Joint Semantic Synthesis and Morphological Analysis of the Derived Word Ryan Cotterell Department of Computer Science Johns Hopkins University ryan.cotterell@jhu.edu Hinrich Schütze CIS LMU Munich inquiries@cislmu.org Abstract Much like sentences are composed of words, words themselves are composed of smaller units. For example, the English word questionably can be analyzed as question+able+ly. However, this structural decomposition of the word does not directly give us a semantic representation of the word’s meaning. Since morphology obeys the principle of compositionality, the semantics of the word can be systematically derived from the meaning of its parts. In this work, we propose a novel probabilistic model of word formation that captures both the analysis of a word w into its constituent segments and the synthesis of the meaning of w from the mean- ings of those segments. Our model jointly learns to segment words into morphemes and compose distributional semantic vectors of those morphemes. We experiment with the model on English CELEX data and German DErivBase (Zeller et al., 2013) data. We show that jointly modeling semantics increases both segmentation accuracy and morpheme F1 by between 3% and 5%. Additionally, we investigate different models of vector compo- sition, showing that recurrent neural networks yield an improvement over simple additive models. Finally, we study the degree to which the representations correspond to a linguist’s notion of morphological productivity. 1 Introduction In most languages, words decompose further into smaller units, termed morphemes. For example, the English word questionably can be analyzed as question+able+ly. This structural decomposition of the word, however, by itself is not a semantic rep- resentation of the word’s meaning;1 we further re- quire an account of how to synthesize the meaning from the decomposition. Fortunately, words—just like phrases—to a large extent obey the principle of compositionality: the semantics of the word can be systematically derived from the meaning of its parts.2 In this work, we propose a novel joint prob- abilistic model of word formation that captures both structural decomposition of a word w into its con- stituent segments and the synthesis of w’s meaning from the meaning of those segments. Morphological segmentation is a structured pre- diction task that seeks to break a word up into its constituent morphemes. The output segmentation has been shown to aid a diverse set of applications, such as automatic speech recognition (Afify et al., 2006), keyword spotting (Narasimhan et al., 2014), machine translation (Clifton and Sarkar, 2011) and parsing (Seeker and Çetinoğlu, 2015). In contrast to much of this prior work, we focus on supervised segmentation, i.e., we provide the model with gold segmentations during training time. Instead of sur- 1There are many different linguistic and computational theo- ries for interpreting the structural decomposition of a word. For example, un- often signifies negation and its effect on semantics can then be modeled by theories based on logic. This work ad- dresses the question of structural decomposition and semantic synthesis in the general framework of distributional semantics. 2Morphological research in theoretical and computational linguistics often focuses on noncompositional or less com- positional phenomena—simply because compositional deriva- tion poses fewer interesting research problems. It is also true that—just as many frequent multiword units are not completely compositional—many frequent derivations (e.g., refusal, fit- ness) are not completely compositional. An indication that non- lexicalized derivations are usually compositional is the fact that standard dictionaries like OUP editors (2010) list derivational affixes with their compositional meaning, without a hedge that they can also occur as part of only partially compositional forms. See also Haspelmath and Sims (2013), §5.3.6. 33 Transactions of the Association for Computational Linguistics, vol. 6, pp. 33–48, 2018. Action Editor: Regina Barzilay. Submission batch: 5/2016; Revision batch: 10/2016; Published 1/2018. c©2018 Association for Computational Linguistics. Distributed under a CC-BY 4.0 license. face segmentation, our model performs canonical segmentation (Cotterell et al., 2016a; Cotterell et al., 2016b; Kann et al., 2016), i.e., it allows the induc- tion of orthographic changes together with the seg- mentation, which is not typical. For the example questionably, our model can restore the deleted char- acters le, yielding the canonical segments question, able and ly. In this work, our primary contribution lies in the integration of continuous semantic vec- tors into supervised morphological segmentation— we present a joint model of morphological analysis and semantic synthesis at the word-level. We experimentally investigate three novel aspects of our model. • First, we show that jointly modeling continu- ous representations of the semantics of mor- phemes and words allows us to improve mor- phological analysis. On the English portion of CELEX (Baayen et al., 1993), we achieve a 5 point improvement in segmentation accuracy and a 3 point improvement in morpheme F1. On the German DErivBase dataset we achieve a 3 point improvement in segmentation accu- racy and a 3 point improvement in morpheme F1. • Second, we explore improved models of vec- tor composition for synthesizing word mean- ing. We find a recurrent neural network im- proves over previously proposed additive mod- els. Moreover, we find that more syntactically oriented vectors (Levy and Goldberg, 2014a) are better suited for morphology than bag-of- word (BOW) models. • Finally, we explore the productivity of English derivational affixes in the context of distribu- tional semantics. 2 Derivational Morphology Two important goals of morphology, the linguistic study of the internal structure of words, are to de- scribe the relation between different words in the lexicon and to decompose them into morphemes, the smallest linguistic unit bearing meaning. Morphol- ogy can be divided into two types: inflectional and derivational. Inflectional morphology is the set of processes through which the word form outwardly displays syntactic information, e.g., verb tense. It follows that an inflectional affix typically neither changes the part-of-speech (POS) nor the semantics of the word. For example, the English verb to run takes various forms: run, runs, ran and running, all of which convey “moving by foot quickly”, but ap- pear in complementary syntactic contexts. Derivation deals with the formation of new words that have semantic shifts in meaning (often includ- ing POS) and is tightly intertwined with lexical se- mantics (Light, 1996). Consider the example of the English noun discontentedness, which is derived from the adjective discontented. It is true that both words share a close semantic relationship, but the transformation is clearly more than a simple inflec- tional marking of syntax. Indeed, we can go one step further and define a chain of words content 7→ contented 7→ discontented 7→ discontentedness. In the computational literature, derivational mor- phology has received less attention than inflectional. There are, however, two bodies of work on deriva- tion in computational linguistics. First, there is a series of papers that explore the relation between lexical semantics and derivation (Lazaridou et al., 2013; Zeller et al., 2014; Padó et al., 2015; Kisse- lew et al., 2015). All of these assume a gold mor- phological analysis and primarily focus on the ef- fect of derivation on distributional semantics. The second body of work, e.g., the unsupervised mor- phological segmenter MORFESSOR (Creutz and La- gus, 2007), does not deal with semantics and makes no distinction between inflectional and derivational morphology.3 Even though the boundary between inflectional and derivational morphology is a con- tinuum rather than a rigid divide (Haspelmath and Sims, 2013), there is still the clear distinction that derivation changes meaning whereas inflection does not. Our goal in this paper is to develop an account of how the meaning of a word form can be computed jointly, combining these two lines of work. Productivity and Semantic Coherence. We highlight two related issues in derivation that moti- vated the development of our model: productivity 3Narasimhan et al. (2015) also make no distinction between inflectional and derivational morphology, but their model is an exception in that it includes vector similarity as a semantic fea- ture. See §5 for discussion. 34 and semantic coherence. Roughly, a productive affix is one that can still actively be employed to form new words in a language. For example, the English nominalizing affix ness (red7→red+ness) can be attached to just about any adjective, including novel forms. In contrast, the archaic English nominal- izing affix th (dear7→dear+th, heal7→heal+th, steal7→steal+th) does not allow us to form new words such as cheapth. This is a crucial issue in derivational morphology since we would not in general want to analyze new words as having been formed from non-productive endings; e.g., we do not want to analyze hearth as hear+th (or wugth as wug+th). Relations such as those between heal and health are lexicalized since they no longer can be derived by productive processes (Bauer, 1983). Under a generative treatment (Chomsky, 1965) of morphology, productivity becomes a central no- tion since a grammar needs to account for active word formation processes in the language (Aronoff, 1976). Defining productivity precisely, however, is tricky; Aronoff (1976) writes, “one of the central mysteries of derivational morphology . . . [is that] . . . though many things are possible in morphology, some are more possible than others.” Nevertheless, speakers often have clear intuitions about which af- fixes in the language are productive.4 Related to productivity is the notion of seman- tic coherence. The principle of compositionality (Frege, 1892; Heim and Kratzer, 1998) applies to interpretation of words just as it does to phrases. In- deed, compositionality is often taken to be a sig- nal for productivity (Aronoff, 1976). When de- ciding whether to further decompose a word, ask- ing whether the parts sum up to the whole is of- ten a good indicator. In the case of questionably 7→ question+able+ly, the compositional meaning is “in a manner that could be questioned”, which corresponds to the meaning of the word. Contrast this with the word unquiet, which means “restless”, rather than “not quiet” and the compound blackmail, which does not refer to a letter written in black ink. The model we will describe in §3 is a joint model of both semantic coherence and segmentation; that 4It is also important to distinguish productivity from creativity—a non-rule-governed form of word formation (Lyons, 1977). As an example of creativity, consider the cre- ation of portmanteaux, e.g., dramedy and soundscape. is, an analysis is judged not only by character-level features, but also by the degree to which the word is semantically compositional. Implicit in such a treatment is the desire to only segment a word if the segmentation is derived from a productive process. While most prior work on morphological segmen- tation has not explicitly modeled productivity,5 we believe, from a computational modeling perspective, segmenting only productive affixes is preferable. This is analogous to the modeling of phrase compo- sitionality in embedding models, where it can be bet- ter to not further decompose noncompositional mul- tiword units like named entities and idiomatic ex- pressions; see, e.g., Mikolov et al. (2013b), Wang et al. (2014), Yin and Schütze (2015), Yaghoobzadeh and Schütze (2015), and Hashimoto and Tsuruoka (2016).6 In this paper, we refer to the semantic aspect of the model either as semantic synthesis or as coherence. These are two ways of looking at semantics that are related as follows. If the synthesis (i.e., composi- tion) of the meaning of the derived form from the meaning of its parts is a regular application of the linguistic rules of derivation, then the meaning so constructed is coherent. These are the cases where a joint model is expected to be beneficial for both segmentation and interpretation. 3 A Joint Model From an NLP perspective, canonical segmentation (Naradowsky and Goldwater, 2009; Cotterell et al., 2016b) is the task that seeks to algorithmically de- compose a word into its canonical sequence of mor- phemes. It is a version of morphological segmenta- tion that requires the learner to handle orthographic changes that take place during word formation. We believe this is a more natural formulation of mor- phological analysis—especially for the processing 5Note that segmenters such as MORFESSOR utilize the prin- ciple of minimum description length, which implicitly encodes productivity, in order to guide segmentation. 6As a reviewer points out, productivity of an affix and se- mantic coherence of the words formed from it are not perfectly aligned. Nonproductive affixes can produce semantically coher- ent words, e.g., warm 7→warm+th. Productive affixes can pro- duce semantically incoherent words, e.g., canny 7→un+canny. Again, this is analogous to multiword units. However, there is a strong correlation and our experiments show that relying on it gives good results. 35 un question able ly su ffi x su ffi x st em pr ef ix unquestionably + + + unquestionablely ⇡ surface form underlying form segmentation vector composition ta rg et Figure 1: A depiction of the joint model that makes the relation between the three factors and the observed sur- face form explicit. We show a simple additive model of composition for ease of explication. of derivational morphology—as it draws heavily on linguistic notions (see §2). The main innovation we present is the augmen- tation of canonical segmentation to take into ac- count semantic coherence and productivity. Con- sider the word hypercuriosity and its canonical seg- mentation hyper+curious+ity; this canonical seg- mentation seeks to decompose the word into its con- stituent morphemes and account for orthographic changes. This amounts to a structural decomposi- tion of the word, i.e., how do we break up the string of characters into chunks? This is similar to the de- composition of a sentence into a parse tree. How- ever, it is also natural to consider the semantic com- positionality of a word, i.e., how is the meaning of the word synthesized from the meaning of the indi- vidual morphemes? We consider both of these questions together in a single model, where we would like to place high probability on canonical segmentations that are also semantically coherent. Returning to hy- percuriosity, we could further decompose it into hyper+cure+ous+ity in analogy to, say, vice 7→ vicious. Nothing about the surface form of curi- ous alone gives us a strong cue that we should rule out the segmentation cure+ous. Turning to distri- butional semantics, however, it is the case that the contexts in which curious occurs are quite different from those in which cure occurs. This gives us a strong cue which segmentation is correct. Formally, given a word string w ∈ Σ∗, where Σ is a discrete alphabet of characters (in English this could be as simple as the 26 letter lowercase alpha- bet), and a word vector v ∈ V , where V is a set of low-dimensional word embeddings, we define the model as: p(v,s, l,u |w) = 1 Zθ(w) exp ( 1 2σ2 ||v −Cβ(s,l)||22 +f(s,l,u)>η + g(u,w)>ω ) . (1) This model is composed of three factors: composi- tion factor ( 1 2σ2 ||v−Cβ(s,l)||22), segmentation factor f and transduction factor g. The parameters of the model are θ = {β,η,ω}, the function Cβ composes morpheme vectors together, s is the segmentation, l is the labeling of the segments, u is the underlying representation and Zθ(w) is the partition function. Note that the conditional distribution p(v | s,l,u,w) is Gaussian distributed by construction. A visualiza- tion of our model is found in Figure 1. This model is a conditional random field (CRF) that is mixed, i.e., it is defined over both discrete and continuous random variables (Koller and Friedman, 2009). We restrict the range of u to be a subset of Σ|w|+k, where k is an insertion limit (Dreyer, 2011). In this work, we take k = 5. Explicitly, the partition function is defined as Zθ(w) = ∫ ∑ l′,s′,u′ exp ( 1 2σ2 ||v′ −Cβ(s′, l′)||22 + f(s′, l′,u′)>η + g(u′,w)>ω ) dv′, (2) which is guaranteed to be finite.7 A CRF is simply the globally renormalized prod- uct of several non-negative factors (Sutton and McCallum, 2006). Our model is composed of three: transduction, segmentation and composition factors—we describe each in turn. 3.1 Transduction Factor The first factor we consider is the transduction factor: exp ( g(u,w)>ω ) , which scores a surface 7Since we have capped the insertion limit, we have a finite number of values that u can take for any w. Thus, it follows that we have a finite number of canonical segmentations s. Hence we take a finite number of Gaussian integrals. These integrals all converge since we have fixed the covariance matrix as σ2I, which is positive definite. 36 representation (SR) w, the character string ob- served in raw text, and an underlying represen- tation (UR), a character string with orthographic processes reversed. The aim of this factor is to place high weight on good pairs, e.g., the pair (w=questionably,u=questionablely), so we can ac- curately restore character-level changes. We encode this portion of the model as a weighted finite-state machine for ease of computation. This factor generalizes probabilistic edit distance (Ristad and Yianilos, 1998) by looking at additional input and output context; see Cotterell et al. (2014) for de- tails. As mentioned above and in contrast to Cot- terell et al. (2014), we bound the insertion limit in the edit distance model.8 Computing the score be- tween two strings u and w requires a dynamic pro- gram that runs in O(|u|·|w|). This is a generalization of the forward algorithm for Hidden Markov Models (HMMs) (Rabiner, 1989). We employ standard feature templates for the task that look at features of edit operations, e.g., substi- tute i for y, in varying context granularities. See Cotterell et al. (2016b) for details. Recent work has also explored weighting of WFST arcs with scores computed by LSTMs (Hochreiter and Schmidhuber, 1997), obviating the need for human selection of feature templates (Rastogi et al., 2016). 3.2 Segmentation Factor The second factor is the segmentation factor: exp ( f(s,l,u)>η ) . The goal of this factor is to score a segmentation s of a UR u. In our example, it scores the input-output pair (u=questionablely, s=question+able+ly). It additionally scores a labeling of the segmentation. Our label set in this work is L = {stem, prefix, suffix}. The proper labeling of the segmentation above is l=question:stem+able:suffix+ly:suffix. The label- ing is critical for our composition functions Cβ (Cot- terell et al., 2015): which vectors are used depends on the label given to the segment; e.g., the vectors of the prefix “post” and the stem “post” are different. We can view this factor as an unnormalized first- 8As our transduction model is an unnormalized factor in a CRF, we do not require the local normalization discussed in Cotterell et al. (2014)—a weight on an edge may be any non- negative real number since we will renormalize later. The un- derlying model, however, remains the same. model composition function stem c = ∑N i=1 1li=stemm li si mult c = ⊙N i=1 m li si add c = ∑N i=1 m li si wadd c = ∑N i=1 αim li si fulladd c = ∑N i=1 Uim li si LDS hi = Xhi−1 + Umlisi RNN hi = tanh(Xhi−1 + Umlisi ) Table 1: Composition models Cβ(s,l) used in this and prior work. The representation of the word is hN for the dynamic and c for the non-dynamic models. Note that for the dynamic models h0 is a learned parameter. order semi-CRF (Sarawagi and Cohen, 2005). Com- putation of the factor again requires dynamic pro- gramming. The algorithm is a different generaliza- tion of the forward algorithm for HMMs, one that extends it to the semi-Markov case. This algorithm runs in O(|u|2 · |L|2). Features. We again use standard feature templates for the task. We create atomic indicator features for the individual segments. We then conjoin the atomic features with left and right context features as well as the label to create more complex feature templates. We also include transition features that fire on pairs of sequential labels. See Cotterell et al. (2015) for details. Recent work has also showed that a neural parameterization can remove the need for manual feature design (Kong et al., 2016). 3.3 Composition Factor The composition factor takes the form of an unnormalized multivariate Gaussian density: exp ( 1 2σ2 ||v −Cβ(s,l)||22 ) , where the mean is computed by the (potentially non-linear) compo- sition function (See Table 1) and the covariance matrix σ2I is a diagonal matrix. The goal of the composition function Cβ(s,l) is to stitch together morpheme embeddings to approximate the vector of the entire word. The simplest form of the composition function Cβ(s,l) is add, an additive model of the morphemes. See Table 1: each vector mlisi refers to a morpheme- 37 specific, label-dependent embedding. If li = stem, then si represents a stem morpheme. Given that our segmentation is canonical, an si that is a stem gen- erally itself is an entry in the lexicon and v(si) ∈ V . If v(si) 6∈ V , then we set v(si) to 0.9 We optimize over vectors with li ∈ {prefix, suffix} as they corre- spond to bound morphemes. We also consider a more expressive composition model, a recurrent neural network (RNN). Let N be the number of segments. Then Cβ(s,l) = hN where hi is a hidden vector, defined by the re- cursion:10 hi = tanh ( Xhi−1 + Um li si ) (Elman, 1990). Again, we optimize the morpheme embed- dings mlisi only when li 6= stem along with the other parameters of the RNN, i.e., the matrices U and X. 4 Inference and Learning Exact inference is intractable since we allow ar- bitrary segment-level features on the canonicalized word forms u. Since the semi-CRF factor has fea- tures that fire on substrings, we would need a dy- namic programming state for each substring of each of the exponentially many settings of u; this breaks the dynamic program. We thus turn to approximate inference through an importance sampling routine (Rubinstein and Kroese, 2011). 4.1 Inference by Importance Sampling Rather than considering all underlying orthographic forms u and segmentations s, we sample from a tractable proposal distribution q—a distribution over canonical segmentations. In the following equations we omit the dependence on w for notational brevity and define h(l,s,u) = f(s,l,u) + g(u,w). Cru- cially, the partition function Zθ(w) is not a function of parameter subvector β and its gradient with re- 9This is not changed in training, so all such v(si) are 0 in the final model. Clearly, this could be improved in future work as a reviewer points out, e.g., by setting such v(si) to an average of a suitable chosen set of known word vectors. 10We do not explore more complex RNNs, e.g., LSTMs (Hochreiter and Schmidhuber, 1997) and GRUs (Cho et al., 2014a) as words in our data have ≤7 morphemes. These archi- tectures make the learning of long distance dependencies eas- ier, but are no more powerful than an Elman RNN, at least in theory. Note that perhaps if applied to languages with richer derivational morphology than English, considering more com- plex neural architectures would make sense. spect to β is 0.11 Recall that computing the gradi- ent of the log-partition function is equivalent to the problem of marginal inference (Wainwright and Jor- dan, 2008). We derive our estimator as follows: ∇θ log Z = E (l,s,u)∼p [h(l,s,u)] (3) = ∑ l,s,u p(l,s,u)h(l,s,u) (4) = ∑ l,s,u q(l,s,u) q(l,s,u) p(l,s,u)h(l,s,u) (5) = E (l,s,u)∼q [ p(l,s,u) q(l,s,u) h(l,s,u) ] , (6) where we have omitted the dependence on w (which we condition on) and v (which we marginalize out). So long as q has support everywhere p does (i.e., p(l,s,u) > 0 ⇒ q(l,s,u) > 0), the estimate is un- biased. Unfortunately, we can only efficiently com- pute p(l,s,u) up to a constant factor, p(l,s,u) = p̄(l,s,u)/Z′θ(w). Thus, we use the indirect impor- tance sampling estimator, 1 ∑M i=1 w (i) M∑ i=1 w(i)h(l(i),s(i),u(i)), (7) where (l(1),s(1),u(1)) . . . (l(M),s(M),u(M)) i.i.d.∼ q and importance weights w(i) are defined as: w(i) = p̄(l(i),s(i),u(i)) q(l(i),s(i),u(i)) . (8) This indirect estimator is biased, but consistent.12 Proposal Distribution. The success of impor- tance sampling depends on the choice of a “good” proposal distribution, i.e., one that ideally is close to p. Since we are fully supervised at training time, we have the option of training locally normalized distributions for the individual components. Con- cretely, we train two proposal distributions q1(u | w) and q2(l,s | u) that take the form of a WFST and a semi-CRF, respectively, using features identical 11The subvector β is responsible for computing only the mean of the Gaussian factor and thus has no impact on its nor- malization coefficient (Murphy, 2012). 12Informally, the indirect importance sampling estimate con- verges to the true expectation as M → ∞ (the definition of statistical consistency). 38 to the joint model. Each of these distributions is tractable—we can compute the marginals with dy- namic programming and thus sample efficiently. To draw samples (l,s,u) ∼ q, we sample sequentially from q1 and then q2, conditioned on the output of q1. 4.2 Learning We optimize the log-likelihood of the model using ADAGRAD (Duchi et al., 2011), which is SGD with a special per-parameter learning rate. The full gra- dient of the objective for one training example is: ∇θ log p(v,s, l,u | w) = f(s,l,u)> + g(u,w)> − 1 σ2 (v −Cβ(s,l))∇θCβ(s,l) −∇θ log Zθ(w), (9) where we use the importance sampling algorithm described in §4.1 to approximate the gradient of the log-partition function, following Bengio and Senecal (2003). Note that ∇θCβ(s,l) depends on the composition function used. In the most com- plicated case when Cβ is a RNN, we can com- pute ∇βCβ(s,l) efficiently with backpropagation through time (Werbos, 1990). We take M = 10 im- portance samples; using so few samples can lead to a poor estimate of the gradient, but for our application it suffices. We employ L2 regularization. 4.3 Decoding Decoding the model is also intractable. To approxi- mate the solution, we again employ importance sam- pling. We take M =10,000 importance samples and select the highest weighted sample. 5 Related Work The idea that vector semantics is useful for mor- phological segmentation is not new. Count vectors (Salton, 1971; Turney and Pantel, 2010) have been shown to be beneficial in the unsupervised induction of morphology (Schone and Jurafsky, 2000; Schone and Jurafsky, 2001). Embeddings were shown to act similarly (Soricut and Och, 2015). Our method differs from this line of research in two key ways. (i) We present a probabilistic model of the pro- cess of synthesizing the word’s meaning from the meaning of its morphemes. Prior work was ei- ther not probabilistic or did not explicitly model morphemes. (ii) Our method is supervised and fo- cuses on derivation. Schone and Jurafsky (2000) and Soricut and Och (2015), being fully unsupervised, do not distinguish between inflection and deriva- tion and Schone and Jurafsky (2001) focus on in- flection. More recently, Narasimhan et al. (2015) look at the unsupervised induction of “morpholog- ical chains” with semantic vectors as a crucial fea- ture. Their goal is to jointly figure out an ordering of word formation and a morphological segmenta- tion, e.g., play7→playful7→playfulness. While it is a rich model like ours, theirs differs in that it is un- supervised and uses vectors as features, rather than explicitly treating vector composition. All of the above work focuses on surface segmentation and not canonical segmentation, as we do. A related line of work that has different goals con- cerns morphological generation. Two recent papers that address this problem using deep learning are Faruqui et al. (2016a) and Faruqui et al. (2016b). In an older line of work, Yarowsky and Wicen- towski (2000) and Wicentowski (2002) exploit log frequency ratios of inflectionally related forms to tease apart that, e.g., the past tense of sing is not singed, but instead sang. Related work by Dreyer and Eisner (2011) uses a Dirichlet process to model a corpus as a “mixture of a paradigm”, allowing for the semi-supervised incorporation of distribu- tional semantics into a structured model of inflec- tional paradigm completion. Our work is also related to recent attempts to in- tegrate morphological knowledge into general em- bedding models. For example, Botha and Blun- som (2014) train a log-bilinear language model that models the composition of morphological structure. Likewise, Luong et al. (2013) train a recursive neural network (Goller and Küchler, 1996) over a heuristi- cally derived tree structure to learn morphological composition over continuous vectors. Our work is different in that we learn a joint model of segmen- tation and composition. Moreover, supervised mor- phological analysis can drastically outperform unsu- pervised analysis (Ruokolainen et al., 2013). Early work by Kay (1977) can be interpreted as finite-state canonical segmentation, but it neither ad- dresses nor experimentally evaluates the question of joint modeling of morphological analysis and se- mantic synthesis. Moreover, we may view canoni- 39 dev test Model Acc F1 Edit Acc F1 Edit E N Semi-CRF (Baseline) 0.55 (.018) 0.75 (.014) 0.80 (.043) 0.54 (.018) 0.75 (.014) 0.78 (.034) Joint (Baseline) 0.77 (.011) 0.87 (.007) 0.41 (.029) 0.77 (.013) 0.87 (.007) 0.43 (.029) Joint + Vec (This Work) 0.83 (.014) 0.91 (.008) 0.31 (.019) 0.82 (.020) 0.90 (.011) 0.32 (.038) Joint + UR (Oracle) 0.94 (.015) 0.96 (.009) 0.07 (.016) 0.94 (.011) 0.96 (.007) 0.07 (.011) Joint + UR + Vec (Oracle) 0.95 (.011) 0.97 (.007) 0.05 (.013) 0.95 (.023) 0.97 (.006) 0.05 (.025) D E Semi-CRF (Baseline) 0.39 (.062) 0.68 (.039) 1.15 (.230) 0.39 (.058) 0.68 (.042) 1.14 (.240) Joint (Baseline) 0.79 (.107) 0.88 (.069) 0.40 (.313) 0.79 (.099) 0.87 (.063) 0.41 (.282) Joint + Vec (This Work) 0.82 (.102) 0.90 (.067) 0.33 (.312) 0.82 (.096) 0.90 (.061) 0.33 (.282) Joint + UR (Oracle) 0.86 (.108) 0.90 (.070) 0.25 (.288) 0.86 (.100) 0.90 (.064) 0.25 (.268) Joint + UR + Vec (Oracle) 0.87 (.106) 0.92 (.069) 0.20 (.285) 0.88 (.096) 0.93 (.062) 0.19 (.263) Table 2: Results for the canonical morphological segmentation task on English and German. Standard deviation is given in parentheses. We compare against two baselines that do not make use of semantic vectors: (i) “Semi-CRF (baseline)”, a semi-CRF that cannot account for orthographic changes and (ii) “Joint (Baseline)”, a version of our joint model without vectors. We also compare against an oracle version with access to gold URs (“Joint + UR (Oracle)”, “Joint + UR + Vec (Oracle)”), revealing that the toughest part of the canonical segmentation task is reversing the orthographic changes. calization as an orthographic analogue to phonology. On this interpretation, the finite-state systems of Ka- plan and Kay (1994), which computationally apply SPE-style phonological rules (Chomsky and Halle, 1968), may be run backwards to get canonical un- derlying forms. 6 Experiments and Results We conduct experiments on English and German derivational morphology. We analyze our joint model’s ability to segment words into their canoni- cal morphemes as well as its ability to composition- ally derive vectors for new words. Finally, we ex- plore the relationship between distributional seman- tics and morphological productivity. For English, we use the pretrained vectors of Levy and Goldberg (2014a) for all experiments. For German, we train word2vec skip-gram vectors on the German Wikipedia. We first describe our En- glish dataset, the subset of the English portion of the CELEX lexical database (Baayen et al., 1993) that was selected by Lazaridou et al. (2013); the dataset contains 10,000 forms. This allows for com- parison with previously proposed methods. We make two modifications. (i) Lazaridou et al. (2013) make the two-morpheme assumption: every word is composed of exactly two morphemes. In general, this is not true, so we further segment all complex words in the corpus. For example, friendless+ness is further segmented into friend+less+ness. To nevertheless allow for fair comparison, we pro- vide versions of our experiments with and without the two-morpheme assumption where appropriate. (ii) Lazaridou et al. (2013) only provide a single train/test split. As we require a held-out develop- ment set for hyperparameter tuning, we randomly allocate a portion of the training data to select the hyperparameters and then retrain the model using these parameters on the original train split. We also report 10-fold cross validation results in addition to Lazaridou et al.’s train/test split. Our German dataset is taken from Zeller et al. (2013) and is described in Cotterell et al. (2016b). It, again, consists of 10,000 derivational forms. We report results on 10-fold cross validation. 6.1 Experiment 1: Canonical Segmentation For our first experiment, we test whether jointly modeling the continuous representations allows us to segment words more accurately. We assume that we are given an embedding for the target word. We estimate the model p(v,s, l,u | w) as described in §4 with L2 regularization λ||θ||22. To evaluate, we decode the distribution p(s,l,u | v,w). We per- form approximate MAP inference with importance sampling—taking the sample with the highest score. 40 EN DE BOW2 BOW5 DEPs SG dev test dev test dev test dev test or ac le stem .403 .402 .374 .376 .422 .422 .400 .405 add .635 .635 .541 .542 .787 .785 .712 .711 LDS .660 .660 .566 .568 .806 .804 .717 .718 RNN .660 .660 .565 .567 .807 .806 .707 .712 jo in t stem .399 .400 .371 .372 .411 .412 .394 .398 add .625 .625 .524 .525 .782 .781 .705 .704 LDS .648 .648 .547 .547 .799 .797 .712 .711 RNN .649 .647 .547 .546 .801 .799 .706 .708 ch ar GRU .586 .585 .452 .452 .769 .768 .675 .667 LSTM .586 .586 .455 .455 .768 .767 .677 .666 Table 3: Vector approximation (measured by mean co- sine similarity) both with (“oracle”) and without (“joint”, “char”) gold morphology. Surprisingly, joint models are close in performance to models with gold morphology. In these experiments, we use the RNN with the de- pendency vectors, the combination of which per- forms best on vector approximation in §6.2. We follow the experimental design of Cotterell et al. (2016b). We compare against two base- lines (marked “Baseline” in Table 2): (i) a “Semi- CRF” segmenter that cannot account for ortho- graphic changes and (ii) the full “Joint” model of Cotterell et al. (2016b).13 We additionally consider an “Oracle” setting, where we give the model the gold underlying orthographic form (“UR”) at both training and test time. This gives us insight into the performance of the transduction factor of our model, i.e., how much could we benefit from a richer model. Our hyperparameters are (i) the regularization coefficient λ and (ii) σ2, the variance of the Gaussian factor. We use grid search to tune them: λ ∈ {0.0, 101, 102, 103, 104, 105}, σ2 ∈ {0.25, 0.5, 0.75, 1.0}. Metrics. We use three metrics to evaluate segmen- tation accuracy. Note that the evaluation of canon- ical segmentation is hard since a system may re- turn a sequence of morphemes whose concatenation is not the same length as the concatenation of the gold morphemes. This rules out metrics for surface segmentation like border F1 (Kurimo et al., 2010), which require the strings to be of the same length. We now define the metrics. (i) Segmentation accuracy measures whether every single canonical morpheme in the returned sequence is correct. It is inflexible: closer answers are penalized the same as 13i.e., a model without the Gaussian factor that scores vectors. more distant answers. (ii) Morpheme F1 (van den Bosch and Daelemans, 1999) takes the predicted se- quence of canonical morphemes, turns it into a set, computes precision and recall in the standard way and based on that then computes F1. This metric gives credit if some of the canonical morphemes were correct. (iii) Levenshtein distance joins the canonical segments with a special symbol # into a single string and computes the Levenshtein distance between predicted and gold strings. Discussion. Results in Table 2 show that jointly modeling semantic coherence improves our ability to analyze words. For test, our proposed joint model (“This Work”) outperforms the baseline supervised canonical segmenter, which is state-of-the-art for the task, by .05 (resp. .03) on accuracy and .03 (resp. .03) on F1 for English (resp. German). We also find that when we give the joint model an oracle UR the vectors generally help less: .01 (resp. .02) on ac- curacy and .01 (resp. .03) on F1 for English (resp. German). This indicates that the chief boon the vec- tor composition factor provides lies in selection of an appropriate UR. Moreover, the up to .15 differ- ence in English between systems with and without the oracle UR suggests that reversing orthographic changes is a particularly difficult part of the task, at least for English. 6.2 Experiment 2: Vector Approximation We adopt the experimental design of Lazaridou et al. (2013). Its aim is to approximate a vector of a derivationally complex word using a learned model of composition. As Lazaridou et al. (2013) assume a gold morphological analysis, we compare two set- tings: (i) oracle morphological analysis and (ii) in- ferred morphological analysis. To the best of our knowledge, (ii) is a novel experimental condition that no previous work has addressed. We consider four composition models (See Ta- ble 1). (i) stem, using just the stem vector. This baseline tells us what happens if we make the incor- rect assumption that derivation behaves like inflec- tion and is not meaning-changing. (ii) add, a purely additive model. This is arguably the simplest way of combining the vectors of the morphemes. (iii) LDS, a linear dynamical system. This is arguably the sim- plest sequence model. (iv) A (simple) RNN. Recur- 41 all HR LR -less in- un- L az ar id ou stem .47 .52 .32 .22 .39 .33 mult .39 .43 .28 .23 .34 .33 dil. .48 .53 .33 .30 .45 .41 wadd .50 .55 .38 .24 .40 .34 fulladd .56 .61 .41 .38 .47 .44 lexfunc .54 .58 .42 .44 .45 .46 B O W 2 stem .43 .44 .38 .32 .43 .51 add .65 .67 .61 .60 .64 .67 LDS .67 .69 .62 .61 .66 .67 RNN .67 .69 .60 .60 .65 .66 c-GRU .59 .60 .55 .59 .55 .57 c-LSTM .52 .53 .50 .55 .50 .50 B O W 5 stem .40 .43 .33 .27 .37 .46 add .56 .59 .51 .46 .55 .59 LDS .58 .61 .51 .48 .57 .60 RNN .58 .61 .50 .48 .56 .58 c-GRU .45 .47 .42 .42 .43 .45 c-LSTM .46 .47 .43 .43 .45 .46 D E P s stem .46 .45 .49 .38 .57 .67 add .79 .79 .77 .78 .80 .80 LDS .80 .81 .77 .79 .81 .81 RNN .81 .82 .77 .79 .80 .81 c-GRU .75 .76 .72 .78 .74 .75 c-LSTM .75 .76 .71 .77 .72 .73 Table 4: Vector approximation (measured by mean cosine similarity) with gold morphology on the train/test split of Lazaridou et al. (2013). HR/LR = high/low-relatedness words. See Lazaridou et al. (2013) for details. rent neural networks are currently the most widely used nonlinear sequence model and simple RNNs are the simplest such models. Part of the motivation for considering a richer class of models lies in our removal of the two- morpheme assumption. Indeed, it is unclear that the wadd and fulladd models (Mitchell and Lapata, 2008) are useful models in the general case of multi- morphemic words—the weights are tied by position, i.e., the first morpheme’s vector (be it a prefix or stem) is always multiplied by the same matrix. Comparison with Lazaridou et al. To compare with Lazaridou et al. (2013), we use their exact train/test split. Those results are reported in Table 4. This dataset enforces that all words are composed of exactly two morphemes. Thus, a word like unques- tionably is segmented as un+questionably, with- out further decomposition. The vectors employed by Lazaridou et al. (2013) are high-dimensional count vectors derived from lemmatized and POS tagged text with a before-and-after window of size 2. They then apply pointwise mutual informa- tion (PMI) weighting and dimensionality reduction by non-negative matrix factorization. In contrast, we employ WORD2VEC (Mikolov et al., 2013a), a model that is also interpretable as the factorization of a PMI matrix (Levy and Goldberg, 2014b). We con- sider three WORD2VEC models: two bag-of-word (BOW) models with before-and-after windows of size 2 and 5 and DEPs (Levy and Goldberg, 2014a), a dependency-based model whose context is derived from dependency parses rather than BOW. In general, the results indicate that the key to better vector approximation is not a richer model of composition, but rather lies in the vectors them- selves. We find that our best model, the RNN, only marginally edges out the LDS. Additionally, looking at the “all” column and the DEPs vectors, the sim- ple additive model is only ≤.02 lower than LDS. In comparison, we observe large differences between the vectors. The RNN+DEPs model is .23 bet- ter than the BOW5 models (.81 vs. .58), .14 better than the BOW2 models (.81 vs. .67) and .25 bet- ter than Lazaridou et al.’s best model (.81 vs. .56). A wider context for BOW (5 instead of 2) yields worse results. This suggests that syntactic infor- mation or at least positional information is neces- sary for improved models of morpheme composi- tion. The test vectors are annotated for relatedness, which is a proxy for semantic coherence. HR (high- relatedness) words were judged to be more compo- sitional than LR (low-relatedness) words. Character-Level Neural Retrofitting. As a fur- ther strong baseline, we consider a retrofitting (Faruqui et al., 2015) approach based on character- level recurrent neural networks. Recently, running a recurrent net over the character stream has become a popular way of incorporating subword information into a model—empirical gains have been observed in a diverse set of NLP tasks: POS tagging (dos Santos and Zadrozny, 2014; Ling et al., 2015), pars- ing (Ballesteros et al., 2015) and language modeling 42 (Kim et al., 2016). To the best of our knowledge, character-level retrofitting is a novel approach. Given a vector v for a word form w, we seek a function to minimize the following objective 1 2 ||v −hN||22, (10) where hN is the final hidden state of a recurrent neu- ral architecture, i.e., hi = σ(Ahi−1 + Bwi), (11) where σ is a non-linearity and wi is the ith char- acter in w, hi−1 is the previous hidden state and A and B are matrices. While we have defined the architecture for a vanilla RNN, we experiment with two more advanced recurrent architectures: GRUs (Cho et al., 2014b) and LSTMs (Hochreiter and Schmidhuber, 1997) as well as deep variants (Sutskever et al., 2014; Gillick et al., 2016; Firat et al., 2016). Importantly, this model has no knowledge of morphology—it can only rely on representations it extracts from the characters. This gives us a clear ablation on the benefit of adding structured morpho- logical knowledge. We optimize the depth and the size of the hidden units on development data using a coarse-grained grid search. We found a depth of 2 and hidden units of size 100 (in both LSTM and GRU) performed best. We trained all models for 100 iterations of Adam (Kingma and Ba, 2015) with L2 regularization with regularization coefficient 0.01. Table 4 shows that the two character-level mod- els (“c-GRU” and “c-LSTM”) perform much worse than our models. This indicates that supervised mor- phological analysis produces higher-quality vector representations than “knowledge-poor” character- level models. However, we note that these character-level models have fewer parameters than our morpheme-level models—there are many more morphemes in a languages than characters. Oracle Morphology. In general, the two- morpheme assumption is incorrect. We consider an expanded setting of Lazaridou et al. (2013)’s task, in which we fully decompose the word, e.g., unquestionably7→un+question+able+ly. These results are reported in Table 3 (top block, “oracle”). We report mean cosine similarity. Standard devia- tions s for 10-fold cross-validation (not shown) are small (≤ .012) with two exceptions: s = .044 for the DEPs-joint-stem results (.411 and .412). The multi-morphemic results mirror those of the bi-morphemic setting of Lazaridou et al. (2013). (i) RNN+DEPs attains an average cosine similarity of around .80 for English. Numbers for German are lower, around .70. (ii) The RNN only marginally edges out LDS for English and is slightly worse for German. Again, this is not surprising as we are mod- eling short sequences. (iii) Certain embeddings lend themselves more naturally to derivational composi- tionality: BOW2 is better than BOW5, DEPs is the clear winner. Inferred Morphology. The final setting we con- sider is the vector approximation task without gold morphology. In this case, we rely on the full joint model p(v,s, l,u | w). At evaluation, we are in- terested in the marginal distribution p(v | w) =∑ s,l,u p(v,s, l,u | w). We then use importance sampling to approximate the mean of this marginal distribution as the predicted embedding, i.e., v̂ = ∫ vp(v | w)dv (12) ≈ 1∑M i=1 w (i) M∑ i=1 w(i)Cβ(l(i),s(i)), (13) where w(i) are the importance weights defined in Equation 8 and l(i) and s(i) are the ith sampled la- beling and segmentation, respectively. Discussion. Surprisingly, Table 3 (joint) shows that relying on the inferred morphology does not drastically affect the results. Indeed, we are often within .01 of the result with gold morphology. Our method can be viewed as a retrofitting procedure (Faruqui et al., 2015), so this result is useful: it indi- cates that joint semantic synthesis and morphologi- cal analysis produces high-quality vectors. 6.3 Experiment 3: Derivational Productivity We now delve into the relation between distribu- tional semantics and morphological productivity. The extent to which jointly modeling semantics aids morphological analysis will be determined by the in- herent compositionality of the words within the vec- tor space. We break down our results on the vector approximation task with gold morphology using the 43 -ly -ness -ize -able in- -ful -ous -less -ic -ist -y un- -ity re- -ion -ment -al -er Affix 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 A v er a g e C o si n e S im il a ri ty Figure 2: The boxplot breaks down the cosine similarity between the approximated vector and the target vector by affix (using gold morphology). We have ordered the affixes such that the better approximated vectors are on the left. dependency vectors and the RNN composer in Fig- ure 2 by selected affixes. We observe a wide range of scores: the most compositional ending ly gives rise to cosine similarities that are 20 points higher than those of the least compositional er. On the left end of Figure 2 we see extremely pro- ductive suffixes. The affix ize is used productively with relatively obscure words in the sciences, e.g., Rao-Blackwellize. Likewise, the affix ness can be applied to almost any adjective without restriction, e.g., Poissonness ‘degree to which data have a Pois- son distribution’. On the right end, we find -ment, -er and re-. The affix -ment is borderline productive (Bauer, 1983)—modern English tends to form novel nominalizations with ness or ity. More interesting are re- and er, both of which are very productive in English. For er, many of the words bringing down the average are simply non-compositional. For ex- ample, homer ‘homerun in baseball’ is not derived from home+er—this is an error in data. We also see examples like cutter. It has a compositional read- ing (e.g., “box cutter”), but also frequently occurs in the non-compositional meaning ‘type of boat’. Fi- nally, proper nouns like Homer and Turner end in er and in our experiments we computed vectors for lowercased words. The affix re- similarly has a large number of non-compositional cases, e.g., remove, relocate, remark. Indeed, to get the compositional reading of remove, the first syllable (rather than the second) is typically stressed to emphasize the prefix. We finally note several limitations of this exper- iment. (i) The ability of our models—even the re- current neural network—to model transformations between vectors is limited. (ii) Our vectors are far from perfect; e.g., sparseness in the training data af- fects quality and some of the words in our corpus are rare. (iii) Semantic coherence is not the only crite- rion for productivity. An example is -th in English. As noted earlier, it is compositional in a word like warmth, but it cannot be used to form new words. 7 Conclusion We have presented a model of the semantics and structure of derivationally complex words. To the best of our knowledge, this is the first attempt to jointly consider, within a single model, (i) the mor- phological decomposition of the word form and (ii) the semantic coherence of the resulting anal- ysis. We found that directly modeling coherence increases segmentation accuracy, improving over a strong baseline. Also, our models show state-of-the- art performance on the derivational vector approxi- mation task introduced by Lazaridou et al. (2013). Future work will focus on the extension of the method to more complex instances of derivational morphology, e.g., compounding and reduplication, and on the extension to additional languages. We also plan to explore the relation between derivation and distributional semantics in greater detail. 44 Acknowledgments. The first author was sup- ported by a DAAD Long-Term Research Grant and an NDSEG fellowship and the second by a Volk- swagenstiftung Opus Magnum grant. We would also like to thank action editor Regina Barzilay for suggesting several changes we incorporated into the work and the three anonymous reviewers. References Mohamed Afify, Ruhi Sarikaya, Hong-Kwang Jeff Kuo, Laurent Besacier, and Yuqing Gao. 2006. On the use of morphological analysis for dialectal Arabic speech recognition. In Ninth International Conference on Spoken Language Processing. Mark Aronoff. 1976. Word Formation in Generative Grammar. MIT Press. Harald Baayen, Richard Piepenbrock, and Hedderik van Rijn. 1993. The CELEX lexical data base on CD- ROM. Miguel Ballesteros, Chris Dyer, and Noah A. Smith. 2015. Improved transition-based parsing by model- ing characters instead of words with LSTMs. In Pro- ceedings of the 2015 Conference on Empirical Meth- ods in Natural Language Processing, pages 349–359, Lisbon, Portugal, September. Association for Compu- tational Linguistics. Laurie Bauer. 1983. English Word-Formation. Cam- bridge University Press. Yoshua Bengio and Jean-Sébastien Senecal. 2003. Quick training of probabilistic neural nets by impor- tance sampling. In Proceedings of the Ninth Interna- tional Conference on Artificial Intelligence and Statis- tics. Jan A. Botha and Phil Blunsom. 2014. Compositional morphology for word representations and language modelling. In International Conference on Machine Learning, pages 1899–1907. Kyunghyun Cho, Bart van Merriënboer, Dzmitry Bah- danau, and Yoshua Bengio. 2014a. On the properties of neural machine translation: Encoder–decoder ap- proaches. Workshop On Syntax, Semantics and Struc- ture in Statistical Translation. Kyunghyun Cho, Bart Van Merriënboer, Caglar Gul- cehre, Dzmitry Bahdanau, Fethi Bougares, Holger Schwenk, and Yoshua Bengio. 2014b. Learning phrase representations using RNN encoder-decoder for statistical machine translation. In Conference on Empirical Methods in Natural Language Processing. Noam Chomsky and Morris Halle. 1968. The sound pat- tern of English. Harper & Row. Noam Chomsky. 1965. Aspects of the Theory of Syntax. MIT Press. Ann Clifton and Anoop Sarkar. 2011. Combin- ing morpheme-based machine translation with post- processing morpheme prediction. In Proceedings of the 49th Annual Meeting of the Association for Com- putational Linguistics: Human Language Technolo- gies, pages 32–42, Portland, Oregon, USA, June. As- sociation for Computational Linguistics. Ryan Cotterell, Nanyun Peng, and Jason Eisner. 2014. Stochastic contextual edit distance and probabilistic FSTs. In Proceedings of the 52nd Annual Meeting of the Association for Computational Linguistics (Vol- ume 2: Short Papers), pages 625–630, Baltimore, Maryland, June. Association for Computational Lin- guistics. Ryan Cotterell, Thomas Müller, Alexander Fraser, and Hinrich Schütze. 2015. Labeled morphological seg- mentation with semi-Markov models. In Proceed- ings of the Nineteenth Conference on Computational Natural Language Learning, pages 164–174, Beijing, China, July. Association for Computational Linguis- tics. Ryan Cotterell, Arun Kumar, and Hinrich Schütze. 2016a. Morphological segmentation inside-out. In Proceedings of the 2016 Conference on Empiri- cal Methods in Natural Language Processing, pages 2325–2330, Austin, Texas, November. Association for Computational Linguistics. Ryan Cotterell, Tim Vieira, and Hinrich Schütze. 2016b. A joint model of orthography and morphological seg- mentation. In Proceedings of the 2016 Conference of the North American Chapter of the Association for Computational Linguistics: Human Language Tech- nologies, pages 664–669, San Diego, California, June. Association for Computational Linguistics. Mathias Creutz and Krista Lagus. 2007. Unsupervised models for morpheme segmentation and morphology learning. ACM Transactions on Speech and Language Processing, 4(1):3. Cı́cero Nogueira dos Santos and Bianca Zadrozny. 2014. Learning character-level representations for part-of- speech tagging. In International Conference on Ma- chine Learning, pages 1818–1826. Markus Dreyer and Jason Eisner. 2011. Discover- ing morphological paradigms from plain text using a Dirichlet process mixture model. In Proceedings of the 2011 Conference on Empirical Methods in Natural Language Processing, pages 616–627. Association for Computational Linguistics, July. Markus Dreyer. 2011. A Non-parametric Model for the Discovery of Inflectional Paradigms from Plain Text using Graphical Models over Strings. Ph.D. thesis, Johns Hopkins University. John Duchi, Elad Hazan, and Yoram Singer. 2011. Adaptive subgradient methods for online learning and 45 stochastic optimization. Journal of Machine Learning Research, 12:2121–2159. Jeffrey L. Elman. 1990. Finding structure in time. Cog- nitive Science, 14(2):179–211. Manaal Faruqui, Jesse Dodge, Sujay Kumar Jauhar, Chris Dyer, Eduard Hovy, and Noah A. Smith. 2015. Retrofitting word vectors to semantic lexicons. In Pro- ceedings of the 2015 Conference of the North Ameri- can Chapter of the Association for Computational Lin- guistics: Human Language Technologies, pages 1606– 1615, Denver, Colorado, May–June. Association for Computational Linguistics. Manaal Faruqui, Ryan McDonald, and Radu Soricut. 2016a. Morpho-syntactic lexicon generation using graph-based semi-supervised learning. Transactions of the Association for Computational Linguistics, 4:1– 16. Manaal Faruqui, Yulia Tsvetkov, Graham Neubig, and Chris Dyer. 2016b. Morphological inflection gen- eration using character sequence to sequence learn- ing. In Proceedings of the 2016 Conference of the North American Chapter of the Association for Com- putational Linguistics: Human Language Technolo- gies, pages 634–643, San Diego, California, June. As- sociation for Computational Linguistics. Orhan Firat, Kyunghyun Cho, and Yoshua Bengio. 2016. Multi-way, multilingual neural machine translation with a shared attention mechanism. In Proceedings of the 2016 Conference of the North American Chap- ter of the Association for Computational Linguistics: Human Language Technologies, pages 866–875, San Diego, California, June. Association for Computa- tional Linguistics. Gottlob Frege. 1892. Über Begriff und Gegenstand. Vierteljahresschrift für wissenschaftliche Philosophie, 16:192–205. Dan Gillick, Cliff Brunk, Oriol Vinyals, and Amarnag Subramanya. 2016. Multilingual language process- ing from bytes. In Proceedings of the 2016 Confer- ence of the North American Chapter of the Associa- tion for Computational Linguistics: Human Language Technologies, pages 1296–1306, San Diego, Califor- nia, June. Association for Computational Linguistics. Christoph Goller and Andreas Küchler. 1996. Learning task-dependent distributed representations by back- propagation through structure. In IEEE International Conference on Neural Networks. Kazuma Hashimoto and Yoshimasa Tsuruoka. 2016. Adaptive joint learning of compositional and non- compositional phrase embeddings. In Proceedings of the 54th Annual Meeting of the Association for Com- putational Linguistics (Volume 1: Long Papers), pages 205–215, Berlin, Germany, August. Association for Computational Linguistics. Martin Haspelmath and Andrea Sims. 2013. Under- standing morphology. Routledge. Irene Heim and Angelika Kratzer. 1998. Semantics in Generative Grammar. Blackwell. Sepp Hochreiter and Jürgen Schmidhuber. 1997. Long short-term memory. Neural Computation, 9(8):1735– 1780. Katharina Kann, Ryan Cotterell, and Hinrich Schütze. 2016. Neural morphological analysis: Encoding- decoding canonical segments. In Proceedings of the 2016 Conference on Empirical Methods in Natural Language Processing, pages 961–967, Austin, Texas, November. Association for Computational Linguis- tics. Ronald M. Kaplan and Martin Kay. 1994. Regular mod- els of phonological rule systems. Computational Lin- guistics, 20(3):331–378. Martin Kay. 1977. Morphological and syntactic analysis. Linguistic Structures Processing, 5:131. Yoon Kim, Yacine Jernite, David Sontag, and Alexan- der M. Rush. 2016. Character-aware neural language models. In Proceedings of the Thirtieth AAAI Confer- ence on Artificial Intelligence, pages 2741–2749. Diederik Kingma and Jimmy Ba. 2015. Adam: A method for stochastic optimization. In International Conference on Learning Representations. Max Kisselew, Sebastian Padó, Alexis Palmer, and Jan Šnajder. 2015. Obtaining a better understanding of distributional models of German derivational morphol- ogy. In Proceedings of the 11th International Confer- ence on Computational Semantics, pages 58–63. Daphne Koller and Nir Friedman. 2009. Probabilistic Graphical Models: Principles and Techniques. MIT Press. Lingpeng Kong, Chris Dyer, and Noah A Smith. 2016. Segmental recurrent neural networks. In 4th Interna- tional Conference on Learning Representations. Mikko Kurimo, Sami Virpioja, Ville Turunen, and Krista Lagus. 2010. Morpho Challenge competition 2005– 2010: Evaluations and results. In Special Interest Group on Computational Morphology and Phonology. Angeliki Lazaridou, Marco Marelli, Roberto Zamparelli, and Marco Baroni. 2013. Compositional-ly derived representations of morphologically complex words in distributional semantics. In Proceedings of the 51st Annual Meeting of the Association for Computational Linguistics (Volume 1: Long Papers), pages 1517– 1526, Sofia, Bulgaria, August. Association for Com- putational Linguistics. Omer Levy and Yoav Goldberg. 2014a. Dependency- based word embeddings. In Proceedings of the 52nd Annual Meeting of the Association for Computational Linguistics (Volume 2: Short Papers), pages 302–308, 46 Baltimore, Maryland, June. Association for Computa- tional Linguistics. Omer Levy and Yoav Goldberg. 2014b. Neural word embedding as implicit matrix factorization. In Ad- vances in Neural Information Processing Systems, pages 2177–2185. Marc Light. 1996. Morphological cues for lexical se- mantics. In Proceedings of the 34th Annual Meeting of the Association for Computational Linguistics, pages 25–31, Santa Cruz, California, USA, June. Associa- tion for Computational Linguistics. Wang Ling, Chris Dyer, Alan W. Black, Isabel Trancoso, Ramon Fermandez, Silvio Amir, Luis Marujo, and Tiago Luis. 2015. Finding function in form: Com- positional character models for open vocabulary word representation. In Proceedings of the 2015 Conference on Empirical Methods in Natural Language Process- ing, pages 1520–1530, Lisbon, Portugal, September. Association for Computational Linguistics. Thang Luong, Richard Socher, and Christopher Man- ning. 2013. Better word representations with recur- sive neural networks for morphology. In Proceedings of the Seventeenth Conference on Computational Nat- ural Language Learning, pages 104–113, Sofia, Bul- garia, August. Association for Computational Linguis- tics. John Lyons. 1977. Semantics. Cambridge University Press. Tomas Mikolov, Kai Chen, Greg Corrado, and Jeffrey Dean. 2013a. Efficient estimation of word representa- tions in vector space. In 2th International Conference on Learning Representations. Tomas Mikolov, Ilya Sutskever, Kai Chen, Gregory S. Corrado, and Jeffrey Dean. 2013b. Distributed rep- resentations of words and phrases and their composi- tionality. In Advances in neural information process- ing systems, pages 3111–3119. Jeff Mitchell and Mirella Lapata. 2008. Vector-based models of semantic composition. In Proceedings of Association of Computational Linguistics, pages 236– 244, Columbus, Ohio, June. Association for Computa- tional Linguistics. Kevin P. Murphy. 2012. Machine Learning: A Proba- bilistic Perspective. MIT Press. Jason Naradowsky and Sharon Goldwater. 2009. Im- proving morphology induction by learning spelling rules. In Twenty-first International Joint Conference on Artificial Intelligence, pages 1531–1536. Karthik Narasimhan, Damianos Karakos, Richard Schwartz, Stavros Tsakalidis, and Regina Barzilay. 2014. Morphological segmentation for keyword spot- ting. In Proceedings of the 2014 Conference on Empirical Methods in Natural Language Processing (EMNLP), pages 880–885, Doha, Qatar, October. As- sociation for Computational Linguistics. Karthik Narasimhan, Regina Barzilay, and Tommi Jaakkola. 2015. An unsupervised method for uncov- ering morphological chains. Transactions of the Asso- ciation for Computational Linguistics, 3:157–167. OUP editors. 2010. New Oxford American Dictionary. Oxford University Press. Sebastian Padó, Alexis Palmer, Max Kisselew, and Jan Šnajder. 2015. Measuring semantic content to as- sess asymmetry in derivation. In Proceedings of the 11th International Conference on Computational Se- mantics. Lawrence R. Rabiner. 1989. A tutorial on hidden Markov models and selected applications in speech recognition. Proceedings of the Institute of Electrical and Electronics Engineers, 77(2):257–286. Pushpendre Rastogi, Ryan Cotterell, and Jason Eisner. 2016. Weighting finite-state transductions with neu- ral context. In Proceedings of the 2016 Conference of the North American Chapter of the Association for Computational Linguistics: Human Language Tech- nologies, pages 623–633, San Diego, California, June. Association for Computational Linguistics. Eric S. Ristad and Peter N. Yianilos. 1998. Learning string-edit distance. IEEE Transactions on Pattern Analysis and Machine Intelligence, 20(5):522–532. Reuven Y. Rubinstein and Dirk P. Kroese. 2011. Sim- ulation and the Monte Carlo method. John Wiley & Sons. Teemu Ruokolainen, Oskar Kohonen, Sami Virpioja, and Mikko Kurimo. 2013. Supervised morphological seg- mentation in a low-resource learning setting using con- ditional random fields. In Proceedings of the Sev- enteenth Conference on Computational Natural Lan- guage Learning, pages 29–37, Sofia, Bulgaria, Au- gust. Association for Computational Linguistics. Gerard Salton, editor. 1971. The SMART Retrieval System—Experiments in Automatic Document Pro- cessing. Prentice Hall. Sunita Sarawagi and William W. Cohen. 2005. Semi- Markov conditional random fields for information ex- traction. In Advances in neural information process- ing systems, pages 1185–1192. Patrick Schone and Daniel Jurafsky. 2000. Knowledge- free induction of morphology using latent semantic analysis. In Proceedings the 4th Conference on Com- putational Natural Language Learning, pages 67–72. Association for Computational Linguistics. Patrick Schone and Daniel Jurafsky. 2001. Knowledge- free induction of inflectional morphologies. In Pro- ceedings of the Second Meeting of the North American 47 Chapter of the Association for Computational Linguis- tics on Language Technologies, pages 1–9. Associa- tion for Computational Linguistics. Wolfgang Seeker and Özlem Çetinoğlu. 2015. A graph- based lattice dependency parser for joint morpholog- ical segmentation and syntactic analysis. Transac- tions of the Association for Computational Linguistics, 3:359–373. Radu Soricut and Franz Och. 2015. Unsupervised mor- phology induction using word embeddings. In Pro- ceedings of the 2015 Conference of the North Ameri- can Chapter of the Association for Computational Lin- guistics: Human Language Technologies, pages 1627– 1637, Denver, Colorado, May–June. Association for Computational Linguistics. Ilya Sutskever, Oriol Vinyals, and Quoc V. Le. 2014. Se- quence to sequence learning with neural networks. In Advances in Neural Information Processing Systems, pages 3104–3112. Charles Sutton and Andrew McCallum. 2006. An in- troduction to conditional random fields for relational learning. In Lise Getoor and Ben Taskar, editors, Introduction to Statistical Relational Learning, pages 93–128. MIT Press. Peter D. Turney and Patrick Pantel. 2010. From fre- quency to meaning: Vector space models of semantics. Journal of Artificial Intelligence Research, 37(1):141– 188. Antal van den Bosch and Walter Daelemans. 1999. Memory-based morphological analysis. In Proceed- ings of the 37th Annual Meeting of the Association for Computational Linguistics, pages 285–292, College Park, Maryland, USA, June. Association for Compu- tational Linguistics. Martin J. Wainwright and Michael I. Jordan. 2008. Graphical models, exponential families, and varia- tional inference. Foundations and Trends in Machine Learning, 1(1-2):1–305. Zhen Wang, Jianwen Zhang, Jianlin Feng, and Zheng Chen. 2014. Knowledge graph and text jointly em- bedding. In Proceedings of the 2014 Conference on Empirical Methods in Natural Language Processing (EMNLP), pages 1591–1601, Doha, Qatar, October. Association for Computational Linguistics. Paul J. Werbos. 1990. Backpropagation through time: What it does and how to do it. Proceedings of the Institute of Electrical and Electronics Engineers, 78(10):1550–1560. Richard Wicentowski. 2002. Modeling and Learning Multilingual Inflectional Morphology in a Minimally Supervised Framework. Ph.D. thesis, Johns Hopkins University. Yadollah Yaghoobzadeh and Hinrich Schütze. 2015. Corpus-level fine-grained entity typing using contex- tual information. In Proceedings of the 2015 Confer- ence on Empirical Methods in Natural Language Pro- cessing, pages 715–725, Lisbon, Portugal, September. Association for Computational Linguistics. David Yarowsky and Richard Wicentowski. 2000. Min- imally supervised morphological analysis by multi- modal alignment. In The 38th Annual Meeting of the Association for Computational Linguistics. Wenpeng Yin and Hinrich Schütze. 2015. Convolutional neural network for paraphrase identification. In Pro- ceedings of the 2015 Conference of the North Ameri- can Chapter of the Association for Computational Lin- guistics: Human Language Technologies, pages 901– 911, Denver, Colorado, May–June. Association for Computational Linguistics. Britta D. Zeller, Jan Šnajder, and Sebastian Padó. 2013. DErivBase: Inducing and evaluating a derivational morphology resource for German. In Proceedings of the 51st Annual Meeting of the Association for Com- putational Linguistics (Volume 1: Long Papers), pages 1201–1211, Sofia, Bulgaria, August. Association for Computational Linguistics. Britta D. Zeller, Sebastian Padó, and Jan Šnajder. 2014. Towards semantic validation of a derivational lexicon. In Proceedings the 25th International Conference on Computational Linguistics, pages 1728–1739, Dublin, Ireland, August. Dublin City University and Associa- tion for Computational Linguistics. 48