Next Article in Journal
Extracting Information on Affective Computing Research from Data Analysis of Known Digital Platforms: Research into Emotional Artificial Intelligence
Previous Article in Journal
Aggregated Gaze Data Visualization Using Contiguous Irregular Cartograms
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Systematic Review

Leveraging Vector Space Similarity for Learning Cross-Lingual Word Embeddings: A Systematic Review

Department of Electrical Engineering and Computer Science, University of Cincinnati, Cincinnati, OH 45221, USA
*
Author to whom correspondence should be addressed.
Digital 2021, 1(3), 145-161; https://doi.org/10.3390/digital1030011
Submission received: 9 April 2021 / Revised: 21 June 2021 / Accepted: 25 June 2021 / Published: 1 July 2021

Abstract

:
This article presents a systematic literature review on quantifying the proximity between independently trained monolingual word embedding spaces. A search was carried out in the broader context of inducing bilingual lexicons from cross-lingual word embeddings, especially for low-resource languages. The returned articles were then classified. Cross-lingual word embeddings have drawn the attention of researchers in the field of natural language processing (NLP). Although existing methods have yielded satisfactory results for resource-rich languages and languages related to them, some researchers have pointed out that the same is not true for low-resource and distant languages. In this paper, we report the research on methods proposed to provide better representation for low-resource and distant languages in the cross-lingual word embedding space.

1. Introduction

Word embeddings aim to provide vector representation of words. They have been successfully used in a variety of NLP applications. Cross-lingual word embeddings are the representations of words belonging to different languages in a joint embedding space [1]. Cross-lingual word embedding models have risen to meet the growing need of transferring knowledge across languages. Such representations facilitate applications such as bilingual lexicon induction (BLI) and machine translation by enabling the comparison of the meaning of words across languages. They also make the transfer of models between languages possible. Models trained on resource-rich languages can be reused as a starting point for a model for low-resource languages. Thus, better models can be trained for low-resource languages despite their lack of resources. Such cross-lingual transfer can be an effective way to bridge the much talked about digital language divide.
We followed the line of research that aims to learn cross-lingual embedding spaces from independently trained monolingual word embeddings. These approaches first use a large monolingual corpora to train monolingual word representations. After that, a transformation matrix is learned that maps the word representation in one language to another.
The linear transformation method proposed by Mikolov et al. [2] is one of the most influential methods for learning a mapping between language spaces. The researchers observed that words with similar meanings display similar geometric arrangements in their respective monolingual spaces and decided that such relationships can be captured by a linear mapping between spaces—namely rotation and scaling—learned from known translation pairs. This mapping can then be applied to map one language space to another and thus to learn the translations for words that were not initially known.
Xing et al. [3] proposed normalizing word vectors so that the inner product falls back to the cosine distance. This choice helps solve the inconsistency between using the inner product during training and the cosine distance during the testing of the word embeddings. They constrain the mapping to be orthogonal so that the length normalization is preserved. Orthogonal transformations also preserve angles between vectors. These properties ensure monolingual invariance after transformation, which Artetxe et al. [4] also proposed to exploit. Moreover, from a computational point of view, orthogonality is desirable because it makes possible to compute the solution under orthogonality constraint efficiently using Singular Value Decomposition (SVD) in linear time with respect to the vocabulary size.
Seed lexicons are also an important aspect of learning a mapping from one language space to another. Three types of seed lexicon have been widely used in research:
Mikolov et al. [2] as well as many of the early approaches employed bilingual lexicons that are generated by translating the most frequent words in the source language to their equivalent in the target language using Google Translate. Lexicons built this way are popularly referred to in the literature as off-the-shelf lexicons. The problem with this approach is that the translations generated by Google Translate are not necessarily in the target language domain space. Mikolov et al. used 5000 translation pairs as the seed lexicon, while some of the later approaches demonstrated that even a transformation learned for a dictionary with 25 seed pairs is a feasible option [5].
Another approach for constructing bilingual lexicons is through weak supervision methods. Weak supervision has been achieved through the use of shared numerals [5], cognates [6], or identically spelled strings [7]. It is easy to obtain such weak supervision and researchers who have used them have reported competitive results with those produced by the use of off-the-shelf lexicons.
Whether based on the off-the-shelf dictionaries or weak-supervision in the form of seed-lexicon to learn a cross-lingual representation of words, the mapping approaches are often based on the assumption that the vector representation of words in different languages are isometric to each other. As a result, they learn an orthogonal mapping from one language to another. This assumption, referred to in the literature as the orthogonality assumption, may not always hold, especially when the language pairs in question are etymologically distant.
On the other hand, Conneau et al. [8] propose a fully unsupervised method of learning a mapping from the source to the target space using adversarial methods. The mapping, which can be seen as the generator in a two-player game, is jointly trained to fool the discriminator, which in turn is trained to distinguish between transformed source embeddings and actual target embeddings. They then propose extracting a synthetic dictionary from the shared space and fine-tuning the induced mapping. Artetxe et al. [9] and Zhang et al. [10] also proposed fully unsupervised methods for learning cross-lingual embedding space. Since they do not rely on existing dictionary pairs between two languages to learn a mapping, it can be seen as a way to bridge the resource gap among languages.
Vulić et al. [11] noted that, although the main motivation behind the fully unsupervised approach is to widen the access to the different language technologies for low-resource languages, the approach failed for such use cases specifically. They empirically showed that supervision through a small seed dictionary or weak-supervision in case of closely related languages yields significantly better results.
This study reviews work where the vector space similarity between word embedding spaces has been a consideration in learning a cross-lingual embedding space. Quantifying and leveraging such similarity can provide better representation for low-resource and distant languages in the cross-lingual embedding space. The rest of this paper is organized into three sections. It starts by explaining the method used to search for relevant studies and the criteria considered for including them in this study. Second, the results of the analysis of the manuscripts and the synthesis of the main findings are presented. The final section draws some conclusions from the existing approaches for leveraging vector space similarities of monolingual word embeddings.

2. Materials and Methods

The search for relevant papers was performed in Scopus [12]. Since this single database compiles manuscripts indexed in the other important engineering databases, the enquiry was limited to this search engine. The goal was to retrieve publications using vector space similarity or distance between monolingual embedding spaces in inducing multilingual word embedding space so that the downstream task—specifically bilingual dictionary induction—performance is improved for cases involving low-resource or distant languages. An important part of this document search was constructing a query string with a combination of keywords and Boolean operators. The final query string used is as follows:
    (( "multilingual" OR "cross-lingual" ) AND "word embedding" ) AND
    ( "*supervised"  OR  "mapping" ) AND  ( "orthogonal*" OR "isomorphi*"
    OR  "isometr*"  OR  "distant language*"  OR  "related language*"
    OR  "etymological*" OR "typological*" OR "low-resource" )
    AND ( "bilingual lexicon induction" OR "BLI" )
As can be seen from the string, the orthogonal or isomorphic assumption was an important consideration. Most of the mapping-based methods for inducing cross-lingual word embedding spaces are based on this assumption, which does not hold for typologically distant languages and affects performance for low-resource languages [7]. The source language was limited to English while the document type was limited to conference papers and journal articles. Similarly, the source type was limited to conference proceedings and journals with the selected papers in the final stage of publication. This search, completed on 27 January 2021 returned 120 documents. In the next step, the citation information along with the keywords and abstracts were exported in a CSV file format. Furthermore, papers were filtered by the title and contents of their abstracts. Papers that did not discuss performance issues of distant or low-resource languages in a cross-lingual word embedding space were excluded. Applying these exclusion criteria reduced the number of documents to 52.
The final step consisted of fully reading these 52 papers. The authors reached a consensus for excluding papers that followed a joint learning of the cross-lingual embedding space as well as papers that did not explicitly propose a method to capture the vector space similarity between monolingual embedding spaces in some way. Finally, 26 papers were considered for detailed analysis and inclusion in this systematic review. The overall search method adopted for this systematic review is depicted as a flow diagram in Figure 1. This flow diagram is based on the Preferred Reporting Items for Systematic Reviews and Meta-Analyses (PRISMA) guidelines for identifying, assessing, and summarizing findings from studies that are related and separate at the same time [13].

3. Results

This section provides an analysis of the 26 papers selected using the selection procedure described in Section 2. Based on the principal motivation of each study and to better convey the contribution of this systematic review, the selected papers were divided into three categories as follows: Category 1 groups the studies that directly deal with the orthogonality assumption and proposes distance measures or the empirical weak orthogonality between language pairs to solve the issue. Category 2 consists of the manuscripts that propose varying the hub space or pivoting language of the joint embedding space and learn some cross lingual information in the process. Category 3 corresponds to papers that propose learning neighborhood-sensitive local maps or nonlinear global maps to deal with cases where the orthogonal assumption does not hold.

3.1. Category 1: Quantifying Space Similarity or Degree of Isomorphism

Zhang et al. [10] proposed to view word embeddings as distributions and to minimize the divergence between them by minimizing their Earth Mover’s Distances (EMD). EMD measures the distance between two probability distributions. Their proposed method is unsupervised as they perform this minimization at the distribution level without any word level supervision. Two approaches are developed towards the fulfillment of their proposed idea: Wasserstein Generative Adversarial Network (WGAN) and EMD Minimization under Orthogonal Transformation (EMDOT). In the WGAN approach, a Generator transforms word embeddings in the source language into target space. The critic estimates the Wasserstein Distance between the transformed source embedding distributions and the target embedding distributions. This estimate is used by the Generator during training. The second approach, EMDOT, introduces an orthogonality constraint on the transformation matrix. It alternatingly optimizes a transport matrix and a transformation matrix while holding the other variables fixed. They also show correlation of Earth Mover’s Distance with typological and geographical dissimilarity between chosen languages and suggest it as a possible distance measure for quantifying linguistic differences. In comparing the two methods, the authors noted that EMDOT often converges to local minima while WGAN seems better at landing in the neighborhood of a good optimum. They report BLI results for five language pairs: Chinese–English, Spanish–English, Italian–English, Japanese–Chinese, and Turkish–English. They offer the low-resource setting of Turkish word embeddings as well as the morphological richness of Turkish as possible reasons for poor results in a Turkish–English BLI task.
Søgaard et al. [7] introduced a metric in order to quantify similarity between word embeddings. They started out by establishing that word embeddings are, in fact, not isomorphic, even in the case of highly related languages such as English and German, in which the monolingual embeddings have been trained using the same hyperparameters. They do this by demonstrating the difference between the nearest neighbor graphs of the k top words of both the English and German monolingual embedding spaces. The graphs built from the top k English words and their German counterparts are not isomorphic as well. The same is true for the top k nouns even though word embeddings are considerably adept at capturing relationship between nouns. For these reasons, the authors propose to measure the potential for unsupervised bilingual dictionary induction by capturing the degrees to which graphs are similar. They base their metric in Laplacian Eigenvalues. First, they build adjacency matrices from the nearest neighbor graphs of two monolingual embedding spaces. They then compute the eigensimilarity of the Laplacians by summing up the squared differences of the k largest eigenvalues. They show strong correlation between BLI performance and their proposed graph similarity metric (0.89). They also propose utilizing identically spelled words present in both languages as a form of inexpensive supervision. They include Estonian, Finnish, Greek, Hungarian, Polish, and Turkish in addition to high-resource European languages in their experimental setup. The language choices were made in order to test whether BLI performance is affected by morphological complexities of the agglutinative languages with mixed or double markings. They investigated the impact of language similarity, among other factors, on BLI performance and reported poor performance for English–Finnish and English–Estonian pairs. To ensure that the results do not reflect the difference in training data size, they conducted another experiment using a large corpus for Finnish, which does not improve performance. Further experiments enabled the authors to opine that unsupervised learning between similar, mixed marking languages such as Finnish and Estonian are possible even though they fail when paired with English.
Sharoff et al. [14] sought to extend the resources of a better resourced language to a lesser resourced one by translation matrix orthogonalization and generalized correction of hubness with Weighted Levenshtein Distance (WLD), which the authors propose as a measure of morphological structure. They proposed collecting a seed bilingual dictionary for lower resourced languages from interlinked Wikipedia article titles. To filter out aligned titles that are not necessarily cognates, the authors first used the estimated probabilities for characters present in the words in the seed dictionary to assess a rough WLD between the words. The translation suggestions were scored by a combination of cosine similarity and WLD. To incorporate weight adjustment resulting from the similarity between cognates into the cross lingual emebdding space, they selected the most similar words from the entire lexicons. This longer lexicon was produced from the cross-lingual embedding space using their WLD scores and then used to realign the existing alignment. This re-alignment process helps minimize the distance between known cognates and to preserve the orthogonality of the weight matrix. They reported BLI performance for the English–Italian pair. They also extracted a full set of possible cognates from a Panslavonic embedding space made up of Balto–Slavonic languages: Belorussian, Czech, Croatian, Lithuanian, Polish, Slovak, Slovene, and Ukrainian. Additionally, they reported Named Entity Recognition (NER) results for languages in the Panslavonic space.
Melis and Jaakkola [15] cast the cross-lingual correspondence problem as an Optimal Transport problem. They built on the idea that metric recovery algorithms give rise to word embeddings and exploit Gromov–Wasserstein (GW) distance [16] to measure the way similarities between word pairs reflect across languages. The GW distance measures how pairwise distances computed within different domains compare to each other. First, the pairwise distances on both the source and target sides were computed. Then, from the results of the first step, GW was computed, making it a quadratic problem. The authors also showed that the GW distance can be a measure of semantic discrepancy between languages. This notion of linguistic distance was based on the relational characterization of vector representation of words in different languages. The pairwise GW distances for word embedding similarity matrices of seven languages confirmed their intuition, where the cluster of Romance languages has the shortest distances among themselves while sharing a large distance to classical Chinese. Their approach has two optimization choices, the first of which is required only in large settings. Running their algorithms on full sets of embeddings is not feasible due to memory constraints. For their experiments, they chose to utilize 20,000 embeddings, which is the largest possible size for the personal computer they used. The other optimization choice is the entropy regularization parameter used in the Sinkhorn iteration of their algorithm. Larger values of this parameter, λ , lead to smoother improvements with faster per-iteration runtime at a small decline in performance. Additionally, convergence is faster for calculation of GW distances between closer languages.
Xu et al. [17] proposed to optimize a bidirectional transformation function between languages. To this end, they calculated the Sinkhorn distance [18] as a measure of distributional similarity and minimized the back-translation losses. Their model takes the monolingual embedding spaces as inputs. Word embeddings are translated from one space to another by transfer functions, which are parameterized by neural networks. The Sinkhorn distance is calculated to measure the distributional similarity between the target space embeddings and the translated source space embeddings. This is performed for both embedding spaces. Similarly, back-translation loss is calculated for both spaces as well. The mapping functions were found to be sensitive to the initial parameter choices since the approach is completely unsupervised. The authors chose to minimize the optimal transport distance through adversarial training, using Wasserstein GAN. This first phase of adversarial training limits the search space for model parameters. The authors found that it boosts the training of their model despite being potentially unstable. They attributed the good results for their English–Spanish BLI task to the similarity between the languages and their training corpora being better aligned than any other language pairs. They also observed a performance difference between the two directions of the English–Italian language pair where Italian to English BLI performance is better than English to Italian. They attributed this discrepancy to there being more unique English words in the evaluation set. They also opined that translating from any other language to English would be easier because of there being multiple valid English equivalents for any word in the other language.
Chen and Cardie [19] commented on the failure of unsupervised multilingual word embedding methods in leveraging the cross-lingual interdependencies and proposed a method that leverages the relation that exists between every language pair. Their two stage algorithm includes training multilingual word embeddings using the adversarial approach and then refining them with pseudo-supervision. For the first stage of their method, Multilingual Adversarial Training (MAT), they applied a series of language discriminators for each language in consideration. They learned the mappings by converting vectors from a random language space to another. The multilingual objective of these mappings leverages the relation between every language pair by confusing a language discriminator between its own vectors and those converted from other language spaces. For the refinement also, the authors proposed a generalized approach. Their proposed Multilingual Pseudo-Supervised Refinement (MPSR) induces a dictionary containing high-confidence translation pairs for all language pairs and uses it as weak-supervision to refine the initial mappings. They observed from their experimental results that pivoting through a third distant language when translating between similar languages degrades the performance. This observation is supported by the BLI task performances between Romance language pairs when they pivot through English. On the other hand, when the source and target languages are distant, a well-chosen pivot can improve the performance. Translation between German and Romance languages improves when English is utilized as the pivot language.
Artetxe et al. [9] based their method on a fully unsupervised initialization that allows them to leverage the structural similarity between embeddings. Then, the solution is improved iteratively using a robust self-learning algorithm. Learning a fully unsupervised mapping between two independently trained monolingual embedding is especially challenging since the embedding matrices are unaligned along both dimensions. To solve this problem, they constructed similarity matrices of both the spaces. The matrix dimensions corresponded to words, thereby reducing the mismatch between the embedding spaces. Instead of trying to achieve matrix similarities in an approximate isometry, they sort the matrices and applied the nearest neighbor retrieval to find the translation pairs. Building the initial dictionary in this method lets them leverage the structural similarity between the monolingual embedding spaces. Their stochastic approach because of its dynamic convergence criterion is able to adapt to the difficulty of the task. One downside to this is that it takes longer to converge for difficult language pairs such as English–Turkish and English–Finnish. They reported a correlation between linguistic distance and execution time.
Zhang et al. [20] observed that orthogonal mapping is effective for language pairs in which word embeddings are naturally isomorphic. They proposed a method called iterative normalization to facilitate orthogonal mapping for non-isomorphic language pairs. They noted that two embedding spaces are easily aligned if the word embeddings in those spaces are both length and center invariant. Length invariance serves the dual purpose of maintaining consistency of monolingual word embeddings and the objectives of the cross-lingual mapping, making alignment with translation pairs easier. The authors also proved that center invariance is a prerequisite for orthogonal alignment. In other words, the mean vectors of different languages should be of the same length for an orthogonal mapping to exist between them. Their proposed approach—Iterative Normalization—transforms monolingual word embeddings to be both length and center invariant. To do this, the vectors are mapped into unit vectors, and the collection of vectors is centered about their mean. From their experiments, they found that length and center invariance also help in learning linear non-orthogonal methods of learning cross lingual word embeddings.
Using the Gromov–Hausdorff (GH) distance, Patra et al. [21] empirically demonstrated that the isometric assumption does not hold between embedding spaces, especially for language pairs that are etymologically distant. The Hausdorff distance measures the diametric distance between two metric spaces. Intuitively, this is the measure of the distance between nearest neighbors that are the furthest apart. Gromov–Hausdorff distance provides an estimate of isometry of two spaces by minimizing the Hausdorff distance over the possible isometric transforms between the two spaces. The authors proposed Bilingual Lexicon Induction with Semi-Supervision (BLISS), which optimizes for the weak orthogonality constraint as well as unsupervised matching of distribution and supervised alignment of embedding spaces. The unsupervised loss component of their proposed method aims to match the distributions of the source and target embedding distributions. They learned a transformation matrix to fool a discriminator trained to distinguish between mapped source embeddings and target embeddings. The supervised loss component of their method is computed by minimizing a similarity function that maximizes the similarity between given aligned bilingual word embeddings. BLISS also incorporates a weak orthogonality constraint in the form of a back-translation loss, which is particularly useful in cases of violation of the orthogonality constraint. Their semi-supervised method shows improvement over both the purely supervised and unsupervised baselines, especially for distant language pairs such as English and Chinese. They further reported observations of the small supervision contributing to the stabilization of the training process.
Zhou et al. [22] proposed expressing the monolingual embedding spaces as probability densities. They matched these densities, defined by a Gaussian mixture model, using a method they call normalizing flow. In this method, they took samples from the Gaussian space at every step of training. By doing this, they moved away from training on fixed training samples. Apart from this idea, their approach consists of the following: the model was allowed to learn in both directions by allowing a back-translation loss along with weak supervision using identical strings and Gaussian mixture weights based on frequency matching. The weak orthogonality constraint introduced via back-translation encourages the mappings in both direction to be orthogonal to each other, which results in improved performance for the distant languages. They reported their empirical finding that Gaussian variance needs to be set higher for a language that is morphologically richer than the other in a language pair.
Grave et al. [23] proposed Procustes in the Wassterstein distance, which is based on the estimation of an orthogonal matrix as well as a permutation matrix. The orthogonal matrix maps the points in the source space, while the permutation matrix helps infer the one-to-one correspondence of the mapped points with the target space points. The Wasserstein distance was chosen as a measure of distance between the two sets of points. They initialized the optimization algorithm with a convex relaxation traditionally used for the graph-isomorphism or graph-matching problem. The authors reported that the refinement procedure also helps improve performance for supervised approaches. They offered the noisy and small supervision lexicons as possible reasons for that. Another observation from their results is the impact of batch size on performance. A larger batch size results in better approximation of squared Wasserstein distance and the real loss function.
Beinborn and Choenni [24] defined the correlation of distances between language families with the variation in the semantic organization of concepts across languages as semantic drift and proposed a method for quantifying it. Instead of relying on the availability of lexicon resources, they extracted translation pairs by taking the nearest semantic neighbor. First, they computed a similarity matrix for every word in the English space known as the representational similarity matrix. This symmetric matrix contains the cosine similarity between every pair of words. They then obtained translations of these English words in other languages by finding their nearest neighbor in the semantic space. Then, the representation similarity matrix was computed for the translated words in different languages. The Spearman correlation of the similarity vectors of these translation pairs expresses the similarity between them. The generalization of this measure was used to express similarity between language pairs. A second-order similarity matrix containing the similarity values between all the pairs of languages under consideration was then used to reconstruct a Phylogenatic tree. This tree was then evaluated against a “gold” tree to confirm its validity. The language clusters from word-based representational similarity are able to distinguish between Western and Eastern European languages as well as Germanic and Latin ones with the exception of English. The authors report that Indonesian, Hebrew, and Turkish are languages that do not fit well to the rest of the languages in the experimental setup. The authors observed that language contact is more crucial to semantic drift than a common ancestor language.
Zhang et al. [25] sought to exploit the target side information while training an unsupervised multilingual word embedding space and proposed a novel Wasserstein GAN. This is based on an autoencoder that enables back-translations with the target side. A weak orthogonality constraint is imposed to learn the initial mappings in both directions. The initial mappings are learned by a Wasserstein GAN-based autoencoder where a generator tries generating fake target samples from source samples. A decoder tries to recover the original source sample while a discriminator fits the Wasserstein distance between the generated and target distributions. This choice sets their work in contrast to other similar works and enables learning a cross-lingual embedding space where the languages are etymologically distant from each other. The back-translation network takes the target side embeddings as input. The initial target-source and source-target mappings learned with weak orthogonality constraints are used as initializing weights between the input and the hidden layer and between the hidden layer and output layer of this back-translation network, respectively. The authors attributed the effectiveness of back-translation with target-side to the two-direction mapping preserving the target-side information for both sides, which cannot be achieved by one transformation matrix and its transpose. It also reuses the target-side samples and improves the quality of the cross lingual word embeddings by utilizing the underlying information. In analyzing the errors made by their system, they found that low-frequency words often obtain wrong translations. They opined that low-frequency words with poor semantic can appear as an outlier in the vector space, which would explain the phenomenon.

3.2. Category 2: Varying Hub Space or Pivot Language

Nakashole and Flauger [26] proposed a knowledge distillation training objective that exploits translation paths for low-resource languages through higher resourced ones. Given that the target language is English, they utilized relatively larger seed dictionaries that a related language shares with both the source language and English to learn a transformation. For example, Spanish, which shares a larger seed dictionary with English and is highly related to Portuguese can be utilized to learn an improved mapping from Portuguese to English compared to using the small seed dictionary between Portuguese and English. Their approach makes the direct mapping from source to target language mimic the predictions made through the trilingual path as well as minimizes a loss function. They defined an additional objective that mimics the predictions through multiple trilingual paths. They reported BLI performances for their proposed trilingual paths where Portuguese, Afrikaans, and Danish distill from better-resourced Spanish, Dutch, and Swedish, respectively. They noted that multiple distillation paths require the model to optimize a more difficult function, which may influence the performance. They lost some of the performance gains when they add French and German to the distillation path opposed to when there was only one distillation path through Spanish from Portuguese to English. However, the path weights reflect the linguistic relations with the path through Spanish having the highest weight followed by French and German.
Doval et al. [27] proposed applying a second transformation to refine the already aligned word embeddings in order to integrate them better with each other. The authors empirically demonstrated that there are significant gaps in the cross-lingual space between words and their translated equivalents. They credited this phenomenon to the constraints imposed upon the initial linear mapping that preserves the structure of the monolingual embeddings. Due to language and training corpora differences, this preservation of structure does not translate into optimal performance in downstream tasks. The authors proposed a method that brings each word pair closer together and makes them meet in the middle. The second mapping is learned from the same bilingual dictionary used to learn the initial mapping. The source word embeddings are now approximated to a point that is defined by the average of the source word embeddings and their respective translation embeddings in the cross-lingual space. While analyzing the source of errors in their BLI results, they found that some of the translated words are related to the gold reference word in the target language. They also observed that, after applying their additional transformation, the performance of the baselines improves and there are multiple instances of a translation changing from a closely related word in the target language to the gold reference itself.
Kementchedjhieva et al. [28] proposed using Generalized Procustes Analysis (GPA) to map two languages onto a third latent space instead of onto each other. This approach makes approximating an alignment easier while yielding better performances in low-resource settings by incorporating supporting languages in the alignment process. The latent space captures properties of each of the embedding spaces as well as the ones that emerge as a result of the combination of the spaces. It is an average of the embedding spaces, which means it is more similar to these embedding spaces than they are to each other. As a result, it is easier to find a good mapping from the individual embedding spaces to the latent space than directly mapping one language space to another. According to the authors, GPA learns an alignment robust to morphologically and semantically related neighboring words as well as antonyms of the correct translations. They attributed the loss of precision to a suboptimal method that learns the alignment and a ceiling effect determining the best alignment that can be learned under the orthogonal constraint. To determine the latter, the authors performed a Procustes fit where they learn alignments in a completely supervised manner where the setup should ensure a 100% precision. The experiments confirmed a ceiling effect since Procustes fit is below 100% for all the language pairs in their experimental setup. The authors also observed a correlation between Procustes fit and precision scores for their weakly supervised GPA and attributed this to linguistic differences between source and target language as well as differences in respective training contents. GPA allows for the simultaneous alignment of three or four languages with the additional languages potentially providing support and boosting performance. They reported the BLI performance of this Multi-support GPA (MGPA) where Afrikaans, Bosnian, Estonian, Hebrew, and Occitan leverage the support of German, Russian, Finnish, Arabic, and Spanish, respectively, which are related languages with richer resources. The performance for Hebrew and Occitan improves in this setting.
Alaux et al. [29] proposed to solve the degradation of translation quality for language pairs that do not involve the hub or the pivot language by mapping all vector sets to a common space. Instead of constraining the mappings directly, their formulation constrains every pair of vectors so that they are aligned well. Regarding the choice of alignment weights, they debated whether it is useful to constrain similar languages with higher weights as they already share similarities and whether increasing weights for distant languages might lead to difficulties in learning an alignment. They found simple weighting schemes to work best without making assumptions about linguistic similarities. They trained their models on 6 and 11 languages. The 11 languages were Czech, Danish, Dutch, English, French, German, Italian, Polish, Portuguese, Russian, and Spanish. The authors reported that adding new, distant languages does not affect the performance.
Heyman et al. [30] pointed out the instability associated with fully unsupervised methods for inducing multilingual word embeddings and proposed a robust framework aimed at mitigating that. They proposed a regularization mechanism in order to find a more robust mapping for languages that are distant to the hub language. This mechanism leverages the structural similarity such a language shares with the rest of the languages in the embedding space. To that effect, they learned a shared space by adding new languages incrementally. They demonstrated that, by gradually adding languages, their method is able to leverage the interdependencies that exist between the languages already in the multilingual space and the new language. From their results, they deduced that adding distant languages later leads to better performance in such an iterative process. In comparing their two approaches, Single Hub Space (SHS) and Incremental Hub Space (IHS), they found that, for IHS, where the matrix growth is linear to the number of languages in the space, computation of SVD becomes more expensive. Value dropping [9], which is a technique of avoiding suboptimal local minima by randomly dropping values from the matrix, also slows down the algorithms. The authors found that incrementally adding languages to a multilingual embedding space serves as a regularization process which, if coupled with adding distant languages later in the iteration, can make value dropping irrelevant for IHS. SHS with value dropping and IHS without it have similar training times.
Bai et al. [31] proposed using a third, latent space for capturing cross-lingual features shared between monolingual word embeddings. They implemented their proposed method with a bilingual adversarial autoencoder. This architecture jointly transforms the source and target monolingual embedding space into the proposed latent space. The adversarial mechanism, along with a distance based objective function ensures that the transformed word embeddings are similar. They claimed that their proposed method is better suited to capture cross-lingual features. They also claimed that their proposed linear mapping with autoencoder constraint is an improvement over existing orthogonal mapping methods citing the weakness of the orthogonal assumption.
Lian et al. [32] proposed to use the Wasserstein barycenter as a “mean” language to enforce the transitive relations between all language pairs in the multilingual embedding space. They reasoned that such a choice has more chances of capturing information related to all the languages than when English is chosen as a pivot language. In contrast to Alaux et al. [29], who learn an arithmetic “mean”, which according to these authors, fails to preserve the distributional properties of different languages, a Wasserstein barycenter was learned. This Wasserstein metric captures the geometry of the distributions of different languages, and the barycenter can be considered a virtual universal language and, consequently, a pivot space. They report results from experiments they carried out to test whether including distant languages improve the BLI task accuracy. Their smaller experimental set included Indo-European languages: Croatian, English, French, German, and Italian. The larger experiment setup included two distant languages Finnish and Turkish. The reported accuracies for both the experimental setup show that translating through the barycenter of all the languages significantly improves the accuracies for several language pairs including German–Russian, English–Italian, English–Russian, English–Croatian, and Croatian–Russian. In cases where the latter setup sees decrease in accuracy, the performance is still comparable.

3.3. Category 3: Local Linear Maps or Non-Linear Global Map

Bai et al. [33] noted that semantically similar words in distant language pairs need not be related by a linear relationship. They built on the idea of Faruqui and Dyer [34] of using canonical correlation analysis to map the word embeddings of two languages in a shared vector space but aimed to instead capture a non-linear relationship using kernel canonical correlation analysis. By demonstrating performance improvements in evaluation tasks, they argued that the relationship between semantically equivalent words in different language can be captured better using non-linear relationships, especially in the cases of typologically distant languages.
Huang et al. [35] went beyond word alignment and proposed alignment on a cluster level so that clusters of words have similar distribution across multiple languages. First, they augmented the monolingual word embeddings with their respective cluster of neighborhood words using an extension of the correlational neural network, CorrNet [36], which were then aligned to form the common semantic space. CorrNet is an autoencoder-based approach of Common representation learning (CRL) that embeds different descriptions of the same data in a common subspace. Intuitively, this means that neighborhood of words are distributed consistently across languages so that mapping between the monolingual embedding spaces and the common semantic space is locally smooth. Language-independent character-level representations of words are also achieved using convolutional neural networks. These character-level representations are also concatenated with the word-level representations in the common space. Clusters are also built based on linguistic properties, which are once again aligned in the shared semantic space. Their approach is able to maintain performance even with a seed dictionary made up of 250 word pairs. They use low-resource name tagging to evaluate the quality of multilingual word embeddings. This task identifies named entities from text and classifies them into types such as Person (PER), Location (LOC), Organization (ORG), etc. They experimented on two sets of languages. The first one consists of Amharic and Tigrinya, languages that share the script (Ge’ez) and language family (proto-Semitic). The second one has English, Turkish, and Uighur, which are high-, medium-, and low-resource languages, respectively, and written in two scripts: Latin for English and Turkish, and Arabic for Uighur.
Nakashole [37] proposed to capture the structural similarity of different languages by learning neighborhood sensitive maps. Instead of the original source language embedding space, they learned their mapping from its neighborhood sensitive representation. To this end, they first learned a dictionary of neighborhoods for the source language embeddings. The distinction of neighborhoods was encouraged by imposing an orthogonality constraint. This dictionary of neighborhoods was multiplied to the source language embeddings to obtain a representation factorized by the neighborhoods. Next, an intermediate representation was achieved by concatenating the factorized representation with the original embeddings. Finally, this intermediate representation was projected into a lower dimension equal to the original embeddings to obtain the final representation on which the transformation was learned. They stressed the significance for learning a neighborhood sensitive map for distant languages since a global linear map may not sufficiently capture the difference in underlying structures for such language embeddings. They noted that nouns and verbs constitute much of the test data and reported higher accuracy in translating nouns than verbs. They also interpreted the discovered neighborhoods. The top word neighborhoods obtained from training the English–German space seem to represent topics. The authors reported variation in the granularity of neighborhoods and their specificity. They report BLI performance on a new English–Portuguese dictionary containing rare words.
Nakashole and Faluger [38] non-linearly mapped one language space to another. They asked whether a single linear map exists that produces comparable performance irrespective of where the translation embeddings fall in the target space. They also pondered over the possible relationship between neighborhood specific maps and the respective neighborhood distances. The experimental results showed that, in small enough neighborhoods, the transformation can be approximated by linear maps and that the amount by which local linear approximations vary correlate with the training neighborhood distances.
Moshtaghi et al. [39] introduced noise tolerance to Generalized Procustes. Their method is fixed for rotational differences between a pair of embedding spaces and accounts for geometrical translation and dilation. In Euclidean geometry, translation denotes a change in location, whereas dilation denotes change in the size of a shape without the shape undergoing a change. Their model LLmap also allows for differences in linear mapping functions in different subspaces. This enables them to perform corrections specific to the regions in the embedding space. They achieved these two objectives by incorporating a noise-tolerant version of Generalized Procustes mapping, which they call Robust Generalized Procustes (RGP), into their Locally Linear Model. RGP enables them to add translation and dilation to the general Procustes. Additionally, they introduced a piece-wise linear mapping between two language spaces. This is based on a Locally Linear Neural Model (LLNF). By increasing the flexibility of the models in tackling the various levels of structural similarities encoded in the embbedding spaces, this model solves the undesired artifacts that may result from a global distance minimization. The computational complexity of the proposed mapping linearly increases with respect to the number of neurons. They dealt with this issue by limiting the number of neurons to four. On the other hand, finding the best dimension to split is the most expensive training step. The authors noted that this was performed as a preprocessing step.
Fan et al. [40] proposed learning multiple mapping matrices with orthogonal constraint since, according to them, a single matrix may not be able to capture the complex linguistic regularities reflected in the semantics of words across different topics and domains. Each of these matrices capture the translation knowledge between two monolingual embedding spaces over distributed latent topics. To discover latent topics, they clustered the words in the source embedding space, resulting in a training set where each translation pair gets distributed across all the latent topics. Learning a single matrix for each of these latent topics harms the performance for low-resource languages. Instead of hard clustering, they introduced a soft piecewise mapping that determines a degree of membership of each translation pair to a number of latent topics predetermined through fine-tuning over the validation set. They reported BLI performance for English to and from Spanish, French, German, Italian, Chinese, and Vietnamese for general domain words. Additionally, they reported English–German and Chinese–English BLI performance for medical words and e-product words, respectively. From the reported results, the authors concluded that their proposed multiple mapping matrices improved BLI performance for low-frequency domain words.

3.4. Other(s)

Gu et al. [41] proposed to apply a model agnostic meta-learning algorithm to facilitate low-resource machine translation and to use universal lexical representation to overcome challenges associated with mismatched input and output in different languages. Meta Learning can be seen as an approach to learn a good initial parameter for fast adaptation. To apply this idea in low-resource machine translation, the authors proposed parameter initialization using high-resource languages and then using the learned parameters in low-resource scenarios. They observe that, in a low-resource scenario, finding a good parameter initialization is strongly correlated with performances of the final model.

4. Discussion

This section presents an analysis of the results reported in Section 3. As mentioned before, the reported 26 papers can be divided into three broad categories. The rationale for this categorization can be explained in light of the seminal papers in the field of cross-lingual word embeddings. Mikolov et al. [2] reported the observation of similar geometric arrangements for semantically equivalent words in monolingual embedding spaces of different languages. They captured this relationship with a linear mapping (rotation and scaling) from one space to another. Xing et al. [3] and Artetxe et al. [4] proposed constraining the linear mapping to be orthogonal to ensure monolingual invariance. Conneau et al. [8], while proposing a fully unsupervised method, also endorsed an orthogonal mapping. Søgaard et al. [7] and Patra et al. [21] demonstrated that the assumption of an orthogonal mapping existing from one language space to another does not always hold, especially for etymologically distant languages. The three categories presented in this systematic review are representative of different ways to address the weak-orthogonality that may exist between the monolingual word embedding spaces. The papers listed under Section 3.1 seek to quantify and optimize for this weak-orthogonality during the training process. Minimization of this measure can lead to better mapping for etymologically distant languages, which is not possible in a cross-lingual model that is agnostic of the inter-lingual distances or similarities. On the other hand, the papers in Section 3.2 reason that mapping a monolingual embedding space onto another that is distant to it makes the mapping less isomorphic, resulting in poor performance. They sought to either have the hub space encode the linguistic properties of all the languages present in the cross-lingual embedding space so that the mappings are more isomorphic or to choose one from the existing monolingual space that can work best as a hub for the set of languages. These two categories try to work around the orthogonality assumption in learning a cross-lingual embedding space. The papers in Section 3.3 argued that there can be no global linear map for majority of language pairs. Hence, they proposed to learn a global, non-linear map, or multiple neighborhood-sensitive linear maps in order to better capture the linguistic properties.

4.1. Analysis of Category 1: Quantifying Space Similarity or Degree of Isomorphism

The manuscripts in Section 3.1 presented various ways of quantifying the degree of orthogonality between the monolingual embedding spaces of distant languages. This varying degree of orthogonality can also be interpreted as linguistic similarity or distance. Some of the distance measures proposed to measure the similarity between vector spaces are Wasserstein distance [10,23], Gromov-Wasserstein distance [15], Eigensimilarity [7], Gromov–Hausdorff distance [21], and Sinkhorn distance [17]. On the other hand, some of these researches propose constructing similarity matrices in both source and target side and taking their correlation as a measure of cross lingual similarity [9,24]. Some other methods introduced a weak-orthogonality constraint as part of the training architecture [21,22,25]. Iterative normalization was proposed in one of these papers to maintain the length and center invariance and to make distant language spaces more isomorphic [20]. These papers and their key ideas are listed in Table 1.
As stated in Section 1, researchers have steadily been trying to make learning cross-lingual mapping less dependent on supervision. A number of researches reported in Section 3.1 proposed fully unsupervised methods for the BLI task [9,10,15,17,19,23,25]. They also sought to explicitly optimize for the cross-lingual similarity or distance. On the other hand, [7,21,22] proposed weakly supervised methods. All of the papers reported BLI task performance for their methods. Turkish and Finnish are frequently reported to be among languages that are difficult to align because of their morphological richness [7,9,10]. Sharoff et al. [14] reported the Named Entity Recognition (NER) task performance for their Panslavonic language space, one of the few instances of cross-lingual transfer in this section. Beinborn et al. [24] moved beyond engineering goals and sought to analyze the relationship between the computational representation of languages. They also worked with the most number of languages as they reconstructed a phylogenetic tree with 28 languages. This can be an interesting avenue of research and can be expanded to include more languages from diverse language families apart from the Romance, Germanic, and Slavic languages.

4.2. Analysis of Category 2: Varying Hub Space or Pivot Language

Traditional mapping-based methods choose one of the language spaces—usually English—as the hub space in the multilingual embedding space, where the rest of the language spaces are mapped [42]. The papers in Section 3.2 move away from that idea. Some of them propose mapping the language spaces not to the source or the target space but instead a third, latent space to preserve language-specific information and to make the mappings more isomorphic [28,29,31,32]. Although similar in their general idea, they differ in nuances. For example, one of the papers proposed constructing the “mean” hub space by taking the arithmetic mean of all the language spaces [29] while another proposed learning a Wasserstein barycenter to better preserve distributional properties of different languages [32]. An incremental hub space was proposed as well, where new languages are added sequentially [30]. Improvement in performance is noted in such a space if distant languages are added later. Another interesting approach was using a related intermediate language as a pivot for low-resource, distant languages [26]. These papers and their key ideas are listed in Table 2.
Of the seven papers listed in Section 3.2, five followed the unsupervised method of learning the cross-lingual space [28,29,30,31,32] while [26,27] using supervision signals. The proposed method of Nakashole and Flauger [26] leverages the comparatively rich resources of a related language in favor of low-resource languages and works for language triplets. Bai et al. [31] and Doval et al. [27] learned a bilingual embedding space while the rest of the researchers optimized for a multilingual setting [29,30,32]. Apart from BLI performances, Heyman et al. [30] reported the performance for multilingual dependency parsing and document classification, which are interesting examples of how multilingual word embeddings can be utilized in cross-lingual transfer. An underlying motivation of these papers is to embed the hub space with the linguistic properties of the languages in the experimental setup. Lian et al. [32] used the term “potential universal language” to express this idea, which can be explored further as a research area.

4.3. Ananlysis of Category 3: Local Linear Maps or Non-Linear Global Map

The papers in Section 3.3 proposed moving away from learning a global linear mapping with orthogonal constraint to learning neighborhood-sensitive local mapping [35,37,40] or non-linear mapping [33,38,39]. These papers and their key ideas are listed in Table 3. All five papers listed under category 3 (see Section 3.3) utilize cross-lingual supervision. While the authors of [37,38,40] learned a bilingual space, those of [33,35,39] learned a multilingual representation. Huang et al. [35] reported the performance for a low-resource NER task. Nakashole [37] and Fan et al. [40] reported BLI performances for rare words and domain-specific words, which is an interesting area of future research. Bai et al. [33] reported multilingual document classification performance. Huang et al. [35] reported NER performance for low-resource proto-Semitic languages Amharic and Tigrinya, while Moshtaghi et al. [39] reported BLI performances for Hindi and Bengali, instances of the otherwise overlooked Indic language family.
Most commonly reported benchmarks for BLI are MUSE [8] and VecMap [9]. The experiments were performed either using the pre-trained fastText embeddings trained on Wikipedia data [43] or embeddings trained under similar settings. The trained cross-lingual word embeddings can be used for cross-lingual transfer. Such a transfer allows a model trained with resource-rich language data to be used by low-resource languages sharing cross-lingual features. Among the downstream tasks reported to evaluate multilingual word embeddings are document classification, NER, dependency parsing and parts-of-speech (POS) tagging [1]. These are traditional NLP tasks now used to evaluate cross-lingual word embeddings. The xling-eval [44] benchmark includes downstream tasks along with BLI. Better alignment of languages in the cross-lingual embedding space has so far resulted in better performance in evaluation and downstream tasks. Hence, learning better alignments by leveraging vector space similarities is a research area worth pursuing.

Author Contributions

Conceptualization, K.B. and A.R.; methodology, K.B. and A.R.; formal analysis, K.B.; data curation, K.B.; data analysis, K.B.; writing and editing, K.B. and A.R. Both authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not Applicable.

Informed Consent Statement

Not Applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ruder, S.; Vulić, I.; Søgaard, A. A survey of cross-lingual word embedding models. J. Artif. Intell. Res. 2019, 65, 569–631. [Google Scholar] [CrossRef] [Green Version]
  2. Mikolov, T.; Quoc, V.L.; Sutskever, I. Exploiting similarities among languages for machine translation. arXiv 2013, arXiv:1309.4168. [Google Scholar]
  3. Xing, C.; Wang, D.; Liu, C.; Lin, Y. Normalized word embedding and orthogonal transform for bilingual word translation. In Proceedings of the 2015 Conference of the North American Chapter of the Association for Computational Linguistics: Human Language Technologies, Denver, CO, USA, 31 May–5 June 2015; 2015. [Google Scholar]
  4. Artetxe, M.; Labaka, G.; Agirre, E. Learning principled bilingual mappings of word embeddings while preserving monolingual invariance. In Proceedings of the 2016 Conference on Empirical Methods in Natural Language Processing, Austin, TX, USA, 1–5 November 2016. [Google Scholar]
  5. Artetxe, M.; Labaka, G.; Agirre, E. Learning bilingual word embeddings with (almost) no bilingual data. In Proceedings of the 55th Annual Meeting of the Association for Computational Linguistics, Volume 1: Long Papers, Vancouver, BC, Canada, 30 July–4 August 2017. [Google Scholar]
  6. Smith, S.L.; Turban, D.H.P.; Hamblin, S.; Hammerla, N.Y. Offline bilingual word vectors, orthogonal transformations and the inverted softmax. arXiv 2017, arXiv:1702.03859. [Google Scholar]
  7. Søgaard, A.; Ruder, S.; Vulić, I. On the limitations of unsupervised bilingual dictionary induction. arXiv 2018, arXiv:1805.03620. [Google Scholar]
  8. Conneau, A.; Lample, G.; Ranzato, M.; Denoyer, L.; Jégou, H. Word Translation without Parallel Data. arXiv 2017, arXiv:1710.04087. [Google Scholar]
  9. Artetxe, M.; Labaka, G.; Agirre, E. A robust self-learning method for fully unsupervised cross-lingual mappings of word embeddings. arXiv 2018, arXiv:1805.06297. [Google Scholar]
  10. Zhang, M.; Liu, Y.; Luan, H.; Sun, M. Earth mover’s distance minimization for unsupervised bilingual lexicon induction. In Proceedings of the 2017 Conference on Empirical Methods in Natural Language Processing, Copenhagen, Denmark, 7–11 September 2017. [Google Scholar]
  11. Vulić, I.; Glavaš, G.; Reichart, R.; Korhonen, A. Do we really need fully unsupervised cross-lingual embeddings? arXiv 2019, arXiv:1909.01638. [Google Scholar]
  12. Advanced Search on Scopus.com. Available online: https://0-www-scopus-com.brum.beds.ac.uk/search/form.uri?display=advanced (accessed on 27 January 2021).
  13. Moher, D.; Liberati, A.; Tetzlaff, J.; Altman, D.G.; PRISMA Group. Preferred reporting items for systematic reviews and meta-analyses: The PRISMA statement. Ann. Int. Med. 2009, 151, 264–269. [Google Scholar] [CrossRef] [Green Version]
  14. Sharoff, S. Language adaptation experiments via cross-lingual embeddings for related languages. In Proceedings of the Eleventh International Conference on Language Resources and Evaluation (LREC 2018), Miyazaki, Japan, 7–12 May 2018. [Google Scholar]
  15. Alvarez-Melis, D.; Jaakkola, T. Gromov-Wasserstein alignment of word embedding spaces. arXiv 2018, arXiv:1809.00013. [Google Scholar]
  16. Mémoli, F. Gromov–Wasserstein distances and the metric approach to object matching. Found. Comput. Math. 2011, 11, 417–487. [Google Scholar] [CrossRef]
  17. Xu, R.; Yang, Y.; Otani, N.; Wu, Y. Unsupervised cross-lingual transfer of word embedding spaces. arXiv 2018, arXiv:1809.03633. [Google Scholar]
  18. Cuturi, M. Sinkhorn distances: Lightspeed computation of optimal transport. In Proceedings of the Advances in Neural Information Processing Systems 26 (NIPS 2013), Lake Tahoe, NV, USA, 5–10 December 2013; pp. 2292–2300. [Google Scholar]
  19. Chen, X.; Cardie, C. Unsupervised multilingual word embeddings. arXiv 2018, arXiv:1808.08933. [Google Scholar]
  20. Zhang, M.; Xu, K.; Kawarabayashi, K.; Jegelka, S.; Boyd-Graber, J. Are Girls Neko or Shōjo? Cross-Lingual Alignment of Non-Isomorphic Embeddings with Iterative Normalization. arXiv 2019, arXiv:1906.01622. [Google Scholar]
  21. Patra, B.; Ruben, J.; Moniz, A.; Garg, S.; Gormley, M.R.; Neubig, G. Bilingual lexicon induction with semi-supervision in non-isometric embedding spaces. arXiv 2019, arXiv:1908.06625. [Google Scholar]
  22. Zhou, C.; Ma, X.; Wang, D.; Neubig, G. Density matching for bilingual word embedding. arXiv 2019, arXiv:1904.02343. [Google Scholar]
  23. Grave, E.; Joulin, A.; Berthet, Q. Unsupervised alignment of embeddings with wasserstein procrustes. In Proceedings of the 22nd International Conference on Artificial Intelligence and Statistics, PMLR, Okinawa, Japan, 16–18 April 2019. [Google Scholar]
  24. Beinborn, L.; Choenni, R. Semantic drift in multilingual representations. Comput. Linguist. 2020, 46, 571–603. [Google Scholar] [CrossRef]
  25. Zhang, Y.; Li, Y.; Zhu, Y.; Hu, X. Wasserstein GAN based on Autoencoder with back-translation for cross-lingual embedding mappings. Pattern Recognit. Lett. 2020, 129, 311–316. [Google Scholar] [CrossRef]
  26. Nakashole, N.; Flauger, R. Knowledge distillation for bilingual dictionary induction. In Proceedings of the 2017 Conference on Empirical Methods in Natural Language Processing, Copenhagen, Denmark, 7–11 September 2017. [Google Scholar]
  27. Doval, Y.; Camacho-Collados, J.; Espinosa-Anke, L.; Schockaert, S. Improving cross-lingual word embeddings by meeting in the middle. arXiv 2018, arXiv:1808.08780. [Google Scholar]
  28. Kementchedjhieva, Y.; Ruder, S.; Cotterell, R.; Søgaard, A. Generalizing procrustes analysis for better bilingual dictionary induction. arXiv 2018, arXiv:1809.00064. [Google Scholar]
  29. Alaux, J.; Grave, E.; Cuturi, M.; Joulin, A. Unsupervised hyperalignment for multilingual word embeddings. arXiv 2018, arXiv:1811.01124. [Google Scholar]
  30. Heyman, G.; Verreet, B.; Vulić, I.; Moens, M. Learning unsupervised multilingual word embeddings with incremental multilingual hubs. In Proceedings of the 2019 Conference of the North American Chapter of the Association for Computational Linguistics: Human Language Technologies, Volume 1 (Long and Short Papers), Minneapolis, MN, USA, 2–7 June 2019. [Google Scholar]
  31. Bai, X.; Cao, H.; Chen, K.; Zhao, T. A bilingual adversarial autoencoder for unsupervised bilingual lexicon induction. IEEE/ACM Trans. Audio Speech Lang. Process. 2019, 27, 1639–1648. [Google Scholar] [CrossRef]
  32. Lian, X.; Jain, K.; Truszkowski, J.; Poupart, P.; Yu, Y. Unsupervised Multilingual Alignment using Wasserstein Barycenter. arXiv 2020, arXiv:2002.00743. [Google Scholar]
  33. Bai, X.; Cao, H.; Zhao, T. Improving vector space word representations via kernel canonical correlation analysis. ACM Trans. Asian Low Resour. Lang. Inf. Process. 2018, 17, 1–16. [Google Scholar] [CrossRef]
  34. Ammar, W.; Mulcaire, G.; Tsvetkov, Y.; Lample, G.; Dyer, C.; Smith, N.A. Massively multilingual word embeddings. arXiv 2016, arXiv:1602.01925. [Google Scholar]
  35. Huang, L.; Cho, K.; Zhang, B.; Ji, H.; Knight, K. Multi-lingual common semantic space construction via cluster-consistent word embedding. arXiv 2018, arXiv:1804.07875. [Google Scholar]
  36. Chandar, S.; Khapra, M.M.; Larochelle, H.; Ravindran, B. Correlational neural networks. Neural Comput. 2016, 28, 257–285. [Google Scholar] [CrossRef]
  37. Nakashole, N. NORMA: Neighborhood sensitive maps for multilingual word embeddings. In Proceedings of the 2018 Conference on Empirical Methods in Natural Language Processing, Brussels, Belgium, 31 October–4 November 2018. [Google Scholar]
  38. Nakashole, N.; Flauger, R. Characterizing departures from linearity in word translation. arXiv 2018, arXiv:1806.04508. [Google Scholar]
  39. Moshtaghi, M. Supervised and Nonlinear Alignment of Two Embedding Spaces for Dictionary Induction in Low Resourced Languages. In Proceedings of the 2019 Conference on Empirical Methods in Natural Language Processing and the 9th International Joint Conference on Natural Language Processing (EMNLP-IJCNLP), Hong Kong, China, 3–7 November 2019. [Google Scholar]
  40. Fan, Y.; Wang, C.; Chen, B.; Hu, Z.; He, X. SPMM: A Soft Piecewise Mapping Model for Bilingual Lexicon Induction. In Proceedings of the 2019 SIAM International Conference on Data Mining, Calgary, AB, Canada, 2–4 May 2019; Society for Industrial and Applied Mathematics: Philadelphia, PA, USA, 2019. [Google Scholar]
  41. Gu, J.; Wang, Y.; Chen, Y.; Cho, K.; Li, V.O.K. Meta-learning for low-resource neural machine translation. arXiv 2018, arXiv:1808.08437. [Google Scholar]
  42. Anastasopoulos, A.; Neubig, G. Should All Cross-Lingual Embeddings Speak English? arXiv 2019, arXiv:1911.03058. [Google Scholar]
  43. Bojanowski, P.; Grave, E.; Joulin, A.; Mikolov, T. Enriching word vectors with subword information. Trans. Assoc. Comput. Linguist. 2017, 5, 135–146. [Google Scholar] [CrossRef] [Green Version]
  44. Glavas, G.; Litschko, R.; Ruder, S.; Vulic, I. How to (properly) evaluate cross-lingual word embeddings: On strong baselines, comparative analyses, and some misconceptions. arXiv 2019, arXiv:1902.00508. [Google Scholar]
Figure 1. Preferred Reporting Items for Systematic Reviews and Meta-Analyses (PRISMA) flowchart of the literature search and study selection.
Figure 1. Preferred Reporting Items for Systematic Reviews and Meta-Analyses (PRISMA) flowchart of the literature search and study selection.
Digital 01 00011 g001
Table 1. Quantifying space similarity or degree of isomorphism.
Table 1. Quantifying space similarity or degree of isomorphism.
AuthorYearKey Idea(s)
Zhang et al. [10]2017Earth mover’s distance minimization
Sharoff [14]2018Weighted Levenshtein distance
Melis and Jaakkola [15]2018Gromov-Wasserstein distance
Xu et al. [17]2018Sinkhorn distance, Back-translation losses
Chen and Cardie [19]2018Adversarial training
Artetxe et al. [9]2018Fully unsupervised initialization
Søgaard et al. [7]2019Eigen similarity of nearest neighbor graphs
Zhang et al. [20]2019Ensuring length and center invariance
Patra et al. [21]2019Gromov-Hausdorff Distance
Zhou et al. [22]2019Embedding space as Gaussian mixture model
Grave et al. [23]2019Permutation matrix, Wasserstein distance
Beinborn and Choenni [24]2020Similarity matrix, Spearman correlation
Zhang et al. [25]2020Wasserstein GAN, Back translation
Table 2. Varying hub space or a pivot language.
Table 2. Varying hub space or a pivot language.
AuthorYearKey Idea(s)
Nakashole and Flauger [26]2017Knowledge distillation
Doval et al. [27]2018Second mapping with same lexicon
Kementchedjhieva et al. [28]2018Generalized Procustes Analysis
Alaux et al. [29]2018Common mapping space
Heyman et al. [30]2019Shared space, incremental mapping
Bai et al. [31]2019Mapping to a third, latent space
Lian et al. [32]2020Wasserstein barycenter as mean language
Table 3. Locally linear maps or non-linear global map.
Table 3. Locally linear maps or non-linear global map.
AuthorYearKey Idea(s)
Bai et al. [33]2018Kernel canonical correlation
Huang et al. [35]2018Cluster level alignment
Nakashole [37]2018Neighborhood sensitive maps
Nakashole and Flauger [38]2018Non-linear mapping
Moshtaghi [39]2019Locally linear mapping
Fan et al. [40]2019Multiple mapping matrices
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bhowmik, K.; Ralescu, A. Leveraging Vector Space Similarity for Learning Cross-Lingual Word Embeddings: A Systematic Review. Digital 2021, 1, 145-161. https://0-doi-org.brum.beds.ac.uk/10.3390/digital1030011

AMA Style

Bhowmik K, Ralescu A. Leveraging Vector Space Similarity for Learning Cross-Lingual Word Embeddings: A Systematic Review. Digital. 2021; 1(3):145-161. https://0-doi-org.brum.beds.ac.uk/10.3390/digital1030011

Chicago/Turabian Style

Bhowmik, Kowshik, and Anca Ralescu. 2021. "Leveraging Vector Space Similarity for Learning Cross-Lingual Word Embeddings: A Systematic Review" Digital 1, no. 3: 145-161. https://0-doi-org.brum.beds.ac.uk/10.3390/digital1030011

Article Metrics

Back to TopTop