essay. The accuracy they got when compared to the human rater score with the system is 0.532.

Cummins et al. ( 2016 ). Proposed Timed Aggregate Perceptron vector model to give ranking to all the essays, and later they converted the rank algorithm to predict the score of the essay. The model trained with features like Word unigrams, bigrams, POS, Essay length, grammatical relation, Max word length, sentence length. It is multi-task learning, gives ranking to the essays, and predicts the score for the essay. The performance evaluated through QWK is 0.69, a substantial agreement between the human rater and the system.

Sultan et al. ( 2016 ). Proposed a Ridge regression model to find short answer scoring with Question Demoting. Question Demoting is the new concept included in the essay's final assessment to eliminate duplicate words from the essay. The extracted features are Text Similarity, which is the similarity between the student response and reference answer. Question Demoting is the number of repeats in a student response. With inverse document frequency, they assigned term weight. The sentence length Ratio is the number of words in the student response, is another feature. With these features, the Ridge regression model was used, and the accuracy they got 0.887.

Contreras et al. ( 2018 ). Proposed Ontology based on text mining in this model has given a score for essays in phases. In phase-I, they generated ontologies with ontoGen and SVM to find the concept and similarity in the essay. In phase II from ontologies, they retrieved features like essay length, word counts, correctness, vocabulary, and types of word used, domain information. After retrieving statistical data, they used a linear regression model to find the score of the essay. The accuracy score is the average of 0.5.

Darwish and Mohamed ( 2020 ) proposed the fusion of fuzzy Ontology with LSA. They retrieve two types of features, like syntax features and semantic features. In syntax features, they found Lexical Analysis with tokens, and they construct a parse tree. If the parse tree is broken, the essay is inconsistent—a separate grade assigned to the essay concerning syntax features. The semantic features are like similarity analysis, Spatial Data Analysis. Similarity analysis is to find duplicate sentences—Spatial Data Analysis for finding Euclid distance between the center and part. Later they combine syntax features and morphological features score for the final score. The accuracy they achieved with the multiple linear regression model is 0.77, mostly on statistical features.

Süzen Neslihan et al. ( 2020 ) proposed a text mining approach for short answer grading. First, their comparing model answers with student response by calculating the distance between two sentences. By comparing the model answer with student response, they find the essay's completeness and provide feedback. In this approach, model vocabulary plays a vital role in grading, and with this model vocabulary, the grade will be assigned to the student's response and provides feedback. The correlation between the student answer to model answer is 0.81.

Classification based Models

Persing and Ng ( 2013 ) used a support vector machine to score the essay. The features extracted are OS, N-gram, and semantic text to train the model and identified the keywords from the essay to give the final score.

Sakaguchi et al. ( 2015 ) proposed two methods: response-based and reference-based. In response-based scoring, the extracted features are response length, n-gram model, and syntactic elements to train the support vector regression model. In reference-based scoring, features such as sentence similarity using word2vec is used to find the cosine similarity of the sentences that is the final score of the response. First, the scores were discovered individually and later combined two features to find a final score. This system gave a remarkable increase in performance by combining the scores.

Mathias and Bhattacharyya ( 2018a ; b ) Proposed Automated Essay Grading Dataset with Essay Attribute Scores. The first concept features selection depends on the essay type. So the common attributes are Content, Organization, Word Choice, Sentence Fluency, Conventions. In this system, each attribute is scored individually, with the strength of each attribute identified. The model they used is a random forest classifier to assign scores to individual attributes. The accuracy they got with QWK is 0.74 for prompt 1 of the ASAS dataset ( https://www.kaggle.com/c/asap-sas/ ).

Ke et al. ( 2019 ) used a support vector machine to find the response score. In this method, features like Agreeability, Specificity, Clarity, Relevance to prompt, Conciseness, Eloquence, Confidence, Direction of development, Justification of opinion, and Justification of importance. First, the individual parameter score obtained was later combined with all scores to give a final response score. The features are used in the neural network to find whether the sentence is relevant to the topic or not.

Salim et al. ( 2019 ) proposed an XGBoost Machine Learning classifier to assess the essays. The algorithm trained on features like word count, POS, parse tree depth, and coherence in the articles with sentence similarity percentage; cohesion and coherence are considered for training. And they implemented K-fold cross-validation for a result the average accuracy after specific validations is 68.12.

Neural network models

Shehab et al. ( 2016 ) proposed a neural network method that used learning vector quantization to train human scored essays. After training, the network can provide a score to the ungraded essays. First, we should process the essay to remove Spell checking and then perform preprocessing steps like Document Tokenization, stop word removal, Stemming, and submit it to the neural network. Finally, the model will provide feedback on the essay, whether it is relevant to the topic. And the correlation coefficient between human rater and system score is 0.7665.

Kopparapu and De ( 2016 ) proposed the Automatic Ranking of Essays using Structural and Semantic Features. This approach constructed a super essay with all the responses. Next, ranking for a student essay is done based on the super-essay. The structural and semantic features derived helps to obtain the scores. In a paragraph, 15 Structural features like an average number of sentences, the average length of sentences, and the count of words, nouns, verbs, adjectives, etc., are used to obtain a syntactic score. A similarity score is used as semantic features to calculate the overall score.

Dong and Zhang ( 2016 ) proposed a hierarchical CNN model. The model builds two layers with word embedding to represents the words as the first layer. The second layer is a word convolution layer with max-pooling to find word vectors. The next layer is a sentence-level convolution layer with max-pooling to find the sentence's content and synonyms. A fully connected dense layer produces an output score for an essay. The accuracy with the hierarchical CNN model resulted in an average QWK of 0.754.

Taghipour and Ng ( 2016 ) proposed a first neural approach for essay scoring build in which convolution and recurrent neural network concepts help in scoring an essay. The network uses a lookup table with the one-hot representation of the word vector of an essay. The final efficiency of the network model with LSTM resulted in an average QWK of 0.708.

Dong et al. ( 2017 ). Proposed an Attention-based scoring system with CNN + LSTM to score an essay. For CNN, the input parameters were character embedding and word embedding, and it has attention pooling layers and used NLTK to obtain word and character embedding. The output gives a sentence vector, which provides sentence weight. After CNN, it will have an LSTM layer with an attention pooling layer, and this final layer results in the final score of the responses. The average QWK score is 0.764.

Riordan et al. ( 2017 ) proposed a neural network with CNN and LSTM layers. Word embedding, given as input to a neural network. An LSTM network layer will retrieve the window features and delivers them to the aggregation layer. The aggregation layer is a superficial layer that takes a correct window of words and gives successive layers to predict the answer's sore. The accuracy of the neural network resulted in a QWK of 0.90.

Zhao et al. ( 2017 ) proposed a new concept called Memory-Augmented Neural network with four layers, input representation layer, memory addressing layer, memory reading layer, and output layer. An input layer represents all essays in a vector form based on essay length. After converting the word vector, the memory addressing layer takes a sample of the essay and weighs all the terms. The memory reading layer takes the input from memory addressing segment and finds the content to finalize the score. Finally, the output layer will provide the final score of the essay. The accuracy of essay scores is 0.78, which is far better than the LSTM neural network.

Mathias and Bhattacharyya ( 2018a ; b ) proposed deep learning networks using LSTM with the CNN layer and GloVe pre-trained word embeddings. For this, they retrieved features like Sentence count essays, word count per sentence, Number of OOVs in the sentence, Language model score, and the text's perplexity. The network predicted the goodness scores of each essay. The higher the goodness scores, means higher the rank and vice versa.

Nguyen and Dery ( 2016 ). Proposed Neural Networks for Automated Essay Grading. In this method, a single layer bi-directional LSTM accepting word vector as input. Glove vectors used in this method resulted in an accuracy of 90%.

Ruseti et al. ( 2018 ) proposed a recurrent neural network that is capable of memorizing the text and generate a summary of an essay. The Bi-GRU network with the max-pooling layer molded on the word embedding of each document. It will provide scoring to the essay by comparing it with a summary of the essay from another Bi-GRU network. The result obtained an accuracy of 0.55.

Wang et al. ( 2018a ; b ) proposed an automatic scoring system with the bi-LSTM recurrent neural network model and retrieved the features using the word2vec technique. This method generated word embeddings from the essay words using the skip-gram model. And later, word embedding is used to train the neural network to find the final score. The softmax layer in LSTM obtains the importance of each word. This method used a QWK score of 0.83%.

Dasgupta et al. ( 2018 ) proposed a technique for essay scoring with augmenting textual qualitative Features. It extracted three types of linguistic, cognitive, and psychological features associated with a text document. The linguistic features are Part of Speech (POS), Universal Dependency relations, Structural Well-formedness, Lexical Diversity, Sentence Cohesion, Causality, and Informativeness of the text. The psychological features derived from the Linguistic Information and Word Count (LIWC) tool. They implemented a convolution recurrent neural network that takes input as word embedding and sentence vector, retrieved from the GloVe word vector. And the second layer is the Convolution Layer to find local features. The next layer is the recurrent neural network (LSTM) to find corresponding of the text. The accuracy of this method resulted in an average QWK of 0.764.

Liang et al. ( 2018 ) proposed a symmetrical neural network AES model with Bi-LSTM. They are extracting features from sample essays and student essays and preparing an embedding layer as input. The embedding layer output is transfer to the convolution layer from that LSTM will be trained. Hear the LSRM model has self-features extraction layer, which will find the essay's coherence. The average QWK score of SBLSTMA is 0.801.

Liu et al. ( 2019 ) proposed two-stage learning. In the first stage, they are assigning a score based on semantic data from the essay. The second stage scoring is based on some handcrafted features like grammar correction, essay length, number of sentences, etc. The average score of the two stages is 0.709.

Pedro Uria Rodriguez et al. ( 2019 ) proposed a sequence-to-sequence learning model for automatic essay scoring. They used BERT (Bidirectional Encoder Representations from Transformers), which extracts the semantics from a sentence from both directions. And XLnet sequence to sequence learning model to extract features like the next sentence in an essay. With this pre-trained model, they attained coherence from the essay to give the final score. The average QWK score of the model is 75.5.

Xia et al. ( 2019 ) proposed a two-layer Bi-directional LSTM neural network for the scoring of essays. The features extracted with word2vec to train the LSTM and accuracy of the model in an average of QWK is 0.870.

Kumar et al. ( 2019 ) Proposed an AutoSAS for short answer scoring. It used pre-trained Word2Vec and Doc2Vec models trained on Google News corpus and Wikipedia dump, respectively, to retrieve the features. First, they tagged every word POS and they found weighted words from the response. It also found prompt overlap to observe how the answer is relevant to the topic, and they defined lexical overlaps like noun overlap, argument overlap, and content overlap. This method used some statistical features like word frequency, difficulty, diversity, number of unique words in each response, type-token ratio, statistics of the sentence, word length, and logical operator-based features. This method uses a random forest model to train the dataset. The data set has sample responses with their associated score. The model will retrieve the features from both responses like graded and ungraded short answers with questions. The accuracy of AutoSAS with QWK is 0.78. It will work on any topics like Science, Arts, Biology, and English.

Jiaqi Lun et al. ( 2020 ) proposed an automatic short answer scoring with BERT. In this with a reference answer comparing student responses and assigning scores. The data augmentation is done with a neural network and with one correct answer from the dataset classifying reaming responses as correct or incorrect.

Zhu and Sun ( 2020 ) proposed a multimodal Machine Learning approach for automated essay scoring. First, they count the grammar score with the spaCy library and numerical count as the number of words and sentences with the same library. With this input, they trained a single and Bi LSTM neural network for finding the final score. For the LSTM model, they prepared sentence vectors with GloVe and word embedding with NLTK. Bi-LSTM will check each sentence in both directions to find semantic from the essay. The average QWK score with multiple models is 0.70.

Ontology based approach

Mohler et al. ( 2011 ) proposed a graph-based method to find semantic similarity in short answer scoring. For the ranking of answers, they used the support vector regression model. The bag of words is the main feature extracted in the system.

Ramachandran et al. ( 2015 ) also proposed a graph-based approach to find lexical based semantics. Identified phrase patterns and text patterns are the features to train a random forest regression model to score the essays. The accuracy of the model in a QWK is 0.78.

Zupanc et al. ( 2017 ) proposed sentence similarity networks to find the essay's score. Ajetunmobi and Daramola ( 2017 ) recommended an ontology-based information extraction approach and domain-based ontology to find the score.

Speech response scoring

Automatic scoring is in two ways one is text-based scoring, other is speech-based scoring. This paper discussed text-based scoring and its challenges, and now we cover speech scoring and common points between text and speech-based scoring. Evanini and Wang ( 2013 ), Worked on speech scoring of non-native school students, extracted features with speech ratter, and trained a linear regression model, concluding that accuracy varies based on voice pitching. Loukina et al. ( 2015 ) worked on feature selection from speech data and trained SVM. Malinin et al. ( 2016 ) used neural network models to train the data. Loukina et al. ( 2017 ). Proposed speech and text-based automatic scoring. Extracted text-based features, speech-based features and trained a deep neural network for speech-based scoring. They extracted 33 types of features based on acoustic signals. Malinin et al. ( 2017 ). Wu Xixin et al. ( 2020 ) Worked on deep neural networks for spoken language assessment. Incorporated different types of models and tested them. Ramanarayanan et al. ( 2017 ) worked on feature extraction methods and extracted punctuation, fluency, and stress and trained different Machine Learning models for scoring. Knill et al. ( 2018 ). Worked on Automatic speech recognizer and its errors how its impacts the speech assessment.

The state of the art

This section provides an overview of the existing AES systems with a comparative study w. r. t models, features applied, datasets, and evaluation metrics used for building the automated essay grading systems. We divided all 62 papers into two sets of the first set of review papers in Table ​ Table5 5 with a comparative study of the AES systems.

State of the art

SystemApproachDatasetFeatures appliedEvaluation metric and results
Mohler and Mihalcea in ( )shortest path similarity, LSA regression modelWord vectorFinds the shortest path
Niraj Kumar and Lipika Dey. In ( )Word-GraphASAP KaggleContent and style-based features63.81% accuracy
Alex Adamson et al. in ( )LSA regression modelASAP KaggleStatistical featuresQWK 0.532
Nguyen and Dery ( )LSTM (single layer bidirectional)ASAP KaggleStatistical features90% accuracy
Keisuke Sakaguchi et al. in ( )Classification modelETS (educational testing services)Statistical, Style based featuresQWK is 0.69
Ramachandran et al. in ( )regression modelASAP Kaggle short AnswerStatistical and style-based featuresQWK 0.77
Sultan et al. in ( )Ridge regression modelSciEntBank answersStatistical featuresRMSE 0.887
Dong and Zhang ( )CNN neural networkASAP KaggleStatistical featuresQWK 0.734
Taghipour and Ngl in ( )CNN + LSTM neural networkASAP KaggleLookup table (one hot representation of word vector)QWK 0.761
Shehab et al. in ( )Learning vector quantization neural network

Mansoura University

student's essays

Statistical featurescorrelation coefficient 0.7665
Cummins et al. in ( )Regression modelASAP KaggleStatistical features, style-based featuresQWK 0.69
Kopparapu and De ( )Neural networkASAP KaggleStatistical features, Style based
Dong, et al. in ( )CNN + LSTM neural networkASAP KaggleWord embedding, content basedQWK 0.764
Ajetunmobi and Daramola ( )WuPalmer algorithmStatistical features
Siyuan Zhao et al. in ( )LSTM (memory network)ASAP KaggleStatistical featuresQWK 0.78
Mathias and Bhattacharyya ( )

Random Forest

Classifier a classification model

ASAP KaggleStyle and Content based featuresClassified which feature set is required
Brian Riordan et al. in ( )CNN + LSTM neural networkASAP Kaggle short AnswerWord embeddingsQWK 0.90
Tirthankar Dasgupta et al. in ( )

CNN

-bidirectional

LSTMs neural network

ASAP KaggleContent and physiological featuresQWK 0.786
Wu and Shih ( )Classification modelSciEntBank answersunigram_recall

Squared correlation coefficient

59.568

unigram_precision
unigram_F_measure
log_bleu_recall
log_bleu_precision

log_bleu_F_measure

BLUE features

Yucheng Wang, etc.in ( )Bi-LSTMASAP KaggleWord embedding sequenceQWK 0.724
Anak Agung Putri Ratna et al. in ( )Winnowing ALGORITHM86.86 accuracy
Sharma and Jayagopi ( )Glove, LSTM neural networkASAP KaggleHand written essay imagesQWK 0.69
Jennifer O. Contreras et al. in ( )

OntoGen (SVM)

Linear Regression

University of Benghazi data setStatistical, style-based features
Mathias, Bhattacharyya ( )GloVe,LSTM neural networkASAP KaggleStatistical features, style featuresPredicted Goodness score for essay
Stefan Ruseti, et al. in ( )BiGRU Siamese architectureAmazon Mechanical Turk online research service. Collected summariesWord embeddingAccuracy 55.2
Zining wang, et al. in ( )

LSTM (semantic)

HAN (hierarchical attention network) neural network

ASAP KaggleWord embeddingQWK 0.83
Guoxi Liang et al. ( )Bi-LSTMASAP KaggleWord embedding, coherence of sentenceQWK 0.801
Ke et al. in ( )Classification modelASAP KaggleContent based

Pearson’s

Correlation Coefficient (PC)-0.39

ME-0.921

Tsegaye Misikir Tashu and Horváth in ( )Unsupervised learning–Locality Sensitivity HashingASAP KaggleStatistical featuresroot mean squared error
Kumar and Dey ( )

Random Forest

CNN, RNN neural network

ASAP Kaggle short AnswerStyle and content-based featuresQWK 0.82
Pedro Uria Rodriguez et al. ( )BERT, XlnetASAP KaggleError correction, sequence learningQWK 0.755
Jiawei Liu et al. ( )CNN, LSTM, BERTASAP Kagglesemantic data, handcrafted features like grammar correction, essay length, number of sentences, etcQWK 0.709
Darwish and Mohamed ( )Multiple Linear RegressionASAP KaggleStyle and content-based featuresQWK 0.77
Jiaqi Lun et al. ( )BERTSemEval-2013Student Answer, Reference AnswerAccuracy 0.8277 (2-way)
Süzen, Neslihan, et al. ( )Text miningintroductory computer science class in the University of North Texas, Student AssignmentsSentence similarityCorrelation score 0.81
Wilson Zhu and Yu Sun in ( )RNN (LSTM, Bi-LSTM)ASAP KaggleWord embedding, grammar count, word countQWK 0.70
Salim Yafet et al. ( )XGBoost machine learning classifierASAP KaggleWord count, POS, parse tree, coherence, cohesion, type token rationAccuracy 68.12
Andrzej Cader ( )Deep Neural NetworkUniversity of Social Sciences in Lodz students’ answersasynchronous featureAccuracy 0.99
Tashu TM, Horváth T ( )

Rule based algorithm,

Similarity based algorithm

ASAP KaggleSimilarity basedAccuracy 0.68
Masaki Uto(B) and Masashi Okano ( )Item Response Theory Models (CNN-LSTM,BERT)ASAP KaggleQWK 0.749

Comparison of all approaches

In our study, we divided major AES approaches into three categories. Regression models, classification models, and neural network models. The regression models failed to find cohesion and coherence from the essay because it trained on BoW(Bag of Words) features. In processing data from input to output, the regression models are less complicated than neural networks. There are unable to find many intricate patterns from the essay and unable to find sentence connectivity. If we train the model with BoW features in the neural network approach, the model never considers the essay's coherence and coherence.

First, to train a Machine Learning algorithm with essays, all the essays are converted to vector form. We can form a vector with BoW and Word2vec, TF-IDF. The BoW and Word2vec vector representation of essays represented in Table ​ Table6. 6 . The vector representation of BoW with TF-IDF is not incorporating the essays semantic, and it’s just statistical learning from a given vector. Word2vec vector comprises semantic of essay in a unidirectional way.

Vector representation of essays

EssayBoW << vector >> Word2vec << vector >> 
Student 1 responseI believe that using computers will benefit us in many ways like talking and becoming friends will others through websites like facebook and mysace << 0.00000 0.00000 0.165746 0.280633 … 0.00000 0.280633 0.280633 0.280633 >> 

 << 3.9792988e-03 − 1.9810481e-03 1.9830784e-03 9.0381579e-04 − 2.9438005e-03 2.1778699e-03 4.4950014e-03 2.9508960e -03 − 2.2331756e-03 − 3.8774475e-03 3.5967759e- 03 − 4.0194849e-03 − 3.0412588e-03 − 2.4055617e-03 4.8296354e-03 2.4813593e-03…

− 2.7158875e-03 − 1.4563646e-03 1.4072991e-03 − 5.2228488e-04

− 2.3597316e-03 6.2979700e-04 − 3.0249553e-03 4.4125126e-04

2.1633594e-03 − 4.9487003e-03 9.9755758e-05 − 2.4388896e-03 >> 

Student 2 responseMore and more people use computers, but not everyone agrees that this benefits society. Those who support advances in technology believe that computers have a positive effect on people << 0.26043 0.26043 0.153814 0.000000 … 0.26043 0.000000 0.000000 0.000000 >  > 

 << 3.9792988e-03 − 1.9810481e- 03 1.9830784e-03 9.0381579e-04

− 2.9438005e-03 2.1778699e-03 4.4950014e-03 2.9508960e-03

− 2.2331756e-03 − 3.8774475e-03 3.5967759e-03 − 4.0194849e-03…

− 2.7158875e-03 − 1.4563646e-03 1.4072991e-03 − 5.2228488e-04

− 2.3597316e-03 6.2979700e-04 − 3.0249553e-03 4.4125126e-04

3.7868773e-03 − 4.4193151e-03 3.0735810e-03 2.5546195e-03

2.1633594e-03 − 4.9487003e-03 9.9755758e-05 − 2.4388896e-03 >> 

In BoW, the vector contains the frequency of word occurrences in the essay. The vector represents 1 and more based on the happenings of words in the essay and 0 for not present. So, in BoW, the vector does not maintain the relationship with adjacent words; it’s just for single words. In word2vec, the vector represents the relationship between words with other words and sentences prompt in multiple dimensional ways. But word2vec prepares vectors in a unidirectional way, not in a bidirectional way; word2vec fails to find semantic vectors when a word has two meanings, and the meaning depends on adjacent words. Table ​ Table7 7 represents a comparison of Machine Learning models and features extracting methods.

Comparison of models

BoWWord2vec
Regression models/classification modelsThe system implemented with Bow features and regression or classification algorithms will have low cohesion and coherenceThe system implemented with Word2vec features and regression or classification algorithms will have low to medium cohesion and coherence
Neural Networks (LSTM)The system implemented with BoW features and neural network models will have low cohesion and coherenceThe system implemented with Word2vec features and neural network model (LSTM) will have medium to high cohesion and coherence

In AES, cohesion and coherence will check the content of the essay concerning the essay prompt these can be extracted from essay in the vector from. Two more parameters are there to access an essay is completeness and feedback. Completeness will check whether student’s response is sufficient or not though the student wrote correctly. Table ​ Table8 8 represents all four parameters comparison for essay grading. Table ​ Table9 9 illustrates comparison of all approaches based on various features like grammar, spelling, organization of essay, relevance.

Comparison of all models with respect to cohesion, coherence, completeness, feedback

AuthorsCohesionCoherenceCompletenessFeed Back
Mohler and Mihalcea ( )LowLowLowLow
Mohler et al. ( )MediumLowMediumLow
Persing and Ng ( )MediumLowLowLow
Adamson et al. ( )LowLowLowLow
Ramachandran et al. ( )MediumMediumLowLow
Sakaguchi et al.. ( ),MediumLowLowLow
Cummins et al. ( )LowLowLowLow
Sultan et al. ( )MediumMediumLowLow
Shehab et al. ( )LowLowLowLow
Kopparapu and De ( )MediumMediumLowLow
Dong an Zhang ( )MediumLowLowLow
Taghipour and Ng ( )MediumMediumLowLow
Zupanc et al. ( )MediumMediumLowLow
Dong et al. ( )MediumMediumLowLow
Riordan et al. ( )MediumMediumMediumLow
Zhao et al. ( )MediumMediumLowLow
Contreras et al. ( )MediumLowLowLow
Mathias and Bhattacharyya ( ; )MediumMediumLowLow
Mathias and Bhattacharyya ( ; )MediumMediumLowLow
Nguyen and Dery ( )MediumMediumMediumMedium
Ruseti et al. ( )MediumLowLowLow
Dasgupta et al. ( )MediumMediumLowLow
Liu et al.( )LowLowLowLow
Wang et al. ( )MediumLowLowLow
Guoxi Liang et al. ( )HighHighLowLow
Wang et al. ( )MediumMediumLowLow
Chen and Li ( )MediumMediumLowLow
Li et al. ( )MediumMediumLowLow
Alva-Manchego et al.( )LowLowLowLow
Jiawei Liu et al. ( )HighHighMediumLow
Pedro Uria Rodriguez et al. ( )MediumMediumMediumLow
Changzhi Cai( )LowLowLowLow
Xia et al. ( )MediumMediumLowLow
Chen and Zhou ( )LowLowLowLow
Kumar et al. ( )MediumMediumMediumLow
Ke et al. ( )MediumLowMediumLow
Andrzej Cader( )LowLowLowLow
Jiaqi Lun et al. ( )HighHighLowLow
Wilson Zhu and Yu Sun ( )MediumMediumLowLow
Süzen, Neslihan et al. ( )MediumLowMediumLow
Salim Yafet et al. ( ) HighMediumLowLow
Darwish and Mohamed ( )MediumLowLowLow
Tashu and Horváth ( )MediumMediumLowMedium
Tashu ( )MediumMediumLowLow
Masaki Uto(B) and Masashi Okano( )MediumMediumMediumMedium
Panitan Muangkammuen and Fumiyo Fukumoto( )MediumMediumMediumLow

comparison of all approaches on various features

ApproachesGrammarStyle (Word choice, sentence structure)Mechanics (Spelling, punctuation, capitalization)DevelopmentBoW (tf-idf)relevance
Mohler and Mihalcea ( )NoNoNoNoYesNo
Mohler et al. ( )YesNoNoNoYesNo
Persing and Ng ( )YesYesYesNoYesYes
Adamson et al. ( )YesNoYesNoYesNo
Ramachandran et al. ( )YesNoYesYesYesYes
Sakaguchi et al. ( ),NoNoYesYesYesYes
Cummins et al. ( )YesNoYesNoYesNo
Sultan et al. ( )NoNoNoNoYesYes
Shehab et al. ( )YesYesYesNoYesNo
Kopparapu and De ( )NoNoNoNoYesNo
Dong and Zhang ( )YesNoYesNoYesYes
Taghipour and Ng ( )YesNoNoNoYesYes
Zupanc et al. ( )NoNoNoNoYesNo
Dong et al. ( )NoNoNoNoNoYes
Riordan et al. ( )NoNoNoNoNoYes
Zhao et al. ( )NoNoNoNoNoYes
Contreras et al. ( )YesNoNoNoYesYes
Mathias and Bhattacharyya ( , )NoYesYesNoNoYes
Mathias and Bhattacharyya ( , )YesNoYesNoYesYes
Nguyen and Dery ( )NoNoNoNoYesYes
Ruseti et al. ( )NoNoNoYesNoYes
Dasgupta et al. ( )YesYesYesYesNoYes
Liu et al.( )YesYesNoNoYesNo
Wang et al. ( )NoNoNoNoNoYes
Guoxi Liang et al. ( )NoNoNoNoNoYes
Wang et al. ( )NoNoNoNoNoYes
Chen and Li ( )NoNoNoNoNoYes
Li et al. ( )YesNoNoNoNoYes
Alva-Manchego et al. ( )YesNoNoYesNoYes
Jiawei Liu et al. ( )YesNoNoYesNoYes
Pedro Uria Rodriguez et al. ( )NoNoNoNoYesYes
Changzhi Cai( )NoNoNoNoNoYes
Xia et al. ( )NoNoNoNoNoYes
Chen and Zhou ( )NoNoNoNoNoYes
Kumar et al. ( )YesYesNoYesYesYes
Ke et al. ( )NoYesNoYesYesYes
Andrzej Cader( )NoNoNoNoNoYes
Jiaqi Lun et al. ( )NoNoNoNoNoYes
Wilson Zhu and Yu Sun ( )NoNoNoNoNoYes
Süzen, Neslihan, et al. ( ) NoNoNoNoYesYes
Salim Yafet et al. ( )YesYesYesNoYesYes
Darwish and Mohamed ( )YesYesNoNoNoYes

What are the challenges/limitations in the current research?

From our study and results discussed in the previous sections, many researchers worked on automated essay scoring systems with numerous techniques. We have statistical methods, classification methods, and neural network approaches to evaluate the essay automatically. The main goal of the automated essay grading system is to reduce human effort and improve consistency.

The vast majority of essay scoring systems are dealing with the efficiency of the algorithm. But there are many challenges in automated essay grading systems. One should assess the essay by following parameters like the relevance of the content to the prompt, development of ideas, Cohesion, Coherence, and domain knowledge.

No model works on the relevance of content, which means whether student response or explanation is relevant to the given prompt or not if it is relevant to how much it is appropriate, and there is no discussion about the cohesion and coherence of the essays. All researches concentrated on extracting the features using some NLP libraries, trained their models, and testing the results. But there is no explanation in the essay evaluation system about consistency and completeness, But Palma and Atkinson ( 2018 ) explained coherence-based essay evaluation. And Zupanc and Bosnic ( 2014 ) also used the word coherence to evaluate essays. And they found consistency with latent semantic analysis (LSA) for finding coherence from essays, but the dictionary meaning of coherence is "The quality of being logical and consistent."

Another limitation is there is no domain knowledge-based evaluation of essays using Machine Learning models. For example, the meaning of a cell is different from biology to physics. Many Machine Learning models extract features with WordVec and GloVec; these NLP libraries cannot convert the words into vectors when they have two or more meanings.

Other challenges that influence the Automated Essay Scoring Systems.

All these approaches worked to improve the QWK score of their models. But QWK will not assess the model in terms of features extraction and constructed irrelevant answers. The QWK is not evaluating models whether the model is correctly assessing the answer or not. There are many challenges concerning students' responses to the Automatic scoring system. Like in evaluating approach, no model has examined how to evaluate the constructed irrelevant and adversarial answers. Especially the black box type of approaches like deep learning models provides more options to the students to bluff the automated scoring systems.

The Machine Learning models that work on statistical features are very vulnerable. Based on Powers et al. ( 2001 ) and Bejar Isaac et al. ( 2014 ), the E-rater was failed on Constructed Irrelevant Responses Strategy (CIRS). From the study of Bejar et al. ( 2013 ), Higgins and Heilman ( 2014 ), observed that when student response contain irrelevant content or shell language concurring to prompt will influence the final score of essays in an automated scoring system.

In deep learning approaches, most of the models automatically read the essay's features, and some methods work on word-based embedding and other character-based embedding features. From the study of Riordan Brain et al. ( 2019 ), The character-based embedding systems do not prioritize spelling correction. However, it is influencing the final score of the essay. From the study of Horbach and Zesch ( 2019 ), Various factors are influencing AES systems. For example, there are data set size, prompt type, answer length, training set, and human scorers for content-based scoring.

Ding et al. ( 2020 ) reviewed that the automated scoring system is vulnerable when a student response contains more words from prompt, like prompt vocabulary repeated in the response. Parekh et al. ( 2020 ) and Kumar et al. ( 2020 ) tested various neural network models of AES by iteratively adding important words, deleting unimportant words, shuffle the words, and repeating sentences in an essay and found that no change in the final score of essays. These neural network models failed to recognize common sense in adversaries' essays and give more options for the students to bluff the automated systems.

Other than NLP and ML techniques for AES. From Wresch ( 1993 ) to Madnani and Cahill ( 2018 ). discussed the complexity of AES systems, standards need to be followed. Like assessment rubrics to test subject knowledge, irrelevant responses, and ethical aspects of an algorithm like measuring the fairness of student response.

Fairness is an essential factor for automated systems. For example, in AES, fairness can be measure in an agreement between human score to machine score. Besides this, From Loukina et al. ( 2019 ), the fairness standards include overall score accuracy, overall score differences, and condition score differences between human and system scores. In addition, scoring different responses in the prospect of constructive relevant and irrelevant will improve fairness.

Madnani et al. ( 2017a ; b ). Discussed the fairness of AES systems for constructed responses and presented RMS open-source tool for detecting biases in the models. With this, one can change fairness standards according to their analysis of fairness.

From Berzak et al.'s ( 2018 ) approach, behavior factors are a significant challenge in automated scoring systems. That helps to find language proficiency, word characteristics (essential words from the text), predict the critical patterns from the text, find related sentences in an essay, and give a more accurate score.

Rupp ( 2018 ), has discussed the designing, evaluating, and deployment methodologies for AES systems. They provided notable characteristics of AES systems for deployment. They are like model performance, evaluation metrics for a model, threshold values, dynamically updated models, and framework.

First, we should check the model performance on different datasets and parameters for operational deployment. Selecting Evaluation metrics for AES models are like QWK, correlation coefficient, or sometimes both. Kelley and Preacher ( 2012 ) have discussed three categories of threshold values: marginal, borderline, and acceptable. The values can be varied based on data size, model performance, type of model (single scoring, multiple scoring models). Once a model is deployed and evaluates millions of responses every time for optimal responses, we need a dynamically updated model based on prompt and data. Finally, framework designing of AES model, hear a framework contains prompts where test-takers can write the responses. One can design two frameworks: a single scoring model for a single methodology and multiple scoring models for multiple concepts. When we deploy multiple scoring models, each prompt could be trained separately, or we can provide generalized models for all prompts with this accuracy may vary, and it is challenging.

Our Systematic literature review on the automated essay grading system first collected 542 papers with selected keywords from various databases. After inclusion and exclusion criteria, we left with 139 articles; on these selected papers, we applied Quality assessment criteria with two reviewers, and finally, we selected 62 writings for final review.

Our observations on automated essay grading systems from 2010 to 2020 are as followed:

  • The implementation techniques of automated essay grading systems are classified into four buckets; there are 1. regression models 2. Classification models 3. Neural networks 4. Ontology-based methodology, but using neural networks, the researchers are more accurate than other techniques, and all the methods state of the art provided in Table ​ Table3 3 .
  • The majority of the regression and classification models on essay scoring used statistical features to find the final score. It means the systems or models trained on such parameters as word count, sentence count, etc. though the parameters extracted from the essay, the algorithm are not directly training on essays. The algorithms trained on some numbers obtained from the essay and hear if numbers matched the composition will get a good score; otherwise, the rating is less. In these models, the evaluation process is entirely on numbers, irrespective of the essay. So, there is a lot of chance to miss the coherence, relevance of the essay if we train our algorithm on statistical parameters.
  • In the neural network approach, the models trained on Bag of Words (BoW) features. The BoW feature is missing the relationship between a word to word and the semantic meaning of the sentence. E.g., Sentence 1: John killed bob. Sentence 2: bob killed John. In these two sentences, the BoW is "John," "killed," "bob."
  • In the Word2Vec library, if we are prepared a word vector from an essay in a unidirectional way, the vector will have a dependency with other words and finds the semantic relationship with other words. But if a word has two or more meanings like "Bank loan" and "River Bank," hear bank has two implications, and its adjacent words decide the sentence meaning; in this case, Word2Vec is not finding the real meaning of the word from the sentence.
  • The features extracted from essays in the essay scoring system are classified into 3 type's features like statistical features, style-based features, and content-based features, which are explained in RQ2 and Table ​ Table3. 3 . But statistical features, are playing a significant role in some systems and negligible in some systems. In Shehab et al. ( 2016 ); Cummins et al. ( 2016 ). Dong et al. ( 2017 ). Dong and Zhang ( 2016 ). Mathias and Bhattacharyya ( 2018a ; b ) Systems the assessment is entirely on statistical and style-based features they have not retrieved any content-based features. And in other systems that extract content from the essays, the role of statistical features is for only preprocessing essays but not included in the final grading.
  • In AES systems, coherence is the main feature to be considered while evaluating essays. The actual meaning of coherence is to stick together. That is the logical connection of sentences (local level coherence) and paragraphs (global level coherence) in a story. Without coherence, all sentences in a paragraph are independent and meaningless. In an Essay, coherence is a significant feature that is explaining everything in a flow and its meaning. It is a powerful feature in AES system to find the semantics of essay. With coherence, one can assess whether all sentences are connected in a flow and all paragraphs are related to justify the prompt. Retrieving the coherence level from an essay is a critical task for all researchers in AES systems.
  • In automatic essay grading systems, the assessment of essays concerning content is critical. That will give the actual score for the student. Most of the researches used statistical features like sentence length, word count, number of sentences, etc. But according to collected results, 32% of the systems used content-based features for the essay scoring. Example papers which are on content-based assessment are Taghipour and Ng ( 2016 ); Persing and Ng ( 2013 ); Wang et al. ( 2018a , 2018b ); Zhao et al. ( 2017 ); Kopparapu and De ( 2016 ), Kumar et al. ( 2019 ); Mathias and Bhattacharyya ( 2018a ; b ); Mohler and Mihalcea ( 2009 ) are used content and statistical-based features. The results are shown in Fig. ​ Fig.3. 3 . And mainly the content-based features extracted with word2vec NLP library, but word2vec is capable of capturing the context of a word in a document, semantic and syntactic similarity, relation with other terms, but word2vec is capable of capturing the context word in a uni-direction either left or right. If a word has multiple meanings, there is a chance of missing the context in the essay. After analyzing all the papers, we found that content-based assessment is a qualitative assessment of essays.
  • On the other hand, Horbach and Zesch ( 2019 ); Riordan Brain et al. ( 2019 ); Ding et al. ( 2020 ); Kumar et al. ( 2020 ) proved that neural network models are vulnerable when a student response contains constructed irrelevant, adversarial answers. And a student can easily bluff an automated scoring system by submitting different responses like repeating sentences and repeating prompt words in an essay. From Loukina et al. ( 2019 ), and Madnani et al. ( 2017b ). The fairness of an algorithm is an essential factor to be considered in AES systems.
  • While talking about speech assessment, the data set contains audios of duration up to one minute. Feature extraction techniques are entirely different from text assessment, and accuracy varies based on speaking fluency, pitching, male to female voice and boy to adult voice. But the training algorithms are the same for text and speech assessment.
  • Once an AES system evaluates essays and short answers accurately in all directions, there is a massive demand for automated systems in the educational and related world. Now AES systems are deployed in GRE, TOEFL exams; other than these, we can deploy AES systems in massive open online courses like Coursera(“ https://coursera.org/learn//machine-learning//exam ”), NPTEL ( https://swayam.gov.in/explorer ), etc. still they are assessing student performance with multiple-choice questions. In another perspective, AES systems can be deployed in information retrieval systems like Quora, stack overflow, etc., to check whether the retrieved response is appropriate to the question or not and can give ranking to the retrieved answers.

Conclusion and future work

As per our Systematic literature review, we studied 62 papers. There exist significant challenges for researchers in implementing automated essay grading systems. Several researchers are working rigorously on building a robust AES system despite its difficulty in solving this problem. All evaluating methods are not evaluated based on coherence, relevance, completeness, feedback, and knowledge-based. And 90% of essay grading systems are used Kaggle ASAP (2012) dataset, which has general essays from students and not required any domain knowledge, so there is a need for domain-specific essay datasets to train and test. Feature extraction is with NLTK, WordVec, and GloVec NLP libraries; these libraries have many limitations while converting a sentence into vector form. Apart from feature extraction and training Machine Learning models, no system is accessing the essay's completeness. No system provides feedback to the student response and not retrieving coherence vectors from the essay—another perspective the constructive irrelevant and adversarial student responses still questioning AES systems.

Our proposed research work will go on the content-based assessment of essays with domain knowledge and find a score for the essays with internal and external consistency. And we will create a new dataset concerning one domain. And another area in which we can improve is the feature extraction techniques.

This study includes only four digital databases for study selection may miss some functional studies on the topic. However, we hope that we covered most of the significant studies as we manually collected some papers published in useful journals.

Below is the link to the electronic supplementary material.

Not Applicable.

Publisher's Note

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

Contributor Information

Dadi Ramesh, Email: moc.liamg@44hsemaridad .

Suresh Kumar Sanampudi, Email: ni.ca.hutnj@idupmanashserus .

  • Adamson, A., Lamb, A., & December, R. M. (2014). Automated Essay Grading.
  • Ajay HB, Tillett PI, Page EB (1973) Analysis of essays by computer (AEC-II) (No. 8-0102). Washington, DC: U.S. Department of Health, Education, and Welfare, Office of Education, National Center for Educational Research and Development
  • Ajetunmobi SA, Daramola O (2017) Ontology-based information extraction for subject-focussed automatic essay evaluation. In: 2017 International Conference on Computing Networking and Informatics (ICCNI) p 1–6. IEEE
  • Alva-Manchego F, et al. (2019) EASSE: Easier Automatic Sentence Simplification Evaluation.” ArXiv abs/1908.04567 (2019): n. pag
  • Bailey S, Meurers D (2008) Diagnosing meaning errors in short answers to reading comprehension questions. In: Proceedings of the Third Workshop on Innovative Use of NLP for Building Educational Applications (Columbus), p 107–115
  • Basu S, Jacobs C, Vanderwende L. Powergrading: a clustering approach to amplify human effort for short answer grading. Trans Assoc Comput Linguist (TACL) 2013; 1 :391–402. doi: 10.1162/tacl_a_00236. [ CrossRef ] [ Google Scholar ]
  • Bejar, I. I., Flor, M., Futagi, Y., & Ramineni, C. (2014). On the vulnerability of automated scoring to construct-irrelevant response strategies (CIRS): An illustration. Assessing Writing, 22, 48-59.
  • Bejar I, et al. (2013) Length of Textual Response as a Construct-Irrelevant Response Strategy: The Case of Shell Language. Research Report. ETS RR-13-07.” ETS Research Report Series (2013): n. pag
  • Berzak Y, et al. (2018) “Assessing Language Proficiency from Eye Movements in Reading.” ArXiv abs/1804.07329 (2018): n. pag
  • Blanchard D, Tetreault J, Higgins D, Cahill A, Chodorow M (2013) TOEFL11: A corpus of non-native English. ETS Research Report Series, 2013(2):i–15, 2013
  • Blood, I. (2011). Automated essay scoring: a literature review. Studies in Applied Linguistics and TESOL, 11(2).
  • Burrows S, Gurevych I, Stein B. The eras and trends of automatic short answer grading. Int J Artif Intell Educ. 2015; 25 :60–117. doi: 10.1007/s40593-014-0026-8. [ CrossRef ] [ Google Scholar ]
  • Cader, A. (2020, July). The Potential for the Use of Deep Neural Networks in e-Learning Student Evaluation with New Data Augmentation Method. In International Conference on Artificial Intelligence in Education (pp. 37–42). Springer, Cham.
  • Cai C (2019) Automatic essay scoring with recurrent neural network. In: Proceedings of the 3rd International Conference on High Performance Compilation, Computing and Communications (2019): n. pag.
  • Chen M, Li X (2018) "Relevance-Based Automated Essay Scoring via Hierarchical Recurrent Model. In: 2018 International Conference on Asian Language Processing (IALP), Bandung, Indonesia, 2018, p 378–383, doi: 10.1109/IALP.2018.8629256
  • Chen Z, Zhou Y (2019) "Research on Automatic Essay Scoring of Composition Based on CNN and OR. In: 2019 2nd International Conference on Artificial Intelligence and Big Data (ICAIBD), Chengdu, China, p 13–18, doi: 10.1109/ICAIBD.2019.8837007
  • Contreras JO, Hilles SM, Abubakar ZB (2018) Automated essay scoring with ontology based on text mining and NLTK tools. In: 2018 International Conference on Smart Computing and Electronic Enterprise (ICSCEE), 1-6
  • Correnti R, Matsumura LC, Hamilton L, Wang E. Assessing students’ skills at writing analytically in response to texts. Elem Sch J. 2013; 114 (2):142–177. doi: 10.1086/671936. [ CrossRef ] [ Google Scholar ]
  • Cummins, R., Zhang, M., & Briscoe, E. (2016, August). Constrained multi-task learning for automated essay scoring. Association for Computational Linguistics.
  • Darwish SM, Mohamed SK (2020) Automated essay evaluation based on fusion of fuzzy ontology and latent semantic analysis. In: Hassanien A, Azar A, Gaber T, Bhatnagar RF, Tolba M (eds) The International Conference on Advanced Machine Learning Technologies and Applications
  • Dasgupta T, Naskar A, Dey L, Saha R (2018) Augmenting textual qualitative features in deep convolution recurrent neural network for automatic essay scoring. In: Proceedings of the 5th Workshop on Natural Language Processing Techniques for Educational Applications p 93–102
  • Ding Y, et al. (2020) "Don’t take “nswvtnvakgxpm” for an answer–The surprising vulnerability of automatic content scoring systems to adversarial input." In: Proceedings of the 28th International Conference on Computational Linguistics
  • Dong F, Zhang Y (2016) Automatic features for essay scoring–an empirical study. In: Proceedings of the 2016 Conference on Empirical Methods in Natural Language Processing p 1072–1077
  • Dong F, Zhang Y, Yang J (2017) Attention-based recurrent convolutional neural network for automatic essay scoring. In: Proceedings of the 21st Conference on Computational Natural Language Learning (CoNLL 2017) p 153–162
  • Dzikovska M, Nielsen R, Brew C, Leacock C, Gi ampiccolo D, Bentivogli L, Clark P, Dagan I, Dang HT (2013a) Semeval-2013 task 7: The joint student response analysis and 8th recognizing textual entailment challenge
  • Dzikovska MO, Nielsen R, Brew C, Leacock C, Giampiccolo D, Bentivogli L, Clark P, Dagan I, Trang Dang H (2013b) SemEval-2013 Task 7: The Joint Student Response Analysis and 8th Recognizing Textual Entailment Challenge. *SEM 2013: The First Joint Conference on Lexical and Computational Semantics
  • Educational Testing Service (2008) CriterionSM online writing evaluation service. Retrieved from http://www.ets.org/s/criterion/pdf/9286_CriterionBrochure.pdf .
  • Evanini, K., & Wang, X. (2013, August). Automated speech scoring for non-native middle school students with multiple task types. In INTERSPEECH (pp. 2435–2439).
  • Foltz PW, Laham D, Landauer TK (1999) The Intelligent Essay Assessor: Applications to Educational Technology. Interactive Multimedia Electronic Journal of Computer-Enhanced Learning, 1, 2, http://imej.wfu.edu/articles/1999/2/04/ index.asp
  • Granger, S., Dagneaux, E., Meunier, F., & Paquot, M. (Eds.). (2009). International corpus of learner English. Louvain-la-Neuve: Presses universitaires de Louvain.
  • Higgins D, Heilman M. Managing what we can measure: quantifying the susceptibility of automated scoring systems to gaming behavior” Educ Meas Issues Pract. 2014; 33 :36–46. doi: 10.1111/emip.12036. [ CrossRef ] [ Google Scholar ]
  • Horbach A, Zesch T. The influence of variance in learner answers on automatic content scoring. Front Educ. 2019; 4 :28. doi: 10.3389/feduc.2019.00028. [ CrossRef ] [ Google Scholar ]
  • https://www.coursera.org/learn/machine-learning/exam/7pytE/linear-regression-with-multiple-variables/attempt
  • Hussein, M. A., Hassan, H., & Nassef, M. (2019). Automated language essay scoring systems: A literature review. PeerJ Computer Science, 5, e208. [ PMC free article ] [ PubMed ]
  • Ke Z, Ng V (2019) “Automated essay scoring: a survey of the state of the art.” IJCAI
  • Ke, Z., Inamdar, H., Lin, H., & Ng, V. (2019, July). Give me more feedback II: Annotating thesis strength and related attributes in student essays. In Proceedings of the 57th Annual Meeting of the Association for Computational Linguistics (pp. 3994-4004).
  • Kelley K, Preacher KJ. On effect size. Psychol Methods. 2012; 17 (2):137–152. doi: 10.1037/a0028086. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kitchenham B, Brereton OP, Budgen D, Turner M, Bailey J, Linkman S. Systematic literature reviews in software engineering–a systematic literature review. Inf Softw Technol. 2009; 51 (1):7–15. doi: 10.1016/j.infsof.2008.09.009. [ CrossRef ] [ Google Scholar ]
  • Klebanov, B. B., & Madnani, N. (2020, July). Automated evaluation of writing–50 years and counting. In Proceedings of the 58th Annual Meeting of the Association for Computational Linguistics (pp. 7796–7810).
  • Knill K, Gales M, Kyriakopoulos K, et al. (4 more authors) (2018) Impact of ASR performance on free speaking language assessment. In: Interspeech 2018.02–06 Sep 2018, Hyderabad, India. International Speech Communication Association (ISCA)
  • Kopparapu SK, De A (2016) Automatic ranking of essays using structural and semantic features. In: 2016 International Conference on Advances in Computing, Communications and Informatics (ICACCI), p 519–523
  • Kumar, Y., Aggarwal, S., Mahata, D., Shah, R. R., Kumaraguru, P., & Zimmermann, R. (2019, July). Get it scored using autosas—an automated system for scoring short answers. In Proceedings of the AAAI Conference on Artificial Intelligence (Vol. 33, No. 01, pp. 9662–9669).
  • Kumar Y, et al. (2020) “Calling out bluff: attacking the robustness of automatic scoring systems with simple adversarial testing.” ArXiv abs/2007.06796
  • Li X, Chen M, Nie J, Liu Z, Feng Z, Cai Y (2018) Coherence-Based Automated Essay Scoring Using Self-attention. In: Sun M, Liu T, Wang X, Liu Z, Liu Y (eds) Chinese Computational Linguistics and Natural Language Processing Based on Naturally Annotated Big Data. CCL 2018, NLP-NABD 2018. Lecture Notes in Computer Science, vol 11221. Springer, Cham. 10.1007/978-3-030-01716-3_32
  • Liang G, On B, Jeong D, Kim H, Choi G. Automated essay scoring: a siamese bidirectional LSTM neural network architecture. Symmetry. 2018; 10 :682. doi: 10.3390/sym10120682. [ CrossRef ] [ Google Scholar ]
  • Liua, H., Yeb, Y., & Wu, M. (2018, April). Ensemble Learning on Scoring Student Essay. In 2018 International Conference on Management and Education, Humanities and Social Sciences (MEHSS 2018). Atlantis Press.
  • Liu J, Xu Y, Zhao L (2019) Automated Essay Scoring based on Two-Stage Learning. ArXiv, abs/1901.07744
  • Loukina A, et al. (2015) Feature selection for automated speech scoring.” BEA@NAACL-HLT
  • Loukina A, et al. (2017) “Speech- and Text-driven Features for Automated Scoring of English-Speaking Tasks.” SCNLP@EMNLP 2017
  • Loukina A, et al. (2019) The many dimensions of algorithmic fairness in educational applications. BEA@ACL
  • Lun J, Zhu J, Tang Y, Yang M (2020) Multiple data augmentation strategies for improving performance on automatic short answer scoring. In: Proceedings of the AAAI Conference on Artificial Intelligence, 34(09): 13389-13396
  • Madnani, N., & Cahill, A. (2018, August). Automated scoring: Beyond natural language processing. In Proceedings of the 27th International Conference on Computational Linguistics (pp. 1099–1109).
  • Madnani N, et al. (2017b) “Building better open-source tools to support fairness in automated scoring.” EthNLP@EACL
  • Malinin A, et al. (2016) “Off-topic response detection for spontaneous spoken english assessment.” ACL
  • Malinin A, et al. (2017) “Incorporating uncertainty into deep learning for spoken language assessment.” ACL
  • Mathias S, Bhattacharyya P (2018a) Thank “Goodness”! A Way to Measure Style in Student Essays. In: Proceedings of the 5th Workshop on Natural Language Processing Techniques for Educational Applications p 35–41
  • Mathias S, Bhattacharyya P (2018b) ASAP++: Enriching the ASAP automated essay grading dataset with essay attribute scores. In: Proceedings of the Eleventh International Conference on Language Resources and Evaluation (LREC 2018).
  • Mikolov T, et al. (2013) “Efficient Estimation of Word Representations in Vector Space.” ICLR
  • Mohler M, Mihalcea R (2009) Text-to-text semantic similarity for automatic short answer grading. In: Proceedings of the 12th Conference of the European Chapter of the ACL (EACL 2009) p 567–575
  • Mohler M, Bunescu R, Mihalcea R (2011) Learning to grade short answer questions using semantic similarity measures and dependency graph alignments. In: Proceedings of the 49th annual meeting of the association for computational linguistics: Human language technologies p 752–762
  • Muangkammuen P, Fukumoto F (2020) Multi-task Learning for Automated Essay Scoring with 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 p 116–123
  • Nguyen, H., & Dery, L. (2016). Neural networks for automated essay grading. CS224d Stanford Reports, 1–11.
  • Palma D, Atkinson J. Coherence-based automatic essay assessment. IEEE Intell Syst. 2018; 33 (5):26–36. doi: 10.1109/MIS.2018.2877278. [ CrossRef ] [ Google Scholar ]
  • Parekh S, et al (2020) My Teacher Thinks the World Is Flat! Interpreting Automatic Essay Scoring Mechanism.” ArXiv abs/2012.13872 (2020): n. pag
  • Pennington, J., Socher, R., & Manning, C. D. (2014, October). Glove: Global vectors for word representation. In Proceedings of the 2014 conference on empirical methods in natural language processing (EMNLP) (pp. 1532–1543).
  • Persing I, Ng V (2013) Modeling thesis clarity in student essays. In:Proceedings of the 51st Annual Meeting of the Association for Computational Linguistics (Volume 1: Long Papers) p 260–269
  • Powers DE, Burstein JC, Chodorow M, Fowles ME, Kukich K. Stumping E-Rater: challenging the validity of automated essay scoring. ETS Res Rep Ser. 2001; 2001 (1):i–44. [ Google Scholar ]
  • Powers DE, Burstein JC, Chodorow M, Fowles ME, Kukich K. Stumping e-rater: challenging the validity of automated essay scoring. Comput Hum Behav. 2002; 18 (2):103–134. doi: 10.1016/S0747-5632(01)00052-8. [ CrossRef ] [ Google Scholar ]
  • Ramachandran L, Cheng J, Foltz P (2015) Identifying patterns for short answer scoring using graph-based lexico-semantic text matching. In: Proceedings of the Tenth Workshop on Innovative Use of NLP for Building Educational Applications p 97–106
  • Ramanarayanan V, et al. (2017) “Human and Automated Scoring of Fluency, Pronunciation and Intonation During Human-Machine Spoken Dialog Interactions.” INTERSPEECH
  • Riordan B, Horbach A, Cahill A, Zesch T, Lee C (2017) Investigating neural architectures for short answer scoring. In: Proceedings of the 12th Workshop on Innovative Use of NLP for Building Educational Applications p 159–168
  • Riordan B, Flor M, Pugh R (2019) "How to account for misspellings: Quantifying the benefit of character representations in neural content scoring models."In: Proceedings of the Fourteenth Workshop on Innovative Use of NLP for Building Educational Applications
  • Rodriguez P, Jafari A, Ormerod CM (2019) Language models and Automated Essay Scoring. ArXiv, abs/1909.09482
  • Rudner, L. M., & Liang, T. (2002). Automated essay scoring using Bayes' theorem. The Journal of Technology, Learning and Assessment, 1(2).
  • Rudner, L. M., Garcia, V., & Welch, C. (2006). An evaluation of IntelliMetric™ essay scoring system. The Journal of Technology, Learning and Assessment, 4(4).
  • Rupp A. Designing, evaluating, and deploying automated scoring systems with validity in mind: methodological design decisions. Appl Meas Educ. 2018; 31 :191–214. doi: 10.1080/08957347.2018.1464448. [ CrossRef ] [ Google Scholar ]
  • Ruseti S, Dascalu M, Johnson AM, McNamara DS, Balyan R, McCarthy KS, Trausan-Matu S (2018) Scoring summaries using recurrent neural networks. In: International Conference on Intelligent Tutoring Systems p 191–201. Springer, Cham
  • Sakaguchi K, Heilman M, Madnani N (2015) Effective feature integration for automated short answer scoring. In: Proceedings of the 2015 conference of the North American Chapter of the association for computational linguistics: Human language technologies p 1049–1054
  • Salim, Y., Stevanus, V., Barlian, E., Sari, A. C., & Suhartono, D. (2019, December). Automated English Digital Essay Grader Using Machine Learning. In 2019 IEEE International Conference on Engineering, Technology and Education (TALE) (pp. 1–6). IEEE.
  • Shehab A, Elhoseny M, Hassanien AE (2016) A hybrid scheme for Automated Essay Grading based on LVQ and NLP techniques. In: 12th International Computer Engineering Conference (ICENCO), Cairo, 2016, p 65-70
  • Shermis MD, Mzumara HR, Olson J, Harrington S. On-line grading of student essays: PEG goes on the World Wide Web. Assess Eval High Educ. 2001; 26 (3):247–259. doi: 10.1080/02602930120052404. [ CrossRef ] [ Google Scholar ]
  • Stab C, Gurevych I (2014) Identifying argumentative discourse structures in persuasive essays. In: Proceedings of the 2014 Conference on Empirical Methods in Natural Language Processing (EMNLP) p 46–56
  • Sultan MA, Salazar C, Sumner T (2016) Fast and easy short answer grading with high accuracy. In: Proceedings of the 2016 Conference of the North American Chapter of the Association for Computational Linguistics: Human Language Technologies p 1070–1075
  • Süzen, N., Gorban, A. N., Levesley, J., & Mirkes, E. M. (2020). Automatic short answer grading and feedback using text mining methods. Procedia Computer Science, 169, 726–743.
  • Taghipour K, Ng HT (2016) A neural approach to automated essay scoring. In: Proceedings of the 2016 conference on empirical methods in natural language processing p 1882–1891
  • Tashu TM (2020) "Off-Topic Essay Detection Using C-BGRU Siamese. In: 2020 IEEE 14th International Conference on Semantic Computing (ICSC), San Diego, CA, USA, p 221–225, doi: 10.1109/ICSC.2020.00046
  • Tashu TM, Horváth T (2019) A layered approach to automatic essay evaluation using word-embedding. In: McLaren B, Reilly R, Zvacek S, Uhomoibhi J (eds) Computer Supported Education. CSEDU 2018. Communications in Computer and Information Science, vol 1022. Springer, Cham
  • Tashu TM, Horváth T (2020) Semantic-Based Feedback Recommendation for Automatic Essay Evaluation. In: Bi Y, Bhatia R, Kapoor S (eds) Intelligent Systems and Applications. IntelliSys 2019. Advances in Intelligent Systems and Computing, vol 1038. Springer, Cham
  • Uto M, Okano M (2020) Robust Neural Automated Essay Scoring Using Item Response Theory. In: Bittencourt I, Cukurova M, Muldner K, Luckin R, Millán E (eds) Artificial Intelligence in Education. AIED 2020. Lecture Notes in Computer Science, vol 12163. Springer, Cham
  • Wang Z, Liu J, Dong R (2018a) Intelligent Auto-grading System. In: 2018 5th IEEE International Conference on Cloud Computing and Intelligence Systems (CCIS) p 430–435. IEEE.
  • Wang Y, et al. (2018b) “Automatic Essay Scoring Incorporating Rating Schema via Reinforcement Learning.” EMNLP
  • Zhu W, Sun Y (2020) Automated essay scoring system using multi-model Machine Learning, david c. wyld et al. (eds): mlnlp, bdiot, itccma, csity, dtmn, aifz, sigpro
  • Wresch W. The Imminence of Grading Essays by Computer-25 Years Later. Comput Compos. 1993; 10 :45–58. doi: 10.1016/S8755-4615(05)80058-1. [ CrossRef ] [ Google Scholar ]
  • Wu, X., Knill, K., Gales, M., & Malinin, A. (2020). Ensemble approaches for uncertainty in spoken language assessment.
  • Xia L, Liu J, Zhang Z (2019) Automatic Essay Scoring Model Based on Two-Layer Bi-directional Long-Short Term Memory Network. In: Proceedings of the 2019 3rd International Conference on Computer Science and Artificial Intelligence p 133–137
  • Yannakoudakis H, Briscoe T, Medlock B (2011) A new dataset and method for automatically grading ESOL texts. In: Proceedings of the 49th annual meeting of the association for computational linguistics: human language technologies p 180–189
  • Zhao S, Zhang Y, Xiong X, Botelho A, Heffernan N (2017) A memory-augmented neural model for automated grading. In: Proceedings of the Fourth (2017) ACM Conference on Learning@ Scale p 189–192
  • Zupanc K, Bosnic Z (2014) Automated essay evaluation augmented with semantic coherence measures. In: 2014 IEEE International Conference on Data Mining p 1133–1138. IEEE.
  • Zupanc K, Savić M, Bosnić Z, Ivanović M (2017) Evaluating coherence of essays using sentence-similarity networks. In: Proceedings of the 18th International Conference on Computer Systems and Technologies p 65–72
  • Dzikovska, M. O., Nielsen, R., & Brew, C. (2012, June). Towards effective tutorial feedback for explanation questions: A dataset and baselines. In  Proceedings of the 2012 Conference of the North American Chapter of the Association for Computational Linguistics: Human Language Technologies  (pp. 200-210).
  • Kumar, N., & Dey, L. (2013, November). Automatic Quality Assessment of documents with application to essay grading. In 2013 12th Mexican International Conference on Artificial Intelligence (pp. 216–222). IEEE.
  • Wu, S. H., & Shih, W. F. (2018, July). A short answer grading system in chinese by support vector approach. In Proceedings of the 5th Workshop on Natural Language Processing Techniques for Educational Applications (pp. 125-129).
  • Agung Putri Ratna, A., Lalita Luhurkinanti, D., Ibrahim I., Husna D., Dewi Purnamasari P. (2018). Automatic Essay Grading System for Japanese Language Examination Using Winnowing Algorithm, 2018 International Seminar on Application for Technology of Information and Communication, 2018, pp. 565–569. 10.1109/ISEMANTIC.2018.8549789.
  • Sharma A., & Jayagopi D. B. (2018). Automated Grading of Handwritten Essays 2018 16th International Conference on Frontiers in Handwriting Recognition (ICFHR), 2018, pp 279–284. 10.1109/ICFHR-2018.2018.00056

Information

  • Author Services

Initiatives

You are accessing a machine-readable page. In order to be human-readable, please install an RSS reader.

All articles published by MDPI are made immediately available worldwide under an open access license. No special permission is required to reuse all or part of the article published by MDPI, including figures and tables. For articles published under an open access Creative Common CC BY license, any part of the article may be reused without permission provided that the original article is clearly cited. For more information, please refer to https://www.mdpi.com/openaccess .

Feature papers represent the most advanced research with significant potential for high impact in the field. A Feature Paper should be a substantial original Article that involves several techniques or approaches, provides an outlook for future research directions and describes possible research applications.

Feature papers are submitted upon individual invitation or recommendation by the scientific editors and must receive positive feedback from the reviewers.

Editor’s Choice articles are based on recommendations by the scientific editors of MDPI journals from around the world. Editors select a small number of articles recently published in the journal that they believe will be particularly interesting to readers, or important in the respective research area. The aim is to provide a snapshot of some of the most exciting work published in the various research areas of the journal.

Original Submission Date Received: .

  • Active Journals
  • Find a Journal
  • Proceedings Series
  • For Authors
  • For Reviewers
  • For Editors
  • For Librarians
  • For Publishers
  • For Societies
  • For Conference Organizers
  • Open Access Policy
  • Institutional Open Access Program
  • Special Issues Guidelines
  • Editorial Process
  • Research and Publication Ethics
  • Article Processing Charges
  • Testimonials
  • Preprints.org
  • SciProfiles
  • Encyclopedia

electronics-logo

Article Menu

automated essay scoring dataset

  • Subscribe SciFeed
  • Recommended Articles
  • Google Scholar
  • on Google Scholar
  • Table of Contents

Find support for a specific problem in the support section of our website.

Please let us know what you think of our products and services.

Visit our dedicated information section to learn more about MDPI.

JSmol Viewer

Revealing glcm metric variations across a plant disease dataset: a comprehensive examination and future prospects for enhanced deep learning applications.

automated essay scoring dataset

1. Introduction

2. methodology, 2.1. data collection.

  • “plant_village dataset”;
  • “a_database_of_leaf_images”;
  • “RoCoLe dataset”;
  • “FGVCx_cassava dataset” and
  • “paddy_doctor dataset”.

2.2. Image Preprocessing and GLCM Metrics

2.3. statistical analysis, 2.4. deep learning model training and evaluation, 3. results and discussion, 3.1. ideals of image properties and distribution across the datasets, 3.1.1. data imbalance, 3.1.2. image resolutions, 3.2. distribution of glcm metrics, 3.2.1. energy, 3.2.2. contrast, 3.2.3. correlation, 3.2.4. homogeneity, 3.2.5. angular second moment (asm), 3.2.6. total variance, 3.2.7. maximum probability, 3.2.8. difference variance, 3.2.9. joint entropy, 3.2.10. difference entropy, 3.3. highest-lowest glcm metric’s scorecard, 3.4. correlation matrix of glcm metrics, 3.4.1. strong correlations, 3.4.2. moderate correlations, 3.4.3. weak correlations, 3.4.4. inverse correlation, 3.4.5. kruskal–wallis test of variance, 3.5. deep learning model’s development and analysis, 4. conclusions, 5. recommendations and future works, author contributions, data availability statement, conflicts of interest.

  • FAO. The State of Food and Agriculture, 1974. Lancet 1975 , 306 , 313–314. [ Google Scholar ] [ CrossRef ]
  • FAO. FAO—News Article: New Standards to Curb the Global Spread of Plant Pests and Diseases ; Food and Agriculture Organization of the United Nations: Roma, Italy, 2018. [ Google Scholar ]
  • Horst, R.K. Plant Diseases and Their Pathogens. In Westcott’s Plant Disease Handbook ; Springer: Berlin/Heidelberg, Germany, 2001; pp. 65–530. [ Google Scholar ] [ CrossRef ]
  • FAO. International Year of Plant Health—Final Report ; FAO: Rome, Italy, 2021. [ Google Scholar ] [ CrossRef ]
  • Buja, I.; Sabella, E.; Monteduro, A.G.; Chiriacò, M.S.; De Bellis, L.; Luvisi, A.; Maruccio, G. Advances in plant disease detection and monitoring: From traditional assays to in-field diagnostics. Sensors 2021 , 21 , 2129. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Strange, R.N.; Scott, P.R. Plant Disease: A Threat to Global Food Security. Annu. Rev. Phytopathol. 2005 , 43 , 83–116. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Witten, I.H.; Cunningham, S.; Holmes, G.; McQueen, R.J.; Smith, L.A. Practical Machine Learning and its Potential Application to Problems in Agriculture. In Practical Machine Learning and Its Potential Application to Problems in Agriculture ; Department of Computer Science, University of Waikato: Hamilton, New Zealand, 1993; Volume 1, pp. 308–325. [ Google Scholar ]
  • Shrivastava, V.K.; Pradhan, M.K. Rice plant disease classification using color features: A machine learning paradigm. J. Plant Pathol. 2021 , 103 , 17–26. [ Google Scholar ] [ CrossRef ]
  • Zhou, H.; Li, Y.; Zhang, Q.; Xu, H.; Su, Y. Soft-sensing of effluent total phosphorus using adaptive recurrent fuzzy neural network with Gustafson-Kessel clustering. Expert Syst. Appl. 2022 , 203 , 117589. [ Google Scholar ] [ CrossRef ]
  • Lecun, Y.; Bengio, Y.; Hinton, G. Deep learning. Nature 2015 , 521 , 436–444. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Tibdewal, M.N.; Kulthe, Y.M.; Bharambe, A.; Farkade, A.; Dongre, A. Deep Learning Models for Classification of Cotton Crop Disease Detection. Zeich. J. 2022 , 8 . [ Google Scholar ]
  • Prakash, N.; Udayakumar, E.; Kumareshan, N. Design and development of Android based Plant disease detection using Arduino. In Proceedings of the 2020 7th International Conference on Smart Structures and Systems, ICSSS 2020, Chennai, India, 23–24 July 2020. [ Google Scholar ] [ CrossRef ]
  • Petchiammal, A.; Briskline Kiruba, S.; Murugan, D. Paddy Leaf diseases identification on Infrared Images based on Convolutional Neural Networks. arXiv 2022 , arXiv:2208.00031. [ Google Scholar ] [ CrossRef ]
  • Ahmad, A.; Saraswat, D.; El Gamal, A. A survey on using deep learning techniques for plant disease diagnosis and recommendations for development of appropriate tools. Smart Agric. Technol. 2023 , 3 , 100083. [ Google Scholar ] [ CrossRef ]
  • Hughes, D.P.; Salathe, M. An open access repository of images on plant health to enable the development of mobile disease diagnostics. arXiv 2015 , arXiv:1511.08060. [ Google Scholar ]
  • Pardede, H.F.; Suryawati, E.; Zilvan, V.; Ramdan, A.; Kusumo, R.B.S.; Heryana, A.; Yuwana, R.S.; Krisnandi, D.; Subekti, A.; Fauziah, F.; et al. Plant diseases detection with low resolution data using nested skip connections. J. Big Data 2020 , 7 , 57. [ Google Scholar ] [ CrossRef ]
  • Thakur, P.S.; Sheorey, T.; Ojha, A. VGG-ICNN: A Lightweight CNN model for crop disease identification. Multimed. Tools Appl. 2023 , 82 , 497–520. [ Google Scholar ] [ CrossRef ]
  • Nagi, R.; Tripathy, S.S. Plant disease identification using fuzzy feature extraction and PNN. Signal Image Video Process. 2023 , 17 , 2809–2815. [ Google Scholar ] [ CrossRef ]
  • Hanh, B.T.; Van Manh, H.; Nguyen, N.V. Enhancing the performance of transferred efficientnet models in leaf image-based plant disease classification. J. Plant Dis. Prot. 2022 , 129 , 623–634. [ Google Scholar ] [ CrossRef ]
  • Wiesner-Hanks, T.; Stewart, E.L.; Kaczmar, N.; Dechant, C.; Wu, H.; Nelson, R.J.; Lipson, H.; Gore, M.A. Image set for deep learning: Field images of maize annotated with disease symptoms. BMC Res. Notes 2018 , 11 , 440. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Barbedo, J.G.A. Impact of dataset size and variety on the effectiveness of deep learning and transfer learning for plant disease classification. Comput. Electron. Agric. 2018 , 153 , 46–53. [ Google Scholar ] [ CrossRef ]
  • Pérez-Enciso, M.; Zingaretti, L.M. A Guide for Using Deep Learning for Complex Trait Genomic Prediction. Genes 2019 , 10 , 553. [ Google Scholar ] [ CrossRef ]
  • Singh, D.; Jain, N.; Jain, P.; Kayal, P.; Kumawat, S.; Batra, N. PlantDoc: A dataset for visual plant disease detection. In Proceedings of the ACM International Conference Proceeding Series, Association for Computing Machinery, Dhaka, Bangladesh, 10–12 January 2020; pp. 249–253. [ Google Scholar ] [ CrossRef ]
  • Oyewola, D.O.; Dada, E.G.; Misra, S.; Damaševičius, R. Detecting cassava mosaic disease using a deep residual convolutional neural network with distinct block processing. Peerj Comput. Sci. 2021 , 7 , e352. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Parraga-Alava, J.; Cusme, K.; Loor, A.; Santander, E. RoCoLe: A robusta coffee leaf images dataset for evaluation of machine learning based methods in plant diseases recognition. Data Brief 2019 , 25 , 104414. [ Google Scholar ] [ CrossRef ]
  • Zhang, X.; Han, L.; Dong, Y.; Shi, Y.; Huang, W.; Han, L.; González-Moreno, P.; Ma, H.; Ye, H.; Sobeih, T. A Deep Learning-Based Approach for Automated Yellow Rust Disease Detection from High-Resolution Hyperspectral UAV Images. Remote Sens. 2019 , 11 , 1554. [ Google Scholar ] [ CrossRef ]
  • Bhakta, I.; Phadikar, S.; Majumder, K. Thermal Image Augmentation with Generative Adversarial Network for Agricultural Disease Prediction. In Lecture Notes in Networks and Systems ; Springer Science and Business Media Deutschland GmbH: Berlin/Heidelberg, Germany, 2022; Volume 480, pp. 345–354. [ Google Scholar ] [ CrossRef ]
  • Mohanty, S.P.; Hughes, D.P.; Salathé, M. Using deep learning for image-based plant disease detection. Front. Plant Sci. 2016 , 7 , 215232. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Chen, M.; French, A.P.; Gao, L.; Ramcharan, A.; Hughes, D.P.; Mccloskey, P.; Baranowski, K.; Mbilinyi, N.; Mrisho, L.; Ndalahwa, M.; et al. A Mobile-Based Deep Learning Model for Cassava Disease Diagnosis. Front. Plant Sci. 2019 , 10 , 272. [ Google Scholar ] [ CrossRef ]
  • Johannes, A.; Picon, A.; Alvarez-Gila, A.; Echazarra, J.; Rodriguez-Vaamonde, S.; Navajas, A.D.; Ortiz-Barredo, A. Automatic plant disease diagnosis using mobile capture devices, applied on a wheat use case. Comput. Electron. Agric. 2017 , 138 , 200–209. [ Google Scholar ] [ CrossRef ]
  • Ahmad, J.; Jan, B.; Farman, H.; Ahmad, W.; Ullah, A. Disease detection in plum using convolutional neural network under true field conditions. Sensors 2020 , 20 , 5569. [ Google Scholar ] [ CrossRef ]
  • Petchiammal; Kiruba, B.; Murugan; Arjunan, P. Paddy Doctor: A Visual Image Dataset for Automated Paddy Disease Classification and Benchmarking. In Proceedings of the 6th Joint International Conference on Data Science and Management of Data (10th ACM IKDD CODS and 28th COMAD), Mumbai, India, 4–7 January 2023; pp. 203–207. [ Google Scholar ] [ CrossRef ]
  • Mikołajczyk, A.; Grochowski, M. Data augmentation for improving deep learning in image classification problem. In Proceedings of the 2018 International Interdisciplinary PhD Workshop, IIPhDW 2018, Swinoujscie, Poland, 9–12 May 2018; pp. 117–122. [ Google Scholar ] [ CrossRef ]
  • Velásquez, A.C.; Castroverde, C.D.M.; He, S.Y. Plant–Pathogen Warfare under Changing Climate Conditions. Curr. Biol. 2018 , 28 , R619–R634. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Wang, Q.; He, G.; Li, F.; Zhang, H. A novel database for plant diseases and pests classification. In Proceedings of the ICSPCC 2020—IEEE International Conference on Signal Processing, Communications and Computing, Proceedings, Macau, China, 21–24 August 2020. [ Google Scholar ] [ CrossRef ]
  • Gadkari, D. Image Quality Analysis Using GLCM. In Electronic Theses and Dissertations ; University of Central Florida: Orlando, FL, USA, 2004; pp. 1–120. [ Google Scholar ]
  • Hall-Beyer, M. Practical guidelines for choosing GLCM textures to use in landscape classification tasks over a range of moderate spatial scales. Int. J. Remote Sens. 2017 , 38 , 1312–1338. [ Google Scholar ] [ CrossRef ]
  • Mall, P.K.; Singh, P.K.; Yadav, D. GLCM based feature extraction and medical X-RAY image classification using machine learning techniques. In Proceedings of the 2019 IEEE Conference on Information and Communication Technology, CICT 2019, Allahabad, India, 6–8 December 2019. [ Google Scholar ] [ CrossRef ]
  • Kadir, A. A Model of Plant Identification System Using GLCM, Lacunarity And Shen Features. Publ. Res. J. Pharm. Biol. Chem. Sci. 2014 , 5 , 1–10. [ Google Scholar ]
  • Shoaib, M.; Shah, B.; EI-Sappagh, S.; Ali, A.; Ullah, A.; Alenezi, F.; Gechev, T.; Hussain, T.; Ali, F. An advanced deep learning models-based plant disease detection: A review of recent research. Front. Plant Sci. 2023 , 14 , 1158933. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Gupta, H.P.; Chopade, S.; Dutta, T. Computational Intelligence in Agriculture. In Emerging Computing Paradigms ; John Wiley and Sons, Ltd.: Hoboken, NJ, USA, 2022; pp. 125–142. [ Google Scholar ] [ CrossRef ]
  • Wang, G.; Sun, Y.; Wang, J. Automatic Image-Based Plant Disease Severity Estimation Using Deep Learning. Comput. Intell. Neurosci. 2017 , 2017 , 2917536. [ Google Scholar ] [ CrossRef ]
  • Barbedo, J.G.A.; Koenigkan, L.V.; Halfeld-Vieira, B.A.; Costa, R.V.; Nechet, K.L.; Godoy, C.V.; Junior, M.L.; Patricio, F.R.A.; Talamini, V.; Chitarra, L.G.; et al. Annotated plant pathology databases for image-based detection and recognition of diseases. IEEE Lat. Am. Trans. 2018 , 16 , 1749–1757. [ Google Scholar ] [ CrossRef ]
  • Barbedo, J.G. Factors influencing the use of deep learning for plant disease recognition. Biosyst. Eng. 2018 , 172 , 84–91. [ Google Scholar ] [ CrossRef ]
  • Redmon, J. Darknet: Open Source Neural Networks in C. 2013–2016. Available online: http://pjreddie.com/darknet/ (accessed on 16 May 2024).
  • Theodoridis, S. Machine Learning: A Bayesian and Optimization Perspective , 2nd ed.; Academic Press: Cambridge, MA, USA, 2020; pp. 1–1131. [ Google Scholar ] [ CrossRef ]
AnnotationNameTotal ImagesImage ResolutionsSettingDisease ClassesPlants Involved
dataset_1plant village dataset54,303[256 × 256] throughoutLab38Apple, Cherry, Corn, grape, Peach, Pepper, Potato, Strawberry, and Tomato
dataset_2a database of leaf images4503[6000 × 4000] throughoutLab22Mango, Arjun, Alstonia Scholaris, Guava, Bael, Jamun, Jatropha, Pongamia Pinnata, Basil, Pomegranate, Lemon, and Chinar
dataset_3RoCoLe dataset1560[2048 × 1152], 768 images,
[1280 × 720], 479 images,
[4128 × 2322], 313 images
Field5Coffee
dataset_4FGVCx cassava dataset537Variable between [213 × 231] to [960 × 540]Field5Cassava
dataset_5paddy doctor dataset16,225[1080 × 1440], 16,219 images,
[1440 × 1080], 6 images
Field13Paddy
FactorsEffectsSource
External factors such as uneven lighting, extensive occlusion, and fuzzy details.Variations in the visual characteristics of affected plants.[ ]
Variations in the presence of illness and the growth of a pest.Subtle differences in the characterization of the same diseases and pests in different regions, resulting in “intra-class distinctions”.[ ]
Similarities in the biological morphology and lifestyles of subclasses of diseases and pests.Problem of “inter-class resemblance”.[ ]
Background disturbances.Makes it harder to detect plant pests and diseases In actual agricultural settings.[ ]
GLCM MetricsHighest in GLCMLowest in GLCM
Energy x x
Contrastx x x
Correlation x x
Homogeneity x x
Angular Second Momentxx x
Total Variance x x
Maximum Probabilityx x
Joint Entropy x x
Difference Variance x x
Difference Entropyx x
Sum of scores:4400425021
Table AnalyzedGLCM 10 Parameters
p-value<0.0001
Exact or approximate p-value?Approximate
p-value summary**** (Highly significant differences among the medians)
Do the medians vary significantly (p < 0.05)?Yes
Number of groups10
Kruskal–Wallis statistic619,192
Data summary
Number of treatments (columns)10
Number of values (total)637,010
DatasetsAccuracyPrecisionRecallF1-Score
D10.91220.91410.91230.9111
D20.90600.91160.90610.9056
D30.66660.73290.66670.6411
D40.58660.58850.58670.5867
D50.58970.59960.58970.5852
The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

Kabir, M.; Unal, F.; Akinci, T.C.; Martinez-Morales, A.A.; Ekici, S. Revealing GLCM Metric Variations across a Plant Disease Dataset: A Comprehensive Examination and Future Prospects for Enhanced Deep Learning Applications. Electronics 2024 , 13 , 2299. https://doi.org/10.3390/electronics13122299

Kabir M, Unal F, Akinci TC, Martinez-Morales AA, Ekici S. Revealing GLCM Metric Variations across a Plant Disease Dataset: A Comprehensive Examination and Future Prospects for Enhanced Deep Learning Applications. Electronics . 2024; 13(12):2299. https://doi.org/10.3390/electronics13122299

Kabir, Masud, Fatih Unal, Tahir Cetin Akinci, Alfredo A. Martinez-Morales, and Sami Ekici. 2024. "Revealing GLCM Metric Variations across a Plant Disease Dataset: A Comprehensive Examination and Future Prospects for Enhanced Deep Learning Applications" Electronics 13, no. 12: 2299. https://doi.org/10.3390/electronics13122299

Article Metrics

Article access statistics, further information, mdpi initiatives, follow mdpi.

MDPI

Subscribe to receive issue release notifications and newsletters from MDPI journals

  • Share full article

Advertisement

Supported by

Is Your Driving Being Secretly Scored?

The insurance industry, hungry for insights into how people drive, has turned to automakers and smartphone apps like Life360.

Kathleen Lomax sitting in the driver’s seat of a car, with her daughter Bridget in the passenger’s seat and her other daughter, Morgan, in the back seat.

By Kashmir Hill

Kashmir Hill is a technology reporter who has been covering the surprising use of data from our cars.

You know you have a credit score. Did you know that you might also have a driver score?

The score reflects the safety of your driving habits — how often you slam on the brakes, speed, look at your phone or drive late at night.

While you can see your credit score, you will have a harder time finding out what your driving score is. But auto insurance companies can get it — and that could affect the rate you pay.

For the last two decades, auto insurers have been trying to get people to enroll in programs, commonly called usage-based insurance plans, that monitor their day-to-day driving so rates better reflect the actual risk. But privacy-minded consumers have been reluctant to sign up.

So the industry has taken a different tack, getting data about how people drive from automakers or from apps that drivers already have on their phones. Experts say most people have no idea the insurance industry can track them this way.

After The New York Times revealed that General Motors was sharing driving behavior with LexisNexis, customers filed dozens of lawsuits and the carmaker ended its contract with the data broker. But data is still being collected from other automakers and it is still being collected from apps.

Driving behavior analysis, or telematics, as the insurance industry calls it, could be better for consumers, leading to personalized rates that are more fair. Plus, if people have to pay more for their risky driving, they may drive more cautiously, leading to safer roads. But this will happen only if drivers are aware that their behavior is being monitored.

We are having trouble retrieving the article content.

Please enable JavaScript in your browser settings.

Thank you for your patience while we verify access. If you are in Reader mode please exit and  log into  your Times account, or  subscribe  for all of The Times.

Thank you for your patience while we verify access.

Already a subscriber?  Log in .

Want all of The Times?  Subscribe .

COMMENTS

  1. The Hewlett Foundation: Automated Essay Scoring

    Develop an automated scoring algorithm for student-written essays. Develop an automated scoring algorithm for student-written essays. ... Develop an automated scoring algorithm for student-written essays. Develop an automated scoring algorithm for student-written essays. code. New Notebook. table_chart. New Dataset. tenancy. New Model. emoji ...

  2. Automated Essay Scoring

    26 papers with code • 1 benchmarks • 1 datasets. Essay scoring: Automated Essay Scoring is the task of assigning a score to an essay, usually in the context of assessing the language ability of a language learner. The quality of an essay is affected by the following four primary dimensions: topic relevance, organization and coherence, word ...

  3. Explainable Automated Essay Scoring: Deep Learning Really Has

    Automated Essay Scoring System, Dataset, and Feature Selection. As previously mentioned, this paper reuses the AES system developed by Kumar and Boulanger (2020). The AES models were trained using the ASAP's seventh essay corpus. These narrative essays were written by Grade-7 students in the setting of state-wide assessments in the United ...

  4. ASAP Dataset

    Edit. ASAP (Automated Student Assessment Prize) There are eight essay sets. Each of the sets of essays was generated from a single prompt. Selected essays range from an average length of 150 to 550 words per response. Some of the essays are dependent upon source information and others are not. All responses were written by students ranging in ...

  5. GitHub

    Kaggle 2012 Hewlett Automated Student Assessment Prize (ASAP) - Automated Essay Scoring - Turanga1/Automated-Essay-Scoring

  6. Xiaochr/LLM-AES

    In this study, we investigate the effectiveness of LLMs, specifically GPT-4 and fine-tuned GPT-3.5, as tools for Automated Essay Scoring (AES). Our comprehensive set of experiments, conducted on both public and private datasets, highlights the remarkable advantages of LLM-based AES systems.

  7. An automated essay scoring systems: a systematic literature review

    Automated essay scoring (AES) is a computer-based assessment system that automatically scores or grades the student responses by considering appropriate features. ... Proposed Automated Essay Grading Dataset with Essay Attribute Scores. The first concept features selection depends on the essay type. So the common attributes are Content ...

  8. Deep Learning in Automated Essay Scoring

    The ASAP essay datasets hosted by Kaggle have often been used in the past as a benchmark dataset to measure the performance of automated essay scoring software. Recently, Zupanc and Bosnić ( 2017 ) have developed an automated essay evaluation tool with semantic analysis (a novelty), which included 72 linguistic and content metrics, 29 ...

  9. Enhancing Automated Essay Scoring Performance via Fine-tuning Pre

    We utilize Quadratic Weighted Kappa to evaluate our model on the Automated Student Assessment Prize dataset. Our model outperforms not only state-of-the-art neural models near 3 percent but also the latest statistic model. ... [Enhancing Automated Essay Scoring Performance via Fine-tuning Pre-trained Language Models with Combination of ...

  10. DREsS: Dataset for Rubric-based Essay Scoring on EFL Writing

    Automated essay scoring (AES) is a useful tool in English as a Foreign Language (EFL) writing education, offering real-time essay scores for students and instructors. However, previous AES models were trained on essays and scores irrelevant to the practical scenarios of EFL writing education and usually provided a single holistic score due to the lack of appropriate datasets. In this paper, we ...

  11. A New Handwritten Essay Dataset for Automatic Essay Scoring with A New

    We propose VisEssay, the first essay-scoring dataset containing handwriting images. VisEssay consists of over 13,000 visual essays originating from 25+ professional in-service teachers whose personal scoring accuracy are recorded by his/her scoring history, together with crowdsourced OCR result per handwriting image. ... Automated essay scoring ...

  12. DREsS: Dataset for Rubric-based Essay Scoring on EFL Writing

    However, previous AES models were trained on essays and scores irrelevant to the practical scenarios of EFL writing education and usually provided a single holistic score due to the lack of appropriate datasets. In this paper, we release DREsS, a large-scale, standard dataset for rubric-based automated essay scoring.

  13. [1804.06898] Neural Automated Essay Scoring and Coherence Modeling for

    Automated Essay Scoring (AES) focuses on automatically analyzing the quality of writing and assigning a score to the text. ... For the synthetic dataset, the high scoring ASAP essays are selected based on the indicated score threshold (inclusive). ''total size'' refers to the number of the ASAP essays selected + + their 10 different ...

  14. Essay Scoring Datasets

    Essay Scoring Datasets. Automated Essay Grading. Automated essay grading is one of the most important educational applications of NLP. Given an essay, an automated essay grader's goal is to provide annotations with respect to various dimensions of essay quality such as coherence, argument strength, and technical quality, all without human ...

  15. Language models and Automated Essay Scoring

    In this paper, we present a new comparative study on automatic essay scoring (AES). The current state-of-the-art natural language processing (NLP) neural network architectures are used in this work to achieve above human-level accuracy on the publicly available Kaggle AES dataset. We compare two powerful language models, BERT and XLNet, and ...

  16. satin-04/Automatic-Essay-Scoring

    Automatic-Essay-Scoring (AES) Automated Essay Scoring (AES) is a tool for evaluating and scoring of essays written in response to specific prompts. It can be defined as the process of scoring written essays using computer programs. The process of automating the assessment process could be useful for both educators and learners since it ...

  17. Automated Essay Scoring

    Now we train the model on the dataset. ... Automated Essay Scoring — Kaggle Competition End to End Project Implementation-Part 3. Kindly go through Part 1, Part 2 and Part 3 for complete ...

  18. Applied Sciences

    In large-scale testing and e-learning environments, automated essay scoring (AES) can relieve the burden upon human raters by replacing human grading with machine grading. However, building AES systems based on deep learning requires a training dataset consisting of essays that are manually rated with scores. In this study, we introduce EssayGAN, an automatic essay generator based on ...

  19. PDF ASAP++: Enriching the ASAP Automated Essay Grading Dataset with Essay

    essay attributes, like content, organization, style, etc. Our contribution is the scoring of individual attributes of the essays, like content, organization, style, etc. in the ASAP dataset for the remainder of the essays. 2. Motivation A lot of the work in essay grading today makes use of the ASAP AEG dataset. However, most of the essays only ...

  20. Improving Automated Essay Scoring by Prompt Prediction and Matching

    Each essay score ranges from 40 to 95 points. We divide the entire dataset at random into the training set, validation set, and test set in the ratio of 6:2:2. To alleviate the problem of insufficient data under a single prompt, we apply the entire training set that consists of different prompts for fine-tuning.

  21. An automated essay scoring systems: a systematic literature review

    Zhu W, Sun Y (2020) Automated essay scoring system using multi-model Machine Learning, david c. wyld et al. (eds): mlnlp, bdiot, itccma, csity, dtmn, aifz, sigpro Google Scholar; Wresch W The Imminence of Grading Essays by Computer-25 Years Later Comput Compos 1993 10 45 58 10.1016/S8755-4615(05)80058-1 Google Scholar Cross Ref

  22. Automated Essay Grading Pre-Processed Dataset

    The dataset is for the work "An Empirical Analysis of Machine Learning Models for Automated Essay Grading" - The dataset is a pre-processed version of the used publically available Hewlett Foundation's Automated Student Assessment Prize (ASAP) dataset for experimental evaluation. Users can freely sign-up on Kaggle and download the raw dataset. This dataset has been extensively used in ...

  23. Applying large language models for automated essay scoring for non

    In a recent study conducted by Okgetheng and Takeuchi (), they assessed the efficacy of Open-Calm language models in grading Japanese essays.The research utilized a dataset of approximately 300 ...

  24. GPT-4 Enhances Japanese Essay Scoring

    The study utilized a dataset of 1,400 story-writing scripts from the International Corpus of Japanese as a Second Language (I-JAS), consisting of 1,000 participants representing 12 different first ...

  25. Optimizing Multi-Stuttered Speech Classification: Leveraging Whisper's

    The automated classification of stuttered speech has significant implications ... The problem has been addressed by firstly curating a dataset of multi-stuttered disfluencies from SEP-28k audio clips. Secondly, employing Whis- ... The evaluation metric acquired was F1-score for different combinations of training data and modifications. For the ...

  26. An automated essay scoring systems: a systematic literature review

    Automated essay scoring (AES) is a computer-based assessment system that automatically scores or grades the student responses by considering appropriate features. ... Proposed Automated Essay Grading Dataset with Essay Attribute Scores. The first concept features selection depends on the essay type. So the common attributes are Content ...

  27. Revealing GLCM Metric Variations across a Plant Disease Dataset: A

    This study highlights the intricate relationship between Gray-Level Co-occurrence Matrix (GLCM) metrics and machine learning model performance in the context of plant disease identification. It emphasizes the importance of rigorous dataset evaluation and selection protocols to ensure reliable and generalizable classification outcomes. Through a comprehensive examination of publicly available ...

  28. [2406.06000] ThaiCoref: Thai Coreference Resolution Dataset

    While coreference resolution is a well-established research area in Natural Language Processing (NLP), research focusing on Thai language remains limited due to the lack of large annotated corpora. In this work, we introduce ThaiCoref, a dataset for Thai coreference resolution. Our dataset comprises 777,271 tokens, 44,082 mentions and 10,429 entities across four text genres: university essays ...

  29. Is Your Driving Being Secretly Scored?

    Significant time spent driving at night hurts a person's score because of poorer visibility and drivers on the road who are more likely to be tired or inebriated. But that, Mr. DeLong pointed ...