Poster abstracts

A longitudinal exploration of the effects of singing on L2 beginner pronunciation

Vita Kogan and Nadezda Bragina, University College London / Queen Mary University

It is not uncommon for adult L2 learners to have accented production even at the late stages of L2 acquisition when all other linguistic systems are mastered (Reiterer et al., 2011). Previous research has shown a connection between musical training and improved L2 pronunciation (Christiner & Reiterer, 2017; Nardo & Reiterer, 2009; Seither-Preisler et al, 2014). However, most of the studies conducted on the topic have focused on laboratory settings and cross-sectional design. Thus, little is known about the effect of regular singing in the instructional SLA, especially with lesser-taught non-Romance languages such as Russian. 

This longitudinal study follows beginner learners of Russian (n=11) for one year as they undergo their typical university-level language course. In contrast to the previous year's group of students (controls), this cohort had 30 minutes of in-class singing weekly as part of their academic syllabus. The production data were collected every week for 24 weeks (12 weeks of Semester 1 and 12 weeks of Semester 2) and constituted both reading and spontaneous speech. The control group did not undertake any training and participated in regular language classes following their normal academic routine. The pronunciation was evaluated: 1) holistically with 6\six na√Øve ratters submitting their impressions of the accentedness on a 5-level Likert scale, and 2) acoustically in Praat with vowels, consonants (voice onset time and palatalization), intonation contour/F0, and lexical stress examined for all 24 production samples for each participant (216 speech samples). The holistic measures did not differ significantly between the treatment and control groups but linear hierarchical regressional analysis demonstrated a significant difference in vowel production (p=.007) and lexical stress (p>.0001). As vowels are typically more sustained or extended during singing in comparison to speaking, learners who sing regularly might have more time to process language-specific acoustic cues. Previous research has also shown that music training is positively related to the acquisition of prosody (e.g., Torppa et al., 2020). These findings suggest that signing should be integrated into the L2 curriculum as a relatively simple and enjoyable way to improve some aspects of L2 production. 

References

Christiner, M., Rüdegger, S., & Reiterer, S. M. (2018). Sing Chinese and tap Tagalog? Predicting individual differences in musical and phonetic aptitude using language families differing by sound-typology. International Journal of Multilingualism, 15(4), 455-471.

Nardo, D., & Reiterer, S. (2009). Musicality and phonetic language aptitude. Language talent and brain activity, 213-256. Reiterer, S. M., Hu, X., Erb, M., Rota, G., Nardo, D., Grodd, W., ... & Ackermann, H. (2011). Individual differences in audio-vocal speech imitation aptitude in late bilinguals: functional neuroimaging and brain morphology. Frontiers in psychology, 2, 271.

Seither-Preisler, A., Parncutt, R., & Schneider, P. (2014). Size and synchronization of auditory cortex promotes musical, literacy, and attentional skills in children. Journal of Neuroscience, 34(33), 10937-10949.

Torppa, R., Faulkner, A., Laasonen, M., Lipsanen, J., & Sammler, D. (2020). Links of prosodic stress perception and musical activities to language skills of children with cochlear implants and normal hearing. Ear and Hearing, 41(2), 395-410. 

How effective is intentional vocabulary learning using word cards?

Sally Alghamdi (Albaha University) and Jeanine Treffers-Daller (University of Reading).  

Research shows that intentional vocabulary learning is more effective than incidental vocabulary learning (Elgort, 2011; M. STRATTON, 2022; Nation, 2013). Moreover, intentional learning using word cards is one of the most effective techniques in vocabulary learning (Nakata, 2019) as it can help commit large number of vocabulary items to long term memory in a short period of time (Dunlosky, Rawson, & Marsh, 2013; Fitzpatrick, Al-Qarni, & Meara, 2008). On the other hand, incidental exposure to vocabulary items in graded readers is viewed as the main means of vocabulary retention in some research (Krashen, 2004; Pitts, White, & Krashen, 1989).

In the current study, seventy-five L1 low intermediate Arabic learners of English were assigned to three groups where they either learned a set of 45 intentionally using word cards or learned them incidentally using graded readers and a control group (CG) that did not receive any vocabulary guidance. The proportion of target words retained by Word Card Group (WCG) was compared to the proportion of target words retained by the Graded Readers Group (GRG) and both were compared with the CG performance. Four test sessions, three immediate post-test and one delayed post-test, were administered to measure vocabulary receptive knowledge with a Lexical Decision Task (LDT), and productive vocabulary knowledge was measured with a Gap-Fill Test (GFT). Analyses with a mixed factorial ANOVA revealed that the WCG outperformed other groups in all sessions in receptive and productive knowledge of target words. We also investigated the effect of word characteristics (morphological complexity and word length) and frequency of exposure on retention for Arab ESL learners. Multiple regression analyses showed that the frequency of exposure to the target words predicted the accuracy scores on the LDT and the scores on the GFT for the WCG, and the number of syllables in a target word predicted the GFT scores for all the groups combined. We conclude by formulating implications for intentional and incidental vocabulary learning. 

References: 

Dunlosky, J., Rawson, K. A., & Marsh, E. J. (2013). Nathan MJ willingham DT improving students’ learning with effective learning techniques: Promising directions from cognitive and educational psychology. Association for Psychological Science, 14(1), 4-58. 

Elgort, I. (2011). Deliberate learning and vocabulary acquisition in a second language. Language Learning, 61(2), 367-413. 

Fitzpatrick, T., Al-Qarni, I., & Meara, P. (2008). Intensive vocabulary learning: A case study. Language Learning Journal, 36(2), 239-248. 

Krashen, S. D. (2004). The power of reading: Insights from the research: Insights from the research ABC-CLIO

M. STRATTON, J. (2022). Intentional and incidental vocabulary learning: The role of historical linguistics in the second language classroom. The Modern Language Journal

Nakata, T. (2019). Learning words with flash cards and word cards. The Routledge Handbook of Vocabulary Studies, 304-319. 

Nation, I. S. P. (2013). Learning vocabulary in another language (2nd ed.). Cambridge: Cambridge University Press. 

Pitts, M., White, H., & Krashen, S. (1989). Acquiring the second language vocabulary through reading: A replication of the clockwork orange study using second language acquirers.  

Language awareness in primary school EFL lessons: teachers' cognitions and practices

Monika Bader (Western Norway University of Applied Sciences), Christine Möller-Omrani (Western Norway University of Applied Sciences) and Simon Borg (Western Norway University of Applied Sciences).  

Language awareness (LA), often defined as ' explicit knowledge about language, and conscious perception and sensitivity in language learning, language teaching and language use’ (ALA, n.d.), plays an important role in foreign language learning and teaching (e.g. Svalberg, 2007; Frijns et. al. 2018), as well as in national curricula, including in Norway, which is the context for the research we present here. How effectively LA is promoted in the classroom is influenced by teachers’ cognitions (Borg, 2019) and pedagogical expertise and this study (which was part of a larger project examining the development of metalinguistic awareness in primary school pupils), examined three questions: 1. To what extent are English lessons for Grade 3 (mean age 8) characterized by pedagogical practices that promote the development of pupils’ LA? 2. Where LA is promoted, (a) what pedagogical practices do teachers adopt and (b) which areas of language are targeted? 3. What cognitions – beliefs, knowledge, understandings and perceptions – related to LA but also to L2 teaching and learning more generally, to young learners and to the educational context (including the curriculum) shape teachers’ pedagogical practices in relation to LA? 

The results presented come from case studies of four teachers of English whose practices and cognitions were explored qualitatively through one background interview (to learn about the teachers’ context and their general views about language awareness), two lesson observations (which were studied for evidence of practices that promote LA) and one post-observation interview (where classroom episodes relevant to LA were discussed with the teacher). 

Overall, English lessons were characterized by modest levels of attention to LA. Lessons were typified by series of language activities which pupils completed, typically with limited explicit attention to language. In discussing their lessons, teachers demonstrated uncertain understandings of what LA meant and of how to interpret references to LA in the curriculum. While teachers were generally positive about the value of LA in the learning of English, various factors combined to limit the extent to which LA was promoted in their lessons. These included their beliefs about young learners and the process of L2 learning, as well as teachers’ lack of theoretical and pedagogical knowledge relevant to LA. 

On the basis of these conclusions, suggestions for the kinds of professional development support that would enable teachers to promote LA more effectively in primary English classrooms will be made. 

References 

Association for Language Awareness (ALA). (n.d.). 

Borg, S. (2019). Language teacher cognition: Perspectives and debates. In X. Gao (Ed.), Second handbook of English language teaching (pp. 1149-1170). Springer International Publishing. 

Frijns, C., Sierens, S., Van Avermaet, P., Sercu, L., & Van Gorp, K. (2018). Serving policy or people? Towards an evidence-based, coherent concept of language awareness for all learners. In C. Hélot, C.Frijns, K.V.Gorp, & S. Sierens (eds.), Language Awareness in Multilingual Classrooms in Europe (pp. 87-116). De Gruyter Mouton. 

Svalberg, A. (2007). Language awareness and language learning. Language Teaching, 40(4), 287-308. doi:10.1017/S0261444807004491  

Acqusition of generic meanings in L3 Italian

Eleonora Boglioni (University of Southampton) and Roumyana Slabakova (University of Southampton).  

Genericity contrasts.

English speakers with Spanish as a second language learning Italian can potentially rely on the L1 and L2 grammars to parse the L3 input. Spanish and Italian share lexical and structural similarities, e.g., in the expression of generic meanings. Generic descriptions can refer to all the members of a species (kind reading: Tigers are becoming extinct) or instances of a species engaging in habitual behaviour (characterizing reading: Lions roar) (Krifka et al., 1995). Importantly, the syntactic realization of both generic meanings varies crosslinguistically. While English bare plural subjects can have kind/characterizing readings (Chierchia, 1998), Spanish and Italian generic subjects require definite articles (Longobardi, 1994). Hence, Spanish and Italian pattern similarly in this respect. Nonetheless, Spanish and Italian pattern differently in allowing bare singular objects with number neutral interpretation in a restricted verb class (e.g., Tengo coche, *I have car), where the object refers to one or more car(s) (Espinal, 2010). To express a similar but not exactly the same meaning, English and Italian require overt determiners. Thus, for each property, two of the languages investigated pattern together, to the exclusion of the third one. L3 acquisition predictions.

Our research investigates crosslinguistic influence (CLI) effects from the background languages to the L3, testing the effects of structural similarity and L3 proficiency. On objects, English–Italian parallels could be beneficial, while negative CLI from Spanish is expected. On subjects, positive influence from Spanish is predicted. 

Methodology.

We tested 60 adult L3 Italian learners with English and Spanish alternatively as L1/L2. To address the role of input, exposure and use, we recruited 30 L1 English speakers in England and 30 L1 Spanish speakers in Spain. These settings could favour English CLI on objects and Spanish CLI on subjects. Comprehension and use of generics were assessed in the L2 and L3 by means of an Acceptability Judgment Task, a Form-to-Meaning Task, and an Elicited Oral Production Task.  Results. We ran linear mixed models (R packages lme4, glmer) on the L3 Italian judgements and oral data (Tables 2–4). On the AJT, significant differences (p < .001) were found on the contrasts between acceptable and unacceptable kind-referring subjects, and number neutral objects (Table 2), in interaction with L3 proficiency. This suggests property-by-property CLI with facilitative effects from Spanish on subjects and from English on objects. In oral production, we observed a main effect for L3 proficiency (Table 4). Individual results (Table 5) point to negative CLI from English (on subjects) and Spanish (on objects). 

Theoretical proposal.

Additional language acquisition involves reorganising the known form–meaning mappings. Taken together, our experimental results point to  Hegemony of the Form. In the case of universal semantic distinctions where all languages express the same meanings, although with different morphosyntactic means (e.g., kind/characterizing genericity), the forms exert strong crosslinguistic influence. Where a different meaning exists in one of the background languages (the Spanish number neutral meaning), it does not exert such a strong crosslinguistic influence. We will discuss implications of this proposal in other additional language acquisition scenarios. 

Internal validity of the new LLAMA (v.3) aptitude tests

Lars Bokander (Jönköping University), Vivienne Rogers (Swansea University), Paul Meara (Swansea University) and Brian Rogers (Swansea University).  

The LLAMA aptitude tests have undergone significant revision since their first iteration (Meara, 2005). The revised version 3 has been available on the Lognostics website since 2021. In this paper we will update the internal validity analysis from Bokander & Bylund (2020) with data from 640 test-takers of the new version. 

We analysed the LLAMA v3 scores with respect to item performance and internal consistency. Item performance was evaluated with classical test theory and Rasch model fit. Internal consistency is influenced by dimensionality, i.e., whether the test measures one or more constructs. Bokander & Bylund (2020) pointed out issues with dimensionality in some LLAMA v.1 subtests. We address this by reporting coefficient omega along with the traditional Chronbach’s alpha. Omega is based on a factor analysis of the items in a test and has less strict statistical assumptions than coefficient alpha, for example, with regards to unidimensionality (Dunn et al., 2014). 

The results indicate several improvements in the new LLAMA tests. Overall, items on average performed better than in LLAMA v1 with respect to Rasch model fit and discrimination between participants. Improved item performance in turn reflects improvements in internal consistency. LLAMA B has not notable changed, and again produced reliable scores. LLAMA D has undergone important changes in how the items are presented to test-takers. This has increased the internal consistency of the scores significantly compared to version 1. Coefficient omega for LLAMA D indicates an internal consistency comparable to that typically found with instruments in SLA research (Plonsky & Derrick, 2016). LLAMA E scores produced high values for both alpha and omega. This improvement is due to a greater score variance and the more normal distribution of scores obtained with the new version. LLAMA F has undergone major changes in item format and scoring procedure. We found a significantly higher internal consistency in our data set than what has previously been reported, with both alpha and omega performing well. Dimensionality related issues seem less pronounced with the new version, although some questions remain about its unidimensionality. 

An implication for language aptitude researchers is that the new LLAMA tests should be able to produce better correlations with other SLA measures, because correlations depend on reliably obtained scores from all instruments in a study. If there actually exists a relationship between two variables, it may still go undetected when unreliable instruments are employed. The risk of this occurring in future aptitude research has now decreased. 

References 

Bokander, L., & Bylund, E. (2020). Probing the internal validity of the LLAMA language aptitude tests. Language Learning, 70(1), 11-47. 

Dunn, T. J., Baguley, T., & Brunsden, V. (2014). From alpha to omega: A practical solution to the pervasive problem of internal consistency estimation. British Journal of Psychology, 105(3), 399-412. 

Meara, P. (2005). LLAMA language aptitude tests: The manual. Lognostics. 

Plonsky, L., & Derrick, D. J. (2016). A meta‐analysis of reliability coefficients in second language research. The Modern Language Journal, 100(2), 538-553.  

Interactive didactics: enhancing L2 speaking complexity and accuracy through a blended drama approach

Simona Bora (University of Essex (UK) / Free University of Bozen (IT)).  

A steadily increasing number of educators in different contexts have been focusing their work on ways in which dramatic approaches can support foreign language learning. Drama is “communication between people” (Via, 1987: 10) and “inextricable part of all social interactions” (Russel di Napoli 2003:17) and therefore offers potentially a very rich interactive context for enhancing students’ L2 oral skills. Although drama in language teaching has gained recognition recently, there have been no quantitative studies conducted to date which investigate the gains made by students learning languages through a blended drama approach in a mandatory high-school curriculum (Bora 2022) in terms of complexity and accuracy. 

This paper sheds light on findings from a longitudinal (20 weeks) mixed-method approach study conducted with high-school intermediate level of language proficiency Italian students. Participants were exposed to a blended drama approach: theatre texts combined with drama activities followed by a stage performance. A control group was taught through a traditional approach instead. Quantitative data has been collected through the implementation of an oral pre-test, a mid-test and a post-test with three tasks: oral proficiency interview, story retelling and guided role-play. The impact of this approach is measured in terms of students’ gains in oral accuracy and complexity when tasks are taken together or separately. Results lend support to previous hypotheses of the effectiveness of a blended drama approach to enhancing L2 speaking skills compared to a traditional one. Implications for teaching practice are further discussed. 

References

Bora, S. F (2022). Drama Pedagogy: Speaking Accuracy and Complexity through Contemporary Theatrical Texts and Performance in Foreign Language Learning. Research in Drama Education: The Journal of Applied Theatre and Performance

Housen, Alex, Folkert Kuiken, Ineke Vedder (2012), Dimensions of L2 Performance and Proficiency. Complexity, Accuracy and Fluency in SLA, Amsterdam: John Benjamins.

Palloti, Gabriele (2015), “CAF: defining, refining and differentiating constructs”, Applied Linguistics, 30(40), 590-601.  

Cognitive engagement and glossing: effects of L1, L2 and intercomprehensible glosses on vocabulary learning. An eye-tracking study.

Ilaria Borro (Università degli Studi di Bergamo).  

Studies of the effectiveness of glossing in L2 reading hypothesized the amount of cognitive engagement with the glossed words to be a relevant factor for vocabulary learning (Boers 2022). Different strategies have been implemented to manipulate cognitive engagement, such as L2 rather than L1 glosses (Kim et al 2020), multiple-choice glosses (e.g. Yoshii 2013), and glosses providing only one of the meanings of polysemic items (Boers 2000).  Findings are currently mixed, albeit encouraging. Notably, none of the existing studies employed real-time measurements of cognitive engagement, which has been assumed rather than assessed.

The present project aims to contribute filling this gap.  One-hundred L2 Italian learners will be assigned to three experimental groups, each exposed to a version of a reading text with glosses requiring a growing amount of cognitive engagement: (i) L1 glosses; (ii) L2 glosses; (iii) glosses in a romance language, unknown but intercomprehensible to the learners due to a degree of semantic transparency with Italian (Blanche-Benveniste et al 1979). A control group will read the text with no glosses.  The reading behavior and cognitive engagement will be assessed through eye-tracking, while the creation of explicit and implicit knowledge of vocabulary will be measured, respectively, with a set of pencil-and-paper tests and a priming protocol test. Immediate and delayed post-test results will be triangulated with eye-tracking data and retrospective verbal report outcomes, to value the participants’ awareness at the point of learning. Expected outcomes include a larger cognitive engagement for L2 and intercomprehensible glosses as compared to L1 glosses, as well as a significant effect of cognitive engagement on vocabulary learning and retain.  The project contributes with empirical evidence to the debate about the role of attention and cognitive engagement for the creation of implicit and explicit knowledge (Long 2017). Moreover, it presents immediate pedagogical implications for a more effective use of glosses in instructed SLA. Finally, it provides data regarding the possible use of intercomprehension as a pedagogic tool for language learning. 

References

Blanche-Benveniste, C., Borel, B., Deulofeu, J., Durand, J., Giacomi, A., Loufrani, C., Meziane, B., & Pazery, N. (1979). Des grilles pour le français parlé, « Recherches sur le français parlé », 2, 163-206.

Boers, F. (2000). Enhancing metaphoric awareness in specialised reading. English for Specific Purposes, 19(2), 137–147.

Boers, F. (2022). Glossing and vocabulary learning. Language Teaching (2022), 55, 1–23 

Kim, H. S., Lee, J. H., & Lee, H. (2020). The relative effects of L1 and L2 glosses on L2 learning: A meta-analysis. Language Teaching Research, Online First.

Long, M. (2017). Geopolitics, methodological issues, and some major research questions. ISLA, 1(1), 7- 44.

Yoshii, M. (2013). Effects of gloss types on vocabulary learning through reading: Comparison of single and multiple gloss types. Special issue on learner-computer interaction in language education: A Festschrift in honor of Robert Fisher, CALICO Journal, 203–229.  

Pragmatic awareness and proficiency: Are highly proficient learners more pragmatically aware?

Athenea Botey (Universitat de Barcelona) and Júlia Barón (Universitat de Barcelona).  

In the field of interlanguage pragmatics (ILP), the relationship between foreign/second language (L2) proficiency and pragmatic competence has long been debated. Research on the interface between linguistic and pragmatic competence in an L2 has yielded results that point to the correlation of learners’ proficiency level and pragmatic development (i.e., Bardovi-Harlig and Dörnyei, 1998; Cook and Liddicoat, 2002; Schauer, 2006) as well as the correlation of grammatical and pragmatic development (i.e., Hoffman-Hicks, 1992; Håkansson and Norrby, 2005; Celaya and Barón, 2015). Nonetheless, a number of studies have found that pragmatic competence and proficiency do not correlate and are in fact independent (i.e., Niezgoda and Röver, 2001; Matsumura, 2003). Most of the research regarding this interface has tested their participants’ productive knowledge of pragmatics. In addition, speech acts such as requests, apologies, suggestions and refusals have been amongst the most frequently investigated in the aforementioned studies. That is why the present study focuses on compliment responses (CRs) and on perceptive knowledge of pragmatics. Therefore, this research aims to investigate the interface between proficiency and ILP by focusing on the pragmatic awareness of the speech act of CRs. 

The participants were 72 in total, of which 8 were native speakers (NSs) of English (mean age= 19.5) and 64 were English as a foreign language (EFL) learners (mean age= 16.6) with Spanish/Catalan as their first language (L1). In order to assess their perceptive pragmatic knowledge of CRs, a pragmatic awareness video elicitation task (PAVET) was administered. When carrying out the PAVET test, participants had to rate the appropriateness of 15 CRs using a Likert-scale that went from 1 (inappropriate) to 6 (very appropriate). The audio-visual compliment-compliment response sequences were extracted from the TV series Gilmore Girls, chosen for its display of various degrees of social distance and power. Besides completing the PAVET, four participants agreed to carry out semi-structured post hoc interviews. The EFL learners were divided into groups of high and low proficiency according to their vocabulary sizes (V_YesNo v1.1 [lognostics.co.uk]) to analyse if and in what ways their responses to the task differed. Their ratings were then compared to a NS benchmark to obtain a sameness score. 

Results indicated a moderate inverse correlation between proficiency level and near-nativeness. Findings, on the one hand, suggest that the L1 of the learners plays an important role in their L2 pragmatic awareness. On the other hand, they point to the direction that proficiency alone is not a determinative factor in order to acquire pragmatic knowledge as there are several other factors that can influence pragmatic development, such as individual differences, the non-linearity of pragmatic development, the use of compensatory nonverbal strategies or exposure to limited pragmatic input in the classroom context. All in all, given the results, it could be implied that pragmatic development and L2 language proficiency are of an independent nature.  

From paper-based to computer-based integrated reading-to-write: Evidence for delivery mode effects at the CEFR B1 level

Tineke Brunfaut (Lancaster University), Luke Harding (Lancaster University) and Aaron O. Batty (Keio University).  

Computer-based tasks are increasingly used in the teaching and assessment of second language (L2) writing, reflecting large-scale changes in day-to-day writing modes, technological developments, increased computer accessibility, administrative efficiency, and—most recently—online delivery due to the Covid19 pandemic. While some writing tasks are conceptualised as computer-based from inception, others have shifted mode from their original paper-based format (either to replace the latter or be used in parallel mode). Research has shown, however, that mode of delivery may have an effect on the L2 writing process and product, and on writers’ perceptions (see Chan, 2018). In the last two decades, the teaching and assessment of writing has furthermore broadened from independent, writing-only tasks to include integrated tasks such as reading-to-write or listening-to-write, reflecting forms of authentic language use. While a sizable body of research on integrated-writing tasks now exists (Plakans, 2015), little is known yet about how these types of tasks operate comparatively in different delivery modes. However, a study by Authors (year) looking into the effect of delivery mode on a suite of independent and integrated writing tasks across CEFR B1-C1 proficiency levels found that while delivery mode had no discernible effect on scores at the B2 and C1 level, learners completing the B1-level tasks scored slightly, but statistically significantly, lower on the computer-based mode than they did on the paper-based mode. 

This presentation reports on a follow-up study, aiming to gain more detailed insights into the observed delivery mode effect of the integrated reading-to-write task at the B1 level. To this end, the discourse characteristics of students’ written samples on the B1 tasks (both independent and integrated ones) were investigated. 106 English-L2 learners based in three European countries each completed one version of the B1 writing-only task and one of the B1 reading-to-write task in the paper-based mode, and another version of each task in the computer-based mode. The tasks were counterbalanced for order of delivery mode and for task version. To establish the discourse characteristics of the written performances, students’ writing samples were analysed using 28 automated discourse measures (using various software tools) and 8 rating scales developed for fine-grained judgements by two experts. The analyses covered the areas of reading-text source use, task fulfilment, textual organisation and structure, and language control. Comparative statistics (Wilcoxon Signed Rank Tests) showed that effects on discourse features were primarily associated with the integrated reading-to-write tasks, rather than the writing-only tasks, thus triangulating the previous score-based findings Authors (year). Specifically, learners gained small advantages in the paper-based mode regarding the selection of relevant content from source materials, and on measures of fluency and accuracy. Implications will be discussed concerning the potential impact of user-interface features and spelling/proofreading challenges for lower proficiency learners (B1) in the screen-based completion of integrated tasks. 

References

Authors (Year). Chan, S. (Ed.) (2018). The comparability of paper-based and computer-based writing: Process and performance [Special Issue]. Assessing Writing, 36.

Plakans, L. (2015). Integrated second language writing assessment: Why? What? How? Linguistics and Language Compass, 9(4), 159-167.  

Exploring effects of early extramural English exposure on university students’ current L2 vocabulary

Nicole Busby (Norwegian University of Science and Technology (NTNU9).  

This exploratory study investigated university students’ reported exposure to English through informal out-of-class (extramural) activities at an early age and how this related to their current vocabulary size in L2 English. Studies have shown that extramural activities can be an important predictor of L2 vocabulary knowledge and that the amount of time spent on these activities is very important (e.g. Sundqvist, 2009; Peters, 2018). Research into effects of extramural exposure on language acquisition has tended to focus on looking for relationships between L2 proficiency and extramural activities that participants engage in at the time of the study, but less is known about the relationship between language proficiency and activities earlier in life which could have contributed to L2 acquisition. 

To investigate this, students at a Norwegian university (N = 34) were asked to complete a survey comprising questions about their early extramural English activities and questions from the Vocabulary Size Test (VST: Nation & Beglar, 2007). Participants were asked to report the earliest extramural activity they felt made an important contribution to their current knowledge of English and the extramural activity they considered most important (if these were different), as well as the age at which they engaged in these activities.  Participants’ mean English vocabulary size, as measured by the VST, was 11,832 words. All participants were able to report specific activities that they felt were important to their acquisition of English. The most frequently reported categories for participants’ earliest English activities were watching TV shows (24%) and reading books (26%). Activities that participants reported as the most important were more varied, with the most common being TV shows (21%), social media (15%) and online videos (15%).  Regression analysis found that the age of reported earliest extramural exposure was a significant predictor for L2 vocabulary size, but that the current age of participants was not a significant predictor of vocabulary size. This is in line with previous research showing that age of onset is an important predictor of L2 acquisition (e.g. Hyltenstam, 1992), but it was interesting to see that this appears to be more important in this sample than the overall time available for language input (as indicated by participant age). Although this approach of asking participants for self-reports of early memories clearly has limitations in terms of accuracy of reporting, results from this exploratory study suggest that investigating early exposure to informal language learning activities could be an important avenue for future research. 

References

Hyltenstam, K. (1992). Non-native features of near-native speakers: On the ultimate attainment of childhood L2 learners. Advances in Psychology, 83, 351-368.

Nation, I.S.P. & Beglar, D. (2007) A vocabulary size test. The Language Teacher, 31(7), 9-13.

Peters, E. (2018). The effect of out-of-class exposure to English language media on learners’ vocabulary knowledge. ITL-International Journal of Applied Linguistics, 169(1), 142-168.

Sundqvist, P. (2009). Extramural English matters: Out-of-school English and its impact on Swedish ninth graders' oral proficiency and vocabulary (Doctoral dissertation, Karlstad University).  

Pragmatic Self-Concepts of Multilingual Children and Adolescents in Germany

Maria Busch (Martin Luther University of Halle-Wittenberg), Luca Plachy (Martin Luther University of Halle-Wittenberg), Matthias Ballod (Martin Luther University of Halle-Wittenberg) and Stephan Sallat (Martin Luther University of Halle-Wittenberg).  

Growing diversity in global contexts leads to higher awareness of linguistic and cultural diversity and increasing research in multilingual language development and foreign & second language acquisition (FSLA). Although psychological factors of FSLA such as motivation and learners attitudes have been largely studied in mostly adult learners, the subjective perspectives of children and adolescents have not been given intense focus. 

This is where so-called self-concepts, which are subjective views on oneself about one's own characteristics, competencies and challenges, come into play (Mercer, 2011). Self-concepts are not only related to motivation and learning processes, they are also connected with communicative interaction. Pragmatic language is a highly relevant aspect for multilingual language development as well (Félix-Brasdefer, 2017; Kecskes, 2018). While verbal and social self-concepts have been studied widely, it is unclear how self-concepts concerning pragmatic language have been taken into account so far. 

The research project “MehrSelbst” (funded by the German Federal Ministry of Education and Research, project duration 10/2022-09/2025) addresses this research gap and investigates research questions on how multilingual children and adolescents in Germany express their self-concepts through self-descriptions and self-evaluations regarding their own pragmatic language. This study therefore explores the perspectives of children and adolescents on their own pragmatic competences in a qualitative research design and triangulates them to their structural language and the evaluation of pragmatics by familiar caregivers and educators. Therefore, this proposed poster aims to present the theoretical background (results of a scoping review) of multilingual pragmatic language development, the interface with self-concept research and methodological considerations. 

References

Félix-Brasdefer, J. César (2017): Interlanguage Pragmatics. In: Yan Huang (Hg.): The Oxford handbook of pragmatics. Oxford, United Kingdom: Oxford University Press (Oxford handbooks in linguistics), p. 416–434. 

Kecskes, I. (2018). Intercultural Pragmatics. In F. Liedtke & A. Tuchen (Hrsg.), Handbuch Pragmatik (p. 140–148). J.B. Metzler. 

Mercer, S. (2011). Language learner self-concept: Complexity, continuity and change. System, 39(3), 335–346.  

A Systematic Review of the Construct Validity of the Academic Pearson Test of English

Phat Cao (University of Nottingham Malaysia), Csaba Zoltan Szabo (University of Nottingham Malaysia) and Barry Lee Reynolds (University of Macau). 

Global-scale language proficiency tests, such as the Academic Pearson Test of English (PTEA), are often judged by the extent to which they accurately measure the underlying theoretical concept they claim to measure, termed construct validity (Chapelle, 2021; Hughes & Hughes, 2020). Construct validity, in turn, yields trust and confidence among various stakeholders. Despite a plethora of studies investigating different aspects of construct validity of the PTEA (e.g., De Jong & Zheng, 2016; Zheng & Mohammadi, 2013), the results have been mixed. One reason for this is the variation in conceptualising test constructs and their characteristics, which often lead to discrepancies in how the data are collected, analysed, and interpreted in the existing studies.

Using a systematic review methodology, this study explored and evaluated the features of PTEA construct validity studies in terms of research objectives, construct characteristics and methodologies employed. Results from the review of 35 studies reveal that a variety of aspects of construct validity have been examined using both qualitative and quantitative designs, with trait-related constructs (i.e., gender factor or cognitive demand) being the predominant research focus and quantitative approaches predominantly employed. The findings additionally highlight discrepancies in the conceptualisation of constructs, resulting in the adopted theoretical frameworks and methods for construct validation as well as a tendency to overlook enabling skills (i.e., grammar, oral fluency, pronunciation, spelling, vocabulary, and written discourse). To contribute to the scholarship on construct validity of the PTEA, recommendations are made for broadening the research objectives and employing a mixed-methods approach in lieu of only quantitative or qualitative analyses. Furthermore, the theoretical frameworks employed for test validation purposes were considered. Implications are also discussed as they are related to construct validity and validation research of the PTEA as well as language testing. 

Exploring the Relationship Between Teachers' and Students’ Emotional Intelligence and Emotional Vocabulary

Allen Chee (University of Nottingham Malaysia), Csaba Szabo (University of Nottingham Malaysia) and Sharimila Ambrose (University of Nottingham Malaysia).  

Research in the affective domain indicates that learners with higher Emotional Intelligence (EI) are more likely to succeed academically due to their ability to regulate emotions, build social relationships, understand, and express their own and others' emotions (Maccann et al., 2020). Meta-studies have shown that there is a significant relationship between EI, learners’ academic achievement and English language abilities (Akpur, 2020; Sánchez-Álvarez et al., 2020). However, learners' ability to recognise, process, and express their own and others' emotions may be impaired by the lack of requisite emotional vocabulary (EV) knowledge (Dewaele, 2015). EV was found to play a vital role in learners’ emotional development and that language was a key element in shaping the development of these concepts (Nook et al., 2020). However, learners must first acquire awareness and knowledge of EV in order to comprehend and verbalise their emotions (Hoemann et al., 2019). Despite the importance of EI and EV in language learning, the relationship between EI and EV is largely unexplored, especially in Malaysia.

This study aimed to investigate the relationship between the four factors of learners’ EI (well-being, self-control, emotionality, sociability) and EV. The first study was conducted with 46 foundation students at a private university in Malaysia. Participants completed an EI questionnaire (TEIQue-SF) and a label generation task as a measure of productive EV knowledge. The results showed a positive and significant correlation between EI and EV. The emotionality factor had the strongest correlation with EV, while language background had no effect. The second study aimed to confirm and expand these results by controlling for language background, dominance, and proficiency in a larger cohort of participants. These findings confirmed the significant relationship between EI and EV. This highlights the importance for curriculum developers and practitioners to support students’ development of EI and EV when developing teaching materials and best teaching practices. 

References

Akpur, U. (2020). A Systematic Review and Meta-Analysis on the Relationship Between Emotional Intelligence and Academic Achievement. Educational Sciences: Theory & Practice, 20(4), 51–64. 

Dewaele, J.M. (2015). On emotions in foreign language learning and use. The Language Teacher, 39(3), 13–15. 

Hoemann, K., Xu, F., & Barrett, L. F. (2019). Emotion words, emotion concepts, and emotional development in children: A constructionist hypothesis. Developmental Psychology, 55(9), 1830–1849.  

Maccann, C., Jiang, Y., Brown, L. E. R., Double, K. S., Bucich, M., & Minbashian, A. (2020). Emotional Intelligence Predicts Academic Performance: A Meta-Analysis. Psychological Bulletin, 146(2), 150–186.  

Nook, E. C., Stavish, C. M., Sasse, S. F., Lambert, H. K., Mair, P., McLaughlin, K. A., & Somerville, L. H. (2020). Charting the development of emotion comprehension and abstraction from childhood to adulthood using observer-rated and linguistic measures. Emotion (Washington, D.C.), 20(5), 773–792. 

Sánchez-Álvarez, N., Berrios Martos, M. P., & Extremera, N. (2020). A Meta-Analysis of the Relationship Between Emotional Intelligence and Academic Performance in Secondary Education: A Multi-Stream Comparison. Frontiers in Psychology, 11, 1517.  

Exploring Linguistic Relativity: The Effect of the French Grammatical Gender System on Bilingual Adults’ Perception of Objects

Zhuohan Chen (Department of Education, university of Oxford).  

A growing body of literature on linguistic relativity (the hypothesis that language influences thoughts) has been focusing on whether there is an effect of the grammatical gender system on perception (Bassetti & Nicoladis, 2016; Lambelet, 2016; Samuel et al., 2019). Evidence suggests that there is an effect of the gender system on perception of both monolinguals and bilinguals (Bassetti & Nicoladis, 2016). However, the findings are potentially limited due to certain methodological limitations (Samuel et al., 2019), for example, the lack of L2 proficiency testing (Bassetti, 2007). Thus, attempting to apply a more rigorous methodology, this study aims to assess the potential effect of the French grammatical gender system on French speakers’ and learners’ perceptions of object gender.

The study involved 140 participants, divided into four groups (N = 35 per group): English monolinguals, French monolinguals, English-French (English-dominant) bilinguals, and French-English (French-dominant) bilinguals. An online experiment was distributed to participants, including a background information questionnaire, English and French vocabulary tests, and a voice distribution task. Quantitative data were analysed using multi-level modelling, relevant regression analyses, and t-tests. Follow-up open-ended question data were coded and analysed using chi-squared tests. The results supported linguistic relativity: The French grammatical gender system affected the perception of French monolinguals and English-dominant bilinguals. The effect of French on French-dominant bilinguals was not reduced by the acquisition of English and seemed to be independent of L2 proficiency. Additional findings included the potential tendency of the French to introduce gender biases (stereotypical association between objects and gender, resulting from the grammatical gender system).

The present study adds supporting evidence to the relativity debate by attempting to apply rigour by utilising a pilot study, a robust sample size, pertinent control items, L2 proficiency testing, and advanced data analysis tools. Pedagogically, findings highlight the need to emphasise the discrepancies between students’ potential preconceived perceptions and the rules of the L2 grammar when teaching an L2 with a different grammatical gender system to L1. The study also shed new light on whether gendered languages draw out potential gender stereotypes among bilinguals.   

References

Bassetti, B. (2007). Bilingualism and thought: Grammatical gender and concepts of objects in Italian-German bilingual children. The International Journal of Bilingualism: Cross-Disciplinary, Cross-Linguistic Studies of Language Behaviour, 11(3), 251-273. 

Bassetti, B., & Nicoladis, E. (2016). Research on grammatical gender and thought in early and emergent bilinguals. The International Journal of Bilingualism: Cross-Disciplinary, Cross-Linguistic Studies of Language Behaviour, 20(1), 3-16. 

Lambelet, A. (2016). Second grammatical gender system and grammatical gender-linked connotations in adult emergent bilinguals with French as a second language. The International Journal of Bilingualism: Cross-Disciplinary, Cross-Linguistic Studies of Language Behavior, 20(1), 62-75. 

Samuel, S., Cole, G., & Eacott, M. J. (2019). Grammatical gender and linguistic relativity: A systematic review. Psychonomic Bulletin and Review, 26(6), 1767-1786.  

Salience in Second Language Acquisition: A Systematic Review

Saioa Cipitria (Vrije Universiteit Brussel), Georgia Knell (Vrije Universiteit Brussel), Esli Struys (Vrije Universiteit Brussel), Ludovic De Cuypere (Vrije Universiteit Brussel) and Alex Housen (Vrije Universiteit Brussel).  

In the field of second language acquisition (SLA), salience refers to the extent to which a linguistic element stands out from its context, and, hence, the likelihood of it being picked up by the learner’s perceptual apparatus. The degree of salience of a particular linguistic feature is believed to drive attention, essential for input (i.e., available stimuli) to become intake (i.e., noticed stimuli) and thus to acquire (i.e., cognitively register) it (Schmidt, 1990). 

The concept of salience in SLA has gained in prominence since the meta-analysis by Goldschneider and DeKeyser (2001), who identify relevant factors for the order of morpheme acquisition in English (e.g., phonological/orthographic substance, frequency, etc.). These factors comprise different aspects of salience. Ellis (2016) classifies salience as experiential (i.e., related to language experience), psycholinguistic (i.e., pertaining to the linguistic context and the learner), and perceptual (i.e., concerning the form). Empirical research on the role of salience in L2 learning remains scarce, in particular concerning the manipulation of salience as an independent variable (yet see contributions to Gass et al., 2018). This systematic review therefore explores the conceptualization and operationalization of salience by reviewing research that directly explores and manipulates salience in the context of L2 language acquisition by considering different types of salience (i.e., experiential, psycholinguistic, and perceptual) and how they affect language processing.

We extracted 993 references from two databases (Web of Science and Scopus), of which 88 studies were accepted for more detailed annotation. Other coding criteria include salience type and the linguistic aspect under study (morphosyntax, lexicon, phonology, etc.). Based on cognitive theories in both SLA and cognitive psychology, we expect a positive correlation to emerge from our review of empirical studies between salience, attention and ultimate acquisition. We further consider how this correlation is mediated by the various types of salience. Finally, we discuss general methodological trends, strengths and shortcomings in research on salience in SLA research as they emerge from our survey. 

Secondary 1 pupils’ spontaneous EFL oral acquisition: the advantage of modelization

Marie-Ange Dat (Nantes Université), Delphine Guedat-Bittighoffer (Université d'Angers) and Rebecca Starkey-Perret (Nantes Université).  

The field of second language (L2) acquisition in guided settings regularly raises the issue of oral fluency. While a broad scientific consensus identifies automation as the key to L2 oral proficiency (DeKeyser, 2009; Paradis, 2009), the question of how to achieve this remains  unanswered. From recommendations of using explicit knowledge (DeKeyser, 2003) to proposals to build skills in the classroom from a modeled language that draws on learners' internal grammar (Germain, 2018; Ellis & Wulff, 2020) by appealing to implicit memory and the process of proceduralization (Paradis, 2009), a wide range of hypotheses remains open.

In our study context (2019-2022), which focuses on 179 French speakers beginning to learn English as a foreign language during the first year of middle school (11.3 years old), we wanted to measure their level of acquisition by comparing (T1 first month of the school year, T2 9 months later) two teaching methods during one school year. The first one represents prototypical foreign language teaching, i.e. an eclectic set composed of the action-oriented approach (CEFR) and explicit elements of the L2. The second is the Neurolinguistic Approach (also called NLA, Germain, 2018) developed in Canada. The primary principles of the approach are : modeling, implicit learning, primacy of oral and authentic situations of interlocution. We ask ourselves whether learning spontaneous speaking based on modelization and implicit memorisation in the context of extensive teaching (4h per week) is more effective in terms of spontaneous interaction than L2 more explicit approaches. The 179 pupils were in 8 classes: 4 experimental classes teaching with NLA, and 4 prototypical control classes. We assessed each pupil's ability to interact orally using the Oral Proficiency Interview (OPI) scale (Macfarlane & Montsion, 2016), the purpose of which is to assess speaking skills in real-life situations. A repeated measures ANOVA analysis shows that highly significantly (p<0.001) pupils in classes that use modelization and implicit learning double their spontaneous interlocution scores compared to more prototypical classes that use explicit learning pathways. 

References

DeKeyser, R. (2003). Implicit and explicit learning. In C. Doughty & M. Long (Eds.), Handbook of second language acquisition (pp. 313–48). Oxford, England: Blackwell.

DeKeyser, R. (2009). Cognitive-psychological processes in second language learning. In M. Long & C. Doughty (Eds.), Handbook of second language teaching (pp. 119–38). Oxford, England: Blackwell.

Ellis, N. C. & Wulff, S. (2020). Usage-based approaches to L2 acquisition. In VanPatten, B., Keating, G. D., & Wulff, S. (Eds.), Theories in Second Language Acquisition: An introduction. (pp. 63-82). New York & London: Routledge.

Germain, C. (2018). The Neurolinguistic Approach (NLA) for Learning and Teaching Foreign Languages – Theory and Practice. Cambridge Scholars Publishing.

MacFarlane, D., & Montsion, L. (2016). Comparaison d’échelles de compétence orale. Ministère de l’Education nationale et du Développement de la Petite Enfance du Nouveau-Brunswick. 

Paradis, M. (2009). Declarative and Procedural Determinants of Second Languages. Amsterdam: John Benjamin.  

How do learner-internal and learner-external individual differences affect adolescent learners’ L2 English speaking development? - A dense longitudinal study.

Vanessa De Wilde (Ghent University).  

Researchers have investigated the role of individual differences in second language learning and found that learners’ development is influenced by both internal differences (e.g. cognitive differences) and external differences (e.g. length of instruction) (Dörnyei, 2014; Muñoz, 2012). Many of these studies investigated second language learning in adults. Recently, some studies have also looked into the role of individual difference variables in language learning in young learners (e.g. De Wilde & Goriot, 2022; Leona et al., 2021) but only few studies have adopted a longitudinal approach (e.g. Pfenninger, 2022).

In the present study we aim to analyze the role of several internal differences (such as cognitive differences and motivation) and external differences (such as out-of-school exposure and instruction) in L2 English speaking development in adolescent learners. In order to be able to capture real-life speaking development, we conducted a dense longitudinal study with frequent measurements of L2 speaking skills.  Learners in the first year of secondary school (11 to 13 years old, n=48) did a speaking task every school week from September to May (25 weeks). The study took place in Flanders and the Netherlands. The participants came from three class groups which were different in terms of instruction: one group had received L2 English instruction in primary school, the second group had just started with formal L2 English lessons and the last group did not have any formal English lessons yet. At the start of the study the participants did multiple tasks looking into individual differences (prior knowledge of English, Dutch vocabulary knowledge, reasoning skills, working memory capacity, motivation, and out-of-school exposure). Generalized additive mixed models (GAMMs) were used to model (non-linear) learning trajectories over time and to investigate the role of internal and external differences in this development.  Results show that learners’ speaking scores are influenced by time and prior L2 English knowledge. Initial differences in learners’ prior knowledge seem to be the main predictor of L2 English speaking skills. Detailed results and implications of the study will be discussed during the presentation and suggestions will be done for future studies. 

References 

De Wilde, V., & Goriot, C. (Eds.). (2022). Second Language Learning Before Adulthood: Individual Differences in Children and Adolescents (Vol. 65). Walter de Gruyter GmbH & Co KG. 

Dörnyei, Z. (2014). The psychology of the language learner: Individual differences in second language acquisition. Routledge. 

Leona, N. L., van Koert, M. J., van der Molen, M. W., Rispens, J. E., Tijms, J., & Snellings, P. (2021). Explaining individual differences in young English language learners’ vocabulary knowledge: The role of Extramural English Exposure and motivation. System, 96, 102402. 

Muñoz, C. (Ed.). (2012). Intensive exposure experiences in second language learning. Multilingual Matters. 

Pfenninger, S. E. (2022). Emergent bilinguals in a digital world: A dynamic analysis of long-term L2 development in (pre) primary school children. International Review of Applied Linguistics in Language Teaching, 60(1), 41-66.  

Investigating native language attrition of late Spanish/English bilinguals in the UK: the case of aspectual marking in Spanish

Laura Dominguez (University of Southampton), E Jamieson (University of Southampton), Glyn Hicks (University of Southampton) and Monika S. Schmid (University of York).  

We investigate possible attrition (i.e. changes in the native grammar) of 30 late Spanish-English bilinguals who have settled in the UK for over 15 years. Specifically, we explore whether aspectual marking is a vulnerable area for the native grammar of these speakers using a multi-method approach to data collection and analysis. We chose this grammatical property as it is predicted to undergo attrition by the Attrition via Acquisition (AvA) model (Hicks & Domínguez 2020) as Spanish and English share the same syntactic/semantic features, but they map differently onto morphosyntactic forms. That is, Spanish speakers do not need to acquire a new feature/structure when using these verb forms in English, but they need to determine in which contexts each aspectual form is grammatical or ungrammatical. Attrition has also been attested (e.g. Cuza (2010)) for this construction for a group of Spanish-English bilinguals in the USA. 

As Table 1 shows, Spanish and English each have two ways to express an event in the present (whether the event is ongoing or habitual) but they differ in the specific contexts in which each of these forms can be used. In English, the simple present form is used for generic/habitual actions, while the periphrastic be+-ing form is used is in the progressive/ongoing aspect. In Spanish, there is an overlap in the aspectual interpretations of the two verb forms: the corresponding estar+-ndo form is also used in the progressive/ongoing aspect, but the simple present can be used for both generic/habitual actions and the progressive/ongoing aspect. 

English                 Spanish Ongoing   Now she is running. Ahora ella está corriendo.                                 Ahora ella corre. 

Habitual  Every day she runs. Todos los días ella corre. 

Table 1. Distribution of verbal forms to express ongoing and habitual actions in English and Spanish 

We predicted that attrition, if attested, would manifest in a preference for the estar+-ndo form over the simple present in ongoing situations as this is the only form available in English in this context. 

We elicit three types of data using a self-paced reading task (online), an acceptability judgment task (offline), and sociolinguistic interviews (oral). These results were compared with similar data from a group of monolingual Spanish speakers who also took part in the study. Modelling the results of the experimental tasks shows no overall attrition at group level. However, two individuals show an English-like pattern in the judgment task. The oral production data also show that, again at an individual level, some speakers may undergo attrition of their aspectual system as a result of the influence of English manifested as a preference for the periphrastic construction in representing ongoing aspectual interpretations in the L1. 

These results support the idea that, in general, morphosyntactic features of an L1 are robust even after extensive exposure to an L2; however, attrition is possible for some speakers at least when the underlying features are shared. 

References 

Cuza, A. (2010). On the L1 attrition of the Spanish present tense. Hispania, 256-272. 

Hicks, G., & Domínguez, L. (2020). A model for L1 grammatical attrition. Second Language Research, 36(2), 143-165  

Investigating sensitivity to partial rules in English wh-questions: Japanese vs. Vietnamese L2 learners

Nigel Duffield (Konan University), Ayumi Matsuo (Kobe College) and Trang Phan (University of Languages & International Studies, Vietnam National University - Hanoi).  

Our project is concerned with L2 learners’ sensitivity to exceptions to the general rules of (short) wh-movement and complementation in English (Chomsky 1957, et seq). (In contrast to the attention paid to constraints on long-distance movement – e.g., L1: Guasti, 1994; L2: Martohardjono, (1993), White & Juffs, 1998 – exceptions to short-movement have been neglected hitherto). The specific focus is on (1) Wh-interactions with finiteness (She knew [why she should eat more vegetables]. ~ She knew [*why to eat more vegetables]; (2) Wh-interactions with negative adverbials (“?What1 under no circumstances2 should3 you … ~ ??Under no circumstances2 what1 should3 you... ~*Under no circumstances2 should3 what1 you ... ~ ?What1 should3 under no circumstances2 you ... give to a dog.”; cf. Haegeman 2012); (3) Wh-interactions in ‘Sluice-Stranding contexts’ (“She asked to attend to something, but I can’t remember what/*what to/??to what?”; Culicover 1998); (4) Raising vs. Control complements (“Amy is *uncertain/certain to lock up. Amy is uncertain/?certain how to lock up.”; cf. Yoshimura et al., 2016). This phase of the project compared Vietnamese and Japanese L2 learners of English: though both L1s are wh-in situ languages, structural commonalities between Vietnamese and English predict an advantage for Vietnamese learners in conditions that probe knowledge of the “left periphery”. 

Method

The current experiment used two formats for eliciting judgments: a pen-and-paper judgment task (P-AJT); and an online version, involving the same materials (O-AJT). Each participant was presented minimal contrast sentences, differing only in the critical contrast, where the former are designated ‘grammatically acceptable’, the latter ‘marginal’ or ‘grammatically unacceptable’ in the literature. Participants were asked to rate the acceptability of each trial on a Likert scale from -2 (‘unacceptable’) to +2 (‘perfectly natural’). 

Items

The stimulus items comprised 348 sentences across four conditions, plus 16 control items (8 minimal pairs, clear grammaticality contrasts). Using a modified Latin Square, six randomized-trial subsets were prepared, each comprising 58+16 items. 

Participants

24 adult native-speakers and 166 L2 learners participated in the study. However, a high percentage of the L2 learners failed to distinguish acceptable from unacceptable items in the control condition [5]. Results from the remaining participants {56 VN; 29 JA} were entered into the main analysis. 

Results

Preliminary descriptive results from the first two conditions (plus control condition) bear out our predictions: on each of the specific contrasts, Vietnamese L2 learners’ judgments pattern with those of native-speakers, albeit in attenuated form, whereas Japanese L2 learners either fail to distinguish between grammatical and ungrammatical sentences, or else diverge sharply from both groups in preferring a clearly marked option (#2-213 order). Full results and inferential statistics will be presented at conference. 

Selected References

Culicover, P. (1999). Syntactic Nuts: hard cases, syntactic theory, …  Oxford: OUP. 

Duffield, N. (2018). Reflections on Psycholinguistic Theories… Cambridge: CUP. 

Haegeman, L. (2012). Adverbial Clauses, Main Clause Phenomena…Oxford: OUP. 

White, L., & Juffs, A. (1998). Constraints on wh-movement …. In S. Flynn et al.

Yoshimura, N., Nakayama, M.,  et al.  (2016). ‘Control and Raising constructions in early L2 English Acquisition’, Second Language, 15. (pp. 53-76).  

What Do Elicited Imitation Tasks for Young EFL Learners Measure?

Akiko Eguchi (Ritsumeikan University) and Remi Murao (Nagoya University).  

Elicited imitation (EI), a testing technique that makes the participants listen to a series of sentences and repeat them verbatim, has been widely used to assess a second language (L2) from various perspectives (e.g., Yan et al., 2016 for a review). The underlying assumption is that EI tasks require learners to process and reconstruct the stimulus sentences with their internal grammar and are applicable to a range of proficient learners by adjusting the stimuli. If this assumption is evidenced, EI can be a promising method to examine scant L2 grammatical knowledge in young learners as repeating phrases and sentences is a familiar activity in classrooms. Compared with numerous L2 adult studies, however, research on using EI with young L2 learners is scarce, and is limited to testing the effects of particular pedagogical interventions by pre-post comparison (e.g., Campfield & Murphy, 2014). Therefore, it is unclear what exactly EI tasks for young learners measure and to what extent primary memory—which is beneficial to rote repetition and differentiated from speech-processing memory—affects EI performance. 

This study clarified what EI tasks measure for young L2 learners by focusing on three variables: syntactic knowledge, vocabulary knowledge, and primary memory. We created 24 sentences for EI tasks using familiar words and various syntactic structures (e.g., “SVO,” “WH do SVO”) that appeared in their textbooks but adjusted to longer sentences to avoid rote repetition (M = 9.4 syllables). EI performance was assessed using a 6-point-scale scoring, assuming that L2 sentential knowledge develops from a single word to a syntactic phrase to a sentence. 

A total of 138 Japanese elementary students (Grade 4 and 5: aged 9–11), who were learning English as a foreign language (EFL), participated in four tasks: (1) EI tasks, (2) untimed sentence correction tasks (SCT) (e.g., *I soccer play. *What color you like?) for syntactic knowledge, (3) vocabulary size test (VST) (Sato, 2019), and (4) digit span task (DST) for primary memory capacity. To determine the extent to which variance in EI scores could be explained by the individual variables, a linear mixed effects model was performed. The results of the final best-fitting model revealed a significant main effect of DST (estimate = 1.72, t = 2.93, p < .01) and the interaction between SCT and VST (estimate = 7.21, t = 4.12, p < .001). This result suggests that EI can successfully measure young L2 learners’ internal grammar integrated with vocabulary knowledge while controlling for the potential confound of rote repetition. 

References

Campfield, D., & Murphy, V. A. (2014). Elicited imitation in search of the influence of linguistic rhythm on child L2 acquisition. System, 41, 207–219.

Sato, T. (2019). Developing a vocabulary size test for Japanese elementary school students. JES Journal, 19, 130–146.

Yan, X., Maeda, Y., Lv, J., & Ginther, A. (2016). Elicited imitation as a measure of second language proficiency: A narrative review and meta-analysis. Language Testing, 33, 497– 528.   

Acquisition of quantifier-negation scope and grammatical morphemes by Chinese learners of Korean

Shuo Feng (Peking University), Meiling Yu (Peking University), Nai'An Ren (Peking University) and Kailun Zhang (Peking University).  

The present study aims to explore the acquisition of universally quantified objects under negation and grammatical morphemes in Korean by Chinese learners. In English, (1) has the not>all reading as the dominant reading. Chinese allows negation and quantifier NP each to precede or follow the other as in (2a-2b), with different scope interpretations. Korean, a SOV language, allows the quantifier NP to precede negation and the preferred reading for (3a) is the all>not reading. However, when the morpheme changes from the object case marker –(l)ul in (3a) to the contrastive focus marker –(n)un in (3b), the preferred reading for (3b) becomes the not>all reading (Park & Dubinsky, 2020). Previous L2 studies found that L2 learners were under strong L1 influence in acquiring quantifier-negation scope and Korean morphemes were challenging to acquire (e.g., Ahn, 2015; Chung & Shin, 2022). However, previous work has focused on a limited set of structures (e.g., quantified subjects) and little research has been carried out with participants with L1s other than English. 

(1) English:          Tom didn’t send all the letter.    (not>all) (2) Chinese: a. Tangmu  meiyou    ji         suoyou   de        xin.     (not>all)         Tom       not-have  send    all         Poss   letter b. Tangmu  suoyou   de        xin      dou   meiyou     ji.     (all>not)         Tom        all         Poss    letter    all     not-have   send (3) Korean: a. Tom-ka      motun  phyenci-lul  an   ponay-ss-ta.      (all>not)         Tom-Subj  all         letter-Do    not   send-Pst-Decl b. Tom-ka      motun  phyenci-nun  an   ponay-ss-ta.    (not>all)         Tom-Subj  all         letter-Cf       not   send-Pst-Decl 

Figure 1. Sample display in the partitioned set context in the SPRT (see the attached pdf.) 

Method

Forty-four Chinese learners of Korean and 30 native Korean speakers completed a self-paced reading task (SPRT), a Korean proficiency task and an acceptability judgment task (AJT). Take the SPRT as an example. Each trial in the SPRT consisted of four steps (see Figure 1). Different scope readings were manipulated in Step 2 where the pictures showed how many objects were affected by an action (either 3 out of 5 or none). Test stimuli were constructed in a 2 x 2 design with Context (2 levels: partitioned set and full set) and Morpheme (2 levels: –(l)ul and –(n)un). Four counterbalanced presentation lists were created using a Latin square block design. The task contained 16 targets, 10 controls and 30 fillers. 

Findings & Conclusion

NNSs were native-like in interpreting quantifier scope in negated sentences with –(l)ul. NNSs’ Korean proficiency played an important role in interpreting quantifier scope, but the influence of proficiency in acquiring –(n)un as a contrastive focus marker was restricted. Although NNSs were not native-like in interpreting sentences with the two morphemes in the offline task, they showed native-like processing pattern in the SPRT. The divergence of NNSs’ performance between the SPRT and the AJT is further discussed in light of NNSs’ less robust encoding of morphological information and increasing processing load in the AJT. 

References

Ahn, H. (2015). Second Language Acquisition of Korean Case by Learners with Different First Languages. Doctoral dissertation, University of Washington.

Chung, E. S., & Shin, J. A. (2022). Native and second language processing of quantifier scope ambiguity. Second Language Research

Jiang, N., Novokshanova, E., Masuda, K., & Wang, X. (2011). Morphological congruency and the acquisition of L2 morphemes. Language Learning, 61(3), 940-967.

Park, K., & Dubinsky, S. (2020). The effects of focus on scope relations between quantifiers and negation in Korean. Proceedings of the Linguistic Society of America, 5(1), 100-106.  

Priming structural similarity in L1-L2 interactions: The use of Present Perfect vs. Simple Past by German learners of English

Regina Grund (University of Würzburg) and Barış Kabak (University of Würzburg).  

The contrast between the English Present Perfect (PP) vs. Simple Past (SP) often poses a challenge for German learners of English, due to not only fuzzy boundaries between the two, but also the presence of structurally similar albeit not always functionally equivalent constructions in German: Perfekt (er hat gekocht ‘he has cooked’) and Präteritum (er koch-te ‘he cook-ed’). While Perfekt is the most commonly used past reference in spoken German, its contextual equivalent in English may correspond to either PP or SP. Although such overlaps might predict an overuse of PP in SP contexts, recent research has shown the opposite: German learners underuse PP in comparison to English monolinguals, possibly due to PP being more complex than SP (e.g., Fuchs et al. 2016). Such corpus-based studies however mainly focused on frequencies and did not investigate to what extent the L1-L2 overlap may influence target-like suppliance in L2 English. 

Here, we investigated potential priming effects of the German Perfekt and Präteritum on the use of PP and SP in L2 English. L1 German university students of English (n=24) were given a written gap-filling task with 24 translanguaging exchanges, each containing a German turn followed by an English response. The latter included a gap requiring the tense that was either structurally similar to the primed tense in German (prime-response match: Perfekt-PP or Präteritum-SP) or dissimilar (prime-response mismatch: Perfekt-SP or Präteritum-PP). 

Mismatch (Perfekt-SP) example: • Was hast du denn so gemacht? [Perfekt primed] • Last week, I ___ (to make) a cake. [SP expected] 

While an increase in the correct suppliance due to syntactic copying was predicted in the match conditions, syntactic copying was expected to provoke incorrect suppliance in the mismatch conditions. Thirteen (near-)native lecturers of English grammar served as controls to verify whether the English turns were PP- or SP-lenient. 

Preliminary results revealed that the error rates were 10% for the match vs. 28% for the mismatch condition. When the German prompt matched the expected English tense, there were more errors in the Präteritum-SP (13%) than in Perfekt-PP (6%) context, suggesting facilitative priming effects for Perfekt. When the prime and the expected tense mismatched, however, 33% of answers in Präteritum-PP were erroneous while the error rate in Perfekt-SP was 23%. Error analyses revealed that while the majority of errors in Präteritum-PP were unrelated to priming (56%), the most frequent error in Perfekt-SP was the incorrect suppliance of PP (76%), implying Perfekt’s unfavorable effect. 

We suggest that a primed Perfekt strongly activates PP in English and is more likely to induce CLI effects than Präteritum. As such, corpus-based SLA studies should be complemented with experimental approaches tapping possible L1-L2 interactions in order to offer a more comprehensive account of equivalence classification in L2 grammar learning, with broader teaching implications where L1 priming can be leveraged for dynamic noticing and feedback. 

Reference

Fuchs, R., et al. (2016). The present perfect in learner Englishes: A corpus-based case study on L1 German intermediate and advanced speech and writing. V. Werner et al. (Eds.), Re-assessing the present perfect. Mouton.  

The interpretation of verbal moods in Spanish: A close replication of Kanwit and Geeslin (2014)

Aarnes Gudmestad (Virginia Tech), Amanda Edmonds (Université Côte d'Azur), Carlos Henderson (University of Lund), Christina Lindqvist (University of Gothenburg) and Şebnem Yalçın (Bogazici University).  

Studies on sociolinguistic variation in a second language (L2) demonstrate that learners develop similar sensitivity to the linguistic and extralinguistic factors that influence variable patterns in their input (Geeslin & Long, 2014). This research has largely focused on learner production or on the selection of variable forms through preference tasks, meaning that little is known about how learners develop the ability to interpret variable structures. In the first study of its kind, Kanwit and Geeslin (2014) investigated the interpretation of a variable structure by L2 learners, focusing on the interpretation of verbal moods (the subjunctive and indicative) in Spanish adverbial clauses. They analyzed L2 development using cross-sectional data and found distinct patterns among lower- and upper-level learners. Crucially, Kanwit and Geeslin (2014) provide researchers with a methodological model for investigating the interpretation of variable forms. 

Given that research on the interpretation of sociolinguistic variation is in early stages, we do not yet know whether Kanwit and Geeslin’s (2014) findings generalize to other learner populations. One approach to verifying research findings is to replicate key studies (Porte & McManus, 2019). We thus conducted the first close replication of Kanwit and Geeslin (2014), modifying one variable—the participants’ first language (L1). We collected data from learners of Spanish with three different L1s. The L1s chosen—French, Swedish, and Turkish—are typologically distinct and have yet to receive much attention in variationist second language acquisition (SLA). This means that investigating these understudied language combinations allows us to explore the generalizability of previous results. 

We collected data from 40 participants in each of the three L1 groups. All participants were enrolled in undergraduate-level Spanish courses. We used Kanwit and Geeslin’s (2014) interpretation task and grammar test and designed our own background questionnaire, as it was crucial for us to obtain detailed information on the participants’ language background. Other than adding a new independent variable (participant’s L1), we followed Kanwit and Geeslin’s (2014) data coding and analysis procedures. The dependent variable was the response for each item on the interpretation task. There were seven independent variables: mood, verbal morphology regularity, clause order, adverb/conjunction, proficiency (measured by the grammar-test score), course level, and participant. We analyzed the data quantitatively and found that learners’ interpretation of verbal moods in Spanish differed as a function of their L1. 

Our study advances knowledge in SLA in three ways. First, we respond to calls for replication studies in SLA (Porte & McManus, 2019). Second, we contribute to the expansion of variationist SLA by conducting much-needed work on the interpretation of variable forms. Third, our study impacts variationist SLA and SLA more broadly by expanding the L1-L2 combinations investigated. 

References

References Geeslin, K. L., & Long, A. Y. (2014). Sociolinguistics and second language acquisition: Learning to use language in context. Routledge.

Kanwit, M., & Geeslin, K. L. (2014). The interpretation of Spanish subjunctive and indicative forms in adverbial clauses: A cross-sectional study. Studies in Second Language Acquisition, 36, 487-533.

Porte, G., & McManus, K. (2019). Doing replication research in applied linguistics. Routledge.  

Investigating the linguistic and social effects of the first year of schooling on the grammar of child heritage speakers: focus on Polish heritage children in the UK

Anna Hart (University of Southampton), Laura Dominguez (University of Southampton) and E Jamieson (University of Southampton).  

Three main accounts have been proposed to explain heritage speaker acquisition: incomplete acquisition (Montrul, 2008; Polinsky, 2006; Silva-Corvalan, 2003), attrition (Polinsky, 2011) and parental input effects (Pascual y Cabo, 2018; Montrul & Sanchez-Walker, 2013; Pires & Rothman, 2009). Few studies have focused on children heritage speakers (Laskowski, 2009: Montrul 2016) although this group of speakers experience a major shift in their linguistic input and social context around the age of 4: they start school in the majority language.

In this study we focus on the acquisition of Case by a group of 30 Polish heritage speaker children (HSC) in England in order to try to tease apart these three possible accounts.  Starting school is a crucial event in the life of HSC: it signals the exposure to the majority language and a significant change in their linguistic input; it also signals a change in their social environment. Hence, we investigate whether starting school in the majority language is a key event with significant consequences for heritage language acquisition, i.e. whether the heritage children’s grammatical system is complete or not before they enter mainstream education in the majority language, and/or whether any changes can be observed after one year of schooling. 

We focus on a group of 30 young Polish heritage children in the UK and investigate their development of Case marking during their first year in primary school. Polish marks Case on nouns with specific morphological forms in a consistent and uniform fashion which is not found in English. Previous studies have shown that adult Polish heritage speakers have problems with particular Case markings (e.g. locative) (Koźmińska, 2015; Wolski-Moskoff, 2019), but not others (i.e. nominative). We tested the productive command of these forms using a narrative retelling task and appropriate grammatical knowledge using an acceptability judgement task. Sociograms were used to investigate the heritage children’s social networks. The children were tested twice, at the start and at the end of their first year in primary school. The Bilingual Language Experience Calculator (Unsworth, 2013) was used to calculate how much input (quantity and quality) they receive. Out of the three tested cases, (nominative, genitive and locative), nominative remains ‘intact’ throughout the first year. There is evidence of attrition for genitive, as 20% of the children show lower accuracy with this case that was 100% accurate at the start. Locative is the most challenging case as the mean accuracy rate is already lower at the beginning of the school year than for the other two cases (92.23%), but it decreases further as the year progresses to 80.54% which points to both incomplete acquisition and attrition for this case. These results reveal that attrition of certain morphosyntactic forms does occur after only 9 months in school but that this change, which is only attested in some cases and for some children, is modulated by both linguistic and social factors.  

L2 Learners with low educational background

Ann-Kristin Helland Gujord (University of Bergen), Åshild Søfteland (Østfold University College) and Linda Evenstad Emilsen (Østfold University College).  

This presentation will report from an ongoing research project on early language development in a hitherto understudied group of adult learners – refugees with little or no previous schooling (e.g., Andringa & Godfroid, 2019; Bigelow & Tarone, 2004; Ortega, 2019; Van de Craats et al., 2006; Young-Scholten, 2013). In this project, Adult Acquisition of Norwegian as a Second Language (ALAN), we track 50 learners’ grammatical, lexical, and pragmatic development during the first year of formal tuition in L2 Norwegian. The learners are participants in the obligatory national Introduction Program, intended to prepare for participation in Norwegian working life or further education. 

To gain insights into how this learner group achieves communicative abilities in an additional language, we collect oral language data four times during the first year of tuition (week 8/16/24/40) through picture descriptions, conversations (with L1-speakers), film retellings (Alone and Hungry, Chaplin, 1936) and collaborations tasks (with other L2-learners). We also collect a broad range of background information. 

For the presentation, we will select around 20 of the participants who have conducted all four measure points and present preliminary analyses of their language development as presented in the Chaplin-film retellings. Considering the emergence of grammatical resources and utterance structures, all participants appear to develop their L2 Norwegian throughout the year; nevertheless, there is great individual variation: While many learner language systems at the final measuring point fit the description of the Basic Variety (Klein & Perdue, 1997, Perdue, 1982; see also Dimroth, 2018), not all participants develop their L2 beyond the Pre-Basic Variety during the year. The individual variation will be discussed considering background variables such as age, years of schooling, literacy, trauma history. 

The longitudinal data provided through ALAN is not only important for generalizability of SLA knowledge, but also for quality in the education offered to LESLLA-learners. Level of education and literacy prior to onset of L2 acquisition are important factors influencing L2 learning, and a prerequisite for quality in the education is that it builds on relevant empirical knowledge of how learners progress towards communicative abilities in the L2. 

References

Andringa, S., & Godfroid, A. (2019). SLA for all? Reproducing second language acquisition. Research in non‐academic samples. Language Learning (Call for participation). 

Bigelow, M., & Tarone, E. (2004). The role of literacy level in second language acquisition: Doesn’t who we study determine what we know? TESOL Quarterly, 38, 689–700.

Dimroth, C. (2018). Beyond Statistical Learning: Communication Principles and Language Internal Factors Shape Grammar in Child and Adult Beginners Learning Polish Through Controlled Exposure, Language Learning, 68(4), 863–905.

Klein, W. & Perdue, C. (1997). The basic variety (or: Couldn’t natural languages be much simpler?), Second Language Research, 14, 301–347.

Perdue, C. (1982). Second language acquisition by adult immigrants: A field manual. European Science Foundation.

Ortega, L. (2019). SLA and the study of equitable multilingualism. Language Learning, 103(1), 23–38. 

Young-Scholten, M. (2013). Low-educated immigrants and the social relevance of second language acquisition research. Second Language Research, 9(4), 441–454.  

Cross-linguistic structural priming as a mechanism of cross-linguistic influence: Asymmetrical effects of L1 activation and inhibition

Carrie Jackson (The Pennsylvania State University) and Holger Hopp (TU Braunschweig).  

Cross-linguistic influence (CLI) is a key characteristic of bilingualism; however, the mechanisms of CLI are still not clear. Recently, several proposals to grammatical CLI have argued that cross-linguistic priming is a key mechanism of CLI in child bilinguals (Nicoladis, 2012; Serratrice, 2016, 2022). We examine the scope of priming as a mechanism of cross-linguistic influence (CLI) in bilinguals by aiming to boost CLI through priming in early and late L2 learners. In two cross-linguistic structural priming studies with 32 less-proficient adolescent (Study 1) and 60 more highly proficient adult German-English learners (Study 2), we assess whether cross-linguistic priming enhances CLI in English for well-formed (adverbial fronting; 1a), dispreferred (TP-LP orders; 1b) and ungrammatical structures (verb-second; 1c; ungrammatical verb raising; 1d). All word orders on the left in (1a-d) are well-formed in German. (1)  a. Last week the man was/ate in the garden.    vs The man was/ate...(adverbial fronting)  b. #The man was/ate last week in the garden. vs in the garden last week.          (TP-LP)  c. *Last week was/ate the man in the garden. vs Last week the man was/ate …     (V2) d. *The man ate often meat.                 vs The man often ate …    (Verb raising) 

Participants were primed with the German word orders (left in (1a-c)) in German, and were then prompted to produce a sentence in English. We measured which of the two respective word orders in (1a-c) they produced in English. L2 learners in both studies showed CLI from German in their English sentence production in a baseline task (Figures 1&2). For grammatical L1-L2 word orders in L2 English, i.e. fronting in (1a) in the priming task, logistic mixed effects modelling shows that less-proficient learners in Study 1 also exhibited short-term cross-linguistic priming, which extended to longer-term priming among the more proficient learners in Study 2 in a posttest.  However, there was no evidence that cross-linguistic priming increased the use of dispreferred (1b, TP-LP orders) or ungrammatical L1-based word orders in L2 English (1c, V2) in either study. Rather, as seen in Figures 1&2, the overall production of these word orders decreased from baseline via the priming task to the posttest. The proportion of ungrammatical word orders not contained in the priming task (1d, Verb raising) did not change from baseline to posttest. Together, these results suggest that, while cross-linguistic priming leads learners to increase the use of shared, grammatical structures, it leads to the inhibition of non-shared, ungrammatical structures in L2 production. We conclude that cross-linguistic priming has asymmetrical effects on CLI of grammatical and ungrammatical L1-based structures in the L2 in that the Interlanguage grammar modulates activation and inhibition of the L1.

References

Nicoladis, E. (2012). Cross-linguistic influence in French–English bilingual children's possessive constructions. BLC

Serratrice, L. (2016). Cross-linguistic influence, cross-linguistic priming and the nature of shared syntactic structures. LAB

Serratrice, L. (2022). What can syntactic priming tell us about crosslinguistic influence? In Syntactic priming in language acquisition: representations, mechanisms and applications. Benjamins.  

It’s about time: Exploring refugee L2 learners’ narratives of time through the lens of agency

Julia Jakob (Faculty of Education, University of Cambridge). 

“We wrote a mock exam [in German class today]. And it was very easy. I asked the teacher if I could go home. Because attendance is mandatory. […] She said yes [this time]. […] But [normally] it’s a waste of time. I have to wait until the others are done. And I can’t do anything.” 

Abud’s expressed frustration over his lack of agency to fill his time in ways he considers useful or rewarding is illustrative of the experiences of many refugees. While narratives of time have been explored in sociology and forced migration studies, they remain underresearched in SLA. This is despite data from forced migration studies showing the potential relevance of this phenomenon for how refugee learners feel able to invest in L2 learning. 

Drawing on qualitative data from semi-structured interviews conducted during five months of fieldwork with ten young men who live in Austria as refugees and engage in formal learning of German as an additional language, this paper will discuss participants’ narratives of time that are associated with their experience of learning German as an additional language. 

In particular, the paper builds on Schlote’s (1996) work in the field of critical sociology of time to interrogate how investment in L2 learning is affected by: 

  • the concession of control over one’s own time to other actors, such as teachers and the state, in the L2 classroom, the broader Austrian refugee L2 education system, and beyond
  • the expressed inability to plan for an often uncertain future in Austria, e.g. due to long waiting times for asylum, L2 courses, or a non-permanent legal status of protection
  • the experiences of (partial) autonomy in structuring one’s own time under consideration of both the above and individual affordances and challenges, with a particular focus on unstructured time in L2 courses, independent L2 study, and opportunities for L2 use 

The subjective experience of time intersects with L2 investment as conceptualized by Darvin & Norton (2015) on several levels. Firstly, narratives of time need to be understood within systemic patterns of control that form a structure that may, to an extent, be challenged and resisted. Secondly, self-positioning and positioning by others interact with narratives of time through notions of productivity and the pace of L2 learning. Thirdly, an unclear future may render the potential benefits of L2 learning uncertain. 

The findings from this small-scale study indicate the need for future research into L2 learners’ narratives of time in (forced) migration contexts. Due to the relevance for issues of L2 learners’ investment, a better understanding of the subjective experiences of time can inform future critical evaluation of integration policymaking and L2 course design and pedagogy. 

References: 

Darvin, R., & Norton, B. (2015). Identity and a Model of Investment in Applied Linguistics. Annual Review of Applied Linguistics, 35, 36–56.

Schlote, A. (1996). Widersprüche sozialer Zeit: Zeitorganisation im Alltag zwischen Herrschaft und Freiheit. VS Verlag für Sozialwissenschaften.  

Can adults learn L2 grammar after prolonged exposure under incidental learning conditions?

Panagiotis Kenanidis (Friedrich-Alexander-Universität Erlangen-Nürnberg), Ewa Dabrowska (FAU Erlangen), Miquel Llompart (Universitat Pompeu Fabra) and Diana Pili-Moss (Leuphana University of Lüneburg).  

While late second language (L2) learning is assumed to be largely explicit, there is some evidence that adults are able to acquire the grammar of novel languages under incidental learning conditions (Rebuschat et al., 2021; Ruiz et al., 2018). In two experiments, we revisit the question of whether adults can learn grammar incidentally and investigate whether word order and morphology are susceptible to incidental learning to the same degree. Additionally, we investigate the extent to which learners’ first language (L1) background and metalinguistic awareness of the target grammatical structures can modulate L2 grammar learning outcomes. 

In experiment 1, forty-one monolingual speakers of English took part in a five-session online study, during which they were exposed to Kepidalo, an artificial language that consisted of nouns, verbs and adjectives and had case marking and variable word order: a canonical Subject-Object-Verb order and a non-canonical Object-Subject-Verb order (1). First, participants were explicitly trained on the nouns of the language. This was followed by a two-alternative forced-choice task (2AFCT) consisting of two blocks. In the first block, participants received vocabulary training while being incidentally exposed to grammar. Two videos, each showing two aliens performing an action, were presented, accompanied by a sentence corresponding to one of them (Figure 1). The videos differed in 1) one of the aliens, or 2) the color of an alien or, 3) the action performed. Block 2 served as a grammatical comprehension test. Here, the two videos differed in that the agent/patient roles were reversed, and no feedback was given. The 2AFCT was repeated in the first four sessions. In session 5, grammatical knowledge was further assessed through a final grammatical comprehension test and a grammaticality judgment task (GJT) including both word order and case marking violations. Despite extensive exposure to input, and although performance on vocabulary increased significantly across sessions, learners’ grammatical comprehension showed little improvement over time, and this was limited to Subject-Object-Verb sentences only. Furthermore, participants were significantly better at detecting word order than case marking violations in the grammaticality judgment task (Figure 2). 

Experiment 2 further increased the amount of incidental exposure to the artificial language by providing two additional sessions and was conducted with native speakers of German. Moreover, testing in the last session was followed by the administration of a post-test awareness questionnaire. Leveraging their prior experience with a morphologically richer L1, participants showed better learning of both word order and case marking compared to their L1 English counterparts, yet continued to display substantial difficulties with the latter (Figure 2). Furthermore, grammar learning was found to be contingent on the learners’ level of awareness. 

Taken together, the results of the two experiments underscore adult learners’ difficulty with case marking and point towards the presence of a threshold in incidental L2 grammar learning, which appears to be tightly linked to prior L1 experience (Ellis, 2006; MacWhinney, 2005). In addition, our findings continue to highlight the facilitative role of awareness on L2 learning outcomes (Andringa, 2020; Schmidt, 2012).  

Testing the interface of implicit and explicit L2 grammar knowledge and their reciprocal relationship: A one-year longitudinal study

Kathy Kim (Boston University).  

In recent years, SLA researchers have suggested that the association between explicit (conscious) and implicit (unconscious) second language (L2) knowledge is reciprocal; that is, awareness (or explicit knowledge) not only facilitates the development of implicit knowledge but can also be a product of implicit knowledge (Ahn, 2020; Andringa, 2020; Kim & Godfroid, 2023). In this project, I build on this line of research on the interface question and explore the extent to which the relational strengths between explicit and implicit L2 knowledge change over the course of one year. Do explicit and implicit knowledge remain reciprocal equally after 6 months and 12 months of using the L2 in an immersion context? I use an autoregressive cross-lagged model (ACLM) to address these questions. 

One hundred forty-nine international students at an English-medium university completed five linguistic tests that measured their explicit and implicit knowledge of L2 English. The untimed written grammaticality judgment test (GJT) and metalinguistic knowledge test (MKT) served as measures of explicit L2 English knowledge. The timed written GJT, oral production, and elicited imitation were administered as implicit L2 English knowledge measures. These measures were provided at three timepoints over a year (T1: January–February, T2: April–May, T3: November–December). 

Across the three waves, accuracy scores improved in most measures, with MKT exhibiting the steepest improvement (40% increase in accuracy from T1 to T3) and oral production showing the least (4% increase in accuracy from T1 to T3). The results of two-factor CFA models—that were performed separately for each timepoint data—suggested that all measures meaningfully explained their corresponding constructs at each timepoint. Building on these CFA models, I am currently evaluating the results of an autoregressive cross-lagged model to explore the causal relationship between the constructs. Preliminary ACLM findings suggest that the strongest predictor of both implicit AND explicit knowledge at time 3 may be the previous level of implicit knowledge and not explicit knowledge. Results suggest that the development of advanced linguistic competence is composed of dynamic interaction between explicit and implicit knowledge with implicit knowledge playing a key role. 

The role of working memory and creativity in written task performance

Judit Kormos (Lancaster University) and Shungo Suzuki (Waseda University).  

Although the role of creativity in second language (L2) learning is relevant both theoretically and pedagogically, creativity is an under-researched area in the field of second language acquisition. Writing in another language requires creativity in knowledge creation and transformation as well as the careful orchestration and coordination of cognitive processes. Therefore, creativity can be expected to play a role in content planning, idea generation and lexical selection processes of L2 writing. Recent research in cognitive psychology also suggests that there is a link between divergent thinking (a key facet of creativity) and working memory capacity (Frith et al., 2021). As L2 writing is a cognitively demanding task that draws on existing L2 knowledge and skills, working memory (WM) resources can also exert a substantial impact on the quality of the written product. Nonetheless, research on the role of WM in L2 writing is scarce and has mostly involved either young L2 learners (Michel et al., 2019) or university students (e.g., Mavrou, 2020; Vasylets & Marin, 2021) and has investigated a limited type of writing tasks. Our project is novel in jointly examining the role of creativity and WM in a large sample of Hungarian teenage learners of English using a narrative and argumentative writing task. 

In our research, 100 Hungarian secondary school students aged between 16-18 at B1-B2 level of language proficiency wrote a story based on six unrelated pictures and an argumentative essay. Participants completed a forward and a backward digit span task and an emotional stroop task and a recently standardized and validated creativity test for the Hungarian context (Fáy et al., 2021) which allowed us to examine not only the role of creative fluency and elaboration but also facets of originality. The accuracy, lexical sophistication, diversity and syntactic complexity of the students’ writing was analysed with automated natural language processing tools: TAALES (Kyle & Crossley, 2015); TAALED (Kyle, Crossley & Jarvis, 2021); TAASSC (Kyle, 2016); GAMET (Crossley et al., 2019). The length of the text in words and sentences was the measure of productivity. A member of the team that developed, validated and standardized the creativity test scored the creativity tests using standardized procedures. The impact of creativity and working memory on the lexical diversity, sophistication, syntactic complexity, accuracy and length of the argumentative and narrative texts was examined using structural equation modelling. 

The results reveal a statistically significant relationship between WM and facets of creative originality as well as a significant, albeit weak, role of WM and creativity in the accuracy, syntactic complexity and lexical diversity of learners’ performance in both tasks. The findings also suggest that creative originality is associated with a wider range of performance measures in the narrative than in the argumentative task. The findings of our research can inform L2 writing pedagogy and yield insights into what areas of L2 writing students with lower working memory capacity and creativity might need additional support.  

Une étude exploratoire sur l’impact de la profondeur orthographique et de la complexité morphologique de la L1 sur le traitement morphographique en français

Pierre Largy (Univerisité Toulouse-Jean Jaurès), Vera Serrau (Univerisité Toulouse-Jean Jaurès) and Cecilia Gunnarsson-Largy (Univerisité Toulouse-Jean Jaurès). 

Le degré de profondeur orthographique d’une langue (distance entre orthographe et prononciation) semble solliciter des modalités différentes de traitement orthographique (Katz & Frost, 1992). Dans ce sens, les locuteurs d’une L1 à orthographe opaque comme le français ou l’anglais devraient traiter les mots majoritairement dans leur structure visuo-orthographique (accès direct aux mots), alors que les locuteurs d’une langue à orthographe transparente comme l’italien ou l’espagnol privilégieraient le traitement des unités sub-lexicales (assemblage entre graphèmes et phonèmes ou entre morphèmes). Des études récentes portant sur l’impact de l’orthographe sur la performance en L2 montrent que les locuteurs d’une L1 à orthographe transparente s’appuient davantage sur la forme écrite des mots (orthographe) et sur la voie sub-lexicale en L2, ce qui influence positivement leur performance orthographique en L2 (van Daal and Wass, 2017) mais négativement leur performance phonologique en L2 (prononciation superflue des lettres silencieuses dans des mots de la L2 : Bassetti & Atkinson, 2015). Or, d’une part, l’impact de la profondeur orthographique de la L1 sur le traitement orthographique en L2 semble avoir été rarement étudié. D’autre part, la plupart des études sur ce sujet se focalisent sur le traitement de mots simples et ne prennent pas en compte l'interaction entre profondeur orthographique et complexité morphologique. La présente étude vise à observer si des effets translinguistiques de transfert cognitif attribuables à la profondeur orthographique et la complexité morphologique (richesse de la morphologie flexionnelle) de la L1 émergent lors du traitement morphographique (= traitement orthographique des mots fléchis) en français L2. Nous nous focalisons sur la performance morphographique dans des verbes fléchis homophoniques vs. hétérophoniques (arrive/arrivent vs. part/partent) en tâche de rappel écrit de phrases dictées en Français L2, auprès de deux groupes de 48 apprenants experts (B2/C1) ayant une L1 à orthographe transparente et à morphologie complexe (L1-TR+MC), l’espagnol ou l’italien, vs. une L1 à orthographe opaque et à morphologie simple (L1-OP+MS), l'anglais. Les résultats de notre étude renvoient à un transfert cognitif, lors du traitement écrit des mots fléchis en Français L2, des mécanismes cognitifs développés en fonction de la profondeur orthographique et de la complexité morphologique de la L1. En effet, la performance morphographique des apprenants italophones/hispanophones est significativement supérieure à celle des apprenants anglophones. De plus, la réussite des apprenants italophones/hispanophones ne varie pas en fonction de l’audibilité de l’accord verbal, alors que celle des apprenants anglophones est affectée négativement par l’homophonie verbale. Nous interprétons ces tendances par le transfert à la L2 de l’appui sur l'orthographe et du recours majoritaire au traitement morphologique auprès des L1-TR+MC. Ce transfert favoriserait la performance morphographique en L2, indépendamment de la disponibilité d’indices audibles de discrimination morphologique. En revanche, le transfert à la L2 de l’appui sur l’oral et du recours majoritaire au traitement lexical auprès des L1-OP+MS complexifierait la performance morphographique en L2, tout particulièrement en absence d’indices audibles de discrimination morphologique (homophonie).  

Effect of first language lexicalisation on second language lexical inferencing and acquisition: A study of French-speaking learners of Chinese as a foreign language

Enhao Léger-Zheng (Laboratoire de NeuroPsychoLinguistique, Université de Toulouse Jean Jaurès) and Olga Théophanous (Laboratoire de NeuroPsychoLinguistique, Université de Toulouse Jean Jaurès).  

Lexical inferencing is an important reading and vocabulary learning strategy in second language (L2) acquisition. However, few studies have investigated the effect of lexicalisation, i.e. the existence of an attested lexical equivalent for an L2 word in learners’ native language (L1), on these processes. Paribakht (2005) observed that Iranian learners of English succeed more readily in inferring the meaning of English words lexicalised in Persian than non-lexicalised ones. Paribakht and Tréville (2007), comparing Iranian and French learners of English, found that both groups obtained better inferencing scores for lexicalised words. However French-speaking learners had generally better performances than Iranians in inferring English words. The similarity L1-L2 (French-English) had thus a positive effect on lexical inferencing. 

The aim of the present study is two-fold. First, it focuses on difficulties French-speaking learners may encounter during the inferencing of unknown Chinese words, lexicalised and non-lexicalised in French. Second, it focuses on the subsequent retention of these inferred words. 33 intermediate level learners of Chinese were asked to infer via the Think-aloud protocol the meaning of 20 Chinese words (10 lexicalised and 10 non-lexicalised in French) integrated in two texts written in Chinese. We examined and classified 654 inferences based on the knowledge sources classification by Paribakht (2005), to which we added four specific sources: form, meaning and pronunciation of component character(s) in a Chinese word, as well as the relationship between component characters in a Chinese word.  Contrary to the studies on English L2 mentioned above, our results show that French-speaking learners of Chinese had better inferencing score for non-lexicalised words. Known characters comprised in an unknown Chinese word are likely to help learners understand the meaning of this word partially, if not totally. However, learners obtained more totally correct answers for lexicalised words than non-lexicalised ones.  As for subsequent retention of inferred words, according to the results of the Vocabulary Knowledge Scale administered three times (once pre-inference and twice post-inference, 2 and 4 weeks later), analysed with the Linear Mixed Model, learners had significantly more difficulties in retaining the meaning of a non-lexicalised word in both short and long terms (p=0.01). Besides, our results reveal that learners’ cognitive investment in inferred words via dictionary consulting during the post-inference phase may have positive impact on lexical gains for these words, but only in terms of lexical form, not meaning. This result, contrary to that of Laufer and Rozovski-Roitblat's study (2015) on English L2, is probably due to the particularity of Chinese characters and lexis. 

References

LAUFER, B. & ROZOVSKI-ROITBLAT, B. (2015). Retention of new words: Quantity of encounters, quality of task, and degree of knowledge. Language Teaching Research, 19, 687 711.

PARIBAKHT, T. (2005). The influence of first language lexicalization on second language lexical inferencing: A study of Farsi-speaking learners of English as a foreign language. Language Learning, 55(4), 701 748.

PARIBAKHT, T. & TREVILLE, M.-C. (2007). L’inférence lexicale chez des locuteurs de français et des locuteurs de persan lors de la lecture de textes anglais : Effet de la lexicalisation en première langue. CMLR, 63(3), 399 428.  

A corpus-based approach to 'singular they' in L2 writing

Yi Liu (University of Sheffield) and Thomas Hammond (University of Sheffield).  

This study uses a learner corpus to investigate what features of the antecedent determine use of 'singular they' in naturalistic L2 writing. Background: In English, 'singular they' is used with grammatically singular antecedents where gender is indeterminate (Paterson 2011). It can be co-referenced with both (in)definite specific (a) and non-specific (b) antecedents: 

(a) I gave a/the student top marks, they were happy ([+/-definite, +specific]) (b) If I gave a/the student top marks, they would be happy ([+/-definite, -specific]) 

Antecedents of 'singular they' can also vary regarding notional number (e.g., someone [-plural] vs. everyone [+plural]). Previous L1-English studies found 'singular they' occurs mostly with indefinite antecedents that are non-specific and notionally plural (e.g., any student, everybody) (Paterson 2014). Similar to English, Mandarin Chinese does not have nominal inflection for gender but third-person pronouns do carry gender information when written. Ta (他: he/gender neutral; 她: she) is used for singular referents, and the plural marker -men (们) can be attached to indicate plurality with grammatically plural referents: 他们 (ta-men: masculine/gender neutral) and 她们 (ta-men: feminine). Previous L2-English studies have largely focused on elicited use of 'singular they' by learners of highly gendered L1s (e.g., Russian and Italian), showing potential L1 influence (Stormbom 2022). Compared to those languages, Mandarin Chinese is more similar to English in terms of gender marking, thus, research is needed to test whether this similarity effects learners’ use of 'singular they'.

Research Question: Is Chinese learners’ use of singular they conditioned by the specificity, definiteness, and notional number of its antecedent, and how does this compare to previous L1/L2 studies?

Methodology: Using the TECCL corpus, we examine the distribution of 'singular they' in Chinese university students’ essays (n=417), analysing the form and properties of their antecedents. We found 1000 tokens of 'they' in total and only 7.5% (n=75) were 'singular they'. Results: The antecedents of 'singular they' manifest in four forms: possessive determiners (e.g., our school), definite NPs (e.g., the government), indefinite pronouns (e.g., anyone), and indefinite NPs (e.g., a family). The distributions regarding the features are shown below:  [+definite] 46% [+specific] 17% [+plural] 32% [-indefinite] 54% [-specific] 83% [-plural] 68%

Conclusions: When using 'singular they', there appears to be no preference for the features [+/-definite] but a preference for non-specific, notionally singular antecedents, which differs from previous L1 and L2 studies. This suggests that Chinese learners do not treat 'singular they' as a direct equivalent of ta-men, but instead conceptualise this as a separate pronominal form from plural they. We discuss this in relation to Homonymy Theory (Whitley 1978; Paterson 2014) and L1 influence in the use of L2 epicene pronouns (singular they and generic he) more generally. 

Selected references

Paterson, L. (2014) British pronoun use, prescription, and processing: Linguistic and social influences affecting ‘they’ and ‘he’. London: Palgrave Macmillan.

Stormbom, C. (2022) Singular they in English as a foreign language. Applied Linguistics Review, vol.13, no.5, pp.873-897.  

Exploring vocabulary attrition through network models: Possibilities, findings and potential

Paul M. Meara (Swansea University) and Imma Miralpeix (Universitat de Barcelona).   

The mental lexicon has often been conceptualized as a network where words are interconnected (Aitchison, 1987). However, research so far has only rarely looked at the possible implications of a networked vocabulary, nor has it explored in depth how network structures can give insight into processes such as language attrition (Meara, 2004). Understanding how language is lost is a concern for psychologists, applied linguists and second language acquisition researchers (Ecke, 2004; Schmid & Köpke, 2019). 

This presentation makes use of simulations (i.e., computer models of vocabulary-like networks) using Cellular Automata (Kauffman, 1993), which are minimally organised networks where each word is linked to a small number of other words, and where words change their activity status depending on the inputs they receive from other words. These models can provide useful analogues for real vocabularies and exhibit some emergent properties which are relevant to understand the mental lexicon (Meara & Miralpeix, 2022). In these networks, each word is linked to two other words and are activated (‘on’) or deactivated (‘off’) in response to the input it receives from its two links. Some words are easily activated, while other words are more difficult to activate. 

Different ways of modelling attrition in a vocabulary network are explored, focusing on the results of the simulations and their implications. For example, it is interesting to note that vocabularies do not get smaller if we simply turn some words ‘off’. In the case of complete deactivation of a network, it is possible for it to reconstruct itself, either partially or completely, if the right kind of stimulus is provided. The simulations also suggest that we need to distinguish between attrition events and vocabulary loss events. Not all attrition events lead to immediate vocabulary loss, but an accumulation of minimal attrition events typically and quickly leads to a catastrophic collapse of activity in a network. The work reported here suggests that vocabulary attrition is a more complex phenomenon than it is usually assumed to be and that simulations, which have not been very much used in applied linguistics research, might offer a useful way of investigating it further. 

References 

Aitchison, J. (1987) Words in the Mind: An Introduction to the Mental Lexicon (4th ed.). Blackwell. 

Ecke, P. (2004). Language attrition and theories of forgetting: A cross-disciplinary review. International Journal of Bilingualism, 8(3), 321–254. 

Kauffman, S. A. (1993). The Origins of Order. Self Organisation in Selection and Evolution. Oxford University Press.

Meara, P.M. (2004). Modelling vocabulary loss. Applied Linguistics, 25(2), 137–155. 

Meara, P.M. & Miralpeix, I. (2022). Vocabulary networks workshop 1: Introduction. Vocabulary Learning and Instruction, 11(1), 17–34. https://doi.org/10.7820/vli.v11.1.meara1

Schmid, M.S. & Köpke, B. (2019). The Oxford Handbook of Language Attrition. Oxford University Press.  

Intervening with debates: EFL students re-engaging in an Exploratory Practice classroom

Elizabeth Machin (None).  

The present study offers insights on engagement in an English as a Foreign Language (EFL) classroom through implementation of the Exploratory Practice Principles at a university in Spain. The account sits within the third strand of the L2 Motivational Self System (L2MSS) (Dörnyei, 2009), the recently re-defined ‘L2 learning experience’ (Dörnyei, 2019), and within the broader domain of the construct of engagement (e.g., Seligman, 2011). The need for understanding evolved out of disaffection: the puzzle of student rejection of classroom communication work in pairs / threes.

Student-led debates (a ‘Potentially Exploitable Pedagogic Activity’ (PEPA)) were explored as a device for restoring engagement. Teacher and students together examined the apparent success of the debates within a classroom discussion and perception questionnaires were individually completed by the students. The debate scripts prepared by groups of students, comments gathered by the students during the classroom discussion and individual freestyle commentary within the questionnaires were analysed using thematic analysis (Braun & Clarke, 2006, 2013). Six themes emerged from the discussion and freestyle commentary: (1) Positive nerves (2) We liked it (3) Respect (4) We learnt new things (5) A challenge and (6) A good form. The present study finds that the students demonstrated motivated L2 behaviour in the group writing of their debate scripts and appeared engaged whilst debating. The students were moved to improvise during the debate, with the data suggesting the form of a debate provided welcome scaffolding. 

References

Braun, V., & Clarke, V. (2006). Using thematic analysis in psychology. Qualitative Research in Psychology. 

Braun, V., & Clarke, V. (2013). Successful qualitative research: A practical guide for beginners. Sage.

Dörnyei, Z. (2009). The L2 Motivational Self System. In Z. Dörnyei & E. Ushioda (Eds.), Motivation, Language Identity and the L2 Self (pp. 9–42). Multilingual Matters

Dörnyei, Z. (2019). Towards a better understanding of the L2 Learning Experience, the Cinderella of the L2 Motivational Self System. Studies in Second Language Learning and Teaching, 9(1), 19–30. 

Seligman, M. E. P. (2011). Flourish: A new understanding of happiness and well-being – and how to achieve them. Nicholas Brealey. 

Early bilingual Papiamento-Dutch reading development in a post-colonial context

Gil-Marie Mercelina (Radboud University), Eliane Segers (Radboud University), Ronald Severing (University of Curaçao) and Ludo Verhoeven (Radboud University).  

Most children in a post-colonial context learn to read in an L2, de facto a foreign language. Recent developments in the Caribbean lead to the introduction of Creoles as language of instruction in the classroom, enabling children to learn to read in the L1 as well. In the Dutch Caribbean Leeward islands, children learn to read in either the L1 Papiamento Creole or the L2 Dutch.

In the present study, we examined the effect of initial language of reading instruction (L1 vs. L2), individual differences (kindergarten speech decoding, vocabulary, rapid naming, short-term memory, phonological awareness, letter knowledge, and oral comprehension), and cross-linguistic transfer on reading development of 289 children. We followed a longitudinal design (kindergarten to second grade) in a natural setting of initial reading instruction in L1 vs L2 using ANOVA and SEM analyses. Results showed kindergartners to perform better on all L1 Papiamento precursor skills. Decoding performance depended on initial language of instruction.  L1-instructed children’s Papiamento decoding was predicted by phonological awareness and letter knowledge. L2-instructed children’s Dutch decoding was predicted by phonological awareness, letter knowledge, short-term memory, and speech decoding. Transfer effects were found from the language of decoding instruction to the other language with L1 precursors and decoding skill contributing to the prediction of L2 decoding but not the other way around. As for reading comprehension, instruction language was significant only for Dutch reading comprehension. Kindergarten Papiamento vocabulary and short-term memory along with word decoding predicted Papiamento reading comprehension. As for Dutch reading comprehension, along with word decoding, both groups relied on kindergarten vocabulary, oral comprehension, and short-term memory while Dutch-instructed children also relied on kindergarten letter knowledge. However, oral comprehension weighed more heavily for the L1-instructed children while word decoding weighed more heavily for the L2-instructed children. Regarding transfer, the L1-instructed group relied on kindergarten precursors while the Dutch instructed group relied on first-grade precursors to predict the other language. To conclude, both groups learned to decode regardless of instruction language, but results reveal the pivotal role and benefit of including the L1 in early bilingual reading development in a post-colonial context while early fostering of vocabulary and automatization of decoding can support later reading comprehension.  

Is L2 grit related to absolute levels of language attainment?

Hitoshi Mikami (Chubu University).  

Grit, mental stamina necessary to pursue long-term goals, has recently received the attention of researchers in the field of second language (L2) acquisition (SLA), for the reason that such mental stamina has particular relevance to a long and continuous process like L2 development. Within this trend, Teimouri et al. (2022) introduced the concept of L2-specific grit, and demonstrated the validity of their language-domain-specific measure of grit called the L2 grit scale. The validity of this new scale has also been supported by the results of other studies, and L2 grit is becoming a hot topic in SLA research.

This study, however, argues that there is a limitation in the previous validation studies. That is, the relevance of L2 grit to absolute levels of language attainment has been left unclear due to the use of self-assessed proficiency. From an assessment perspective, it is crucial to clarify how learner characteristics are related to objective L2 proficiency as well as to self-perceived proficiency. In this context, the aim of the present study was to investigate how L2 grit is related to absolute language proficiency.       This study was conducted in cooperation with the English department of a Japanese university. One hundred and six students took a proficiency test and then responded to a questionnaire. L2 proficiency was measured using a test called TOEIC, which is one of the most popular standardized language tests assessing non-native speakers’ English proficiency. TOEIC estimates one’s English proficiency based on the scores of two sections (i.e., reading and listening comprehension sections). Consequently, three types of proficiency scores were obtained from the test: the reading and listening scores (hereafter Reading and Listening), and the overall score (hereafter Overall Proficiency). To explore how grit is related to the proficiency measures, grit and L2 grit were measured using the localized versions of their respective scales. Partial correlation analysis was first carried out to confirm how L2 grit is related to the three proficiency measures (Reading, Listening, and Overall Proficiency) when compared to grit. Control variables were gender and year at university. Following this, a series of hierarchical regression analyses were performed to evaluate the incremental prediction of the test scores from grit when L2 grit was used as the first step regression model.      The main findings can be summarized as follows. First, the results of the partial correlation analyses showed that L2 grit always has stronger correlations with the test scores than grit. Second, the results of the regression analyses confirmed that L2 grit was a consistent predictor of Reading, Listening, and Overall Proficiency. These results not only provide additional support for the validity of the L2 grit scale, but also demonstrate that gritty L2 learners are indeed likely to achieve a higher linguistic level in their target language. 

Reference

Teimouri, Y., Plonsky, L., & Tabandeh, F. (2022). L2 grit: Passion and perseverance for second-language learning. Language Teaching Research, 26(5), 893–918. 

Comprehension of prefaced disagreements in French by advanced L2 learners

Simone Morehed (Université de Fribourg).  

This paper presents a study on the comprehension of disagreements by advanced L2 learners of French. The aim is to study at which moment in the sequential context they understand the upcoming disagreement, and through which (para)verbal resources. Conversation analysis studies describe disagreements as “non-preferred actions”, since they generally go against the social cohesion and thus are prefaced with hesitations, pauses and mitigators (Schegloff 2007). The preface signals the upcoming “problem” and delays and attenuates the disagreement. The preface’s importance is shown in psycholinguistic studies stating that L1 speakers start deciphering the incoming utterance from its start (Barthel et al. 2016). The above studies raise the questions on how L2 learners understand different types of disagreement markers in preface position: which (para)verbal resources do they attend to, and is the preface sufficient to identify the disagreement or is more context needed? What are the similarities and differences between L2 learners and L1 speakers of French? For L2 learners’ interactional and pragmatic competences, comprehension is crucial, without which the learner has difficulties interacting appropriately (Taguchi & Roever 2017).

We analysed disagreements from spoken authentic interactions of corpora of L1 French with the method of conversation analysis. The disagreement markers were categorized and include different types, from monosyllabic hesitation markers like “ben”, to constructions like the partial agreement “oui mais” and the epistemic stance “je sais pas”. Based on the analyses, an experiment was designed and conducted with 200 adult advanced L2 learners of French (L1 German or Italian), and 100 L1 speakers. The experiment contained two tasks: an online questionnaire and a stimulated recall. The questionnaire included authentic spoken disagreement sequences in L1 French. The length of the initial disagreement utterance was altered to investigate where in the sequential context the disagreement was understood: from the first part of the preface to the whole utterance. The participants answered written questions reflecting their comprehension of the disagreement. In the stimulated recall, the participants motivated their answers, where we analyse which (para)verbal resources in the disagreement sequences they identified and attended to. This paper will present the results that are currently being analysed. The first results show that the L2 learners generally need more interactional context to understand the disagreement than the L1 speakers, for whom (part of) the preface often is sufficient. In the stimulated recalls, all three participant groups often mention verbal resources, whereas paraverbal resources, e.g. intonation, are mentioned less. The ongoing analyses will investigate in detail how the different markers are interpreted, and further compare the three participant groups. The results can give insights into L2 learners’ comprehension difficulties in authentic oral interactions, and specifically which disagreement markers pose problem or are interpreted differently from L1 speakers. 

References

Barthel, M., Sauppe, S., Levinson, S. C., & Meyer, A. S. (2016). The timing of utterance planning in task-oriented dialogue: Evidence from a novel list-completion paradigm. Frontiers in Psychology, 7. 

Schegloff, E. A. (2007). Sequence Organization in Interaction. Cambridge: Cambridge University Press. Taguchi, N., & Roever, C. (2017). Second language pragmatics. New York: Oxford University Press. 

Do executive function capacities mediate noticing during face-to-face oral interaction?

Jonathan Moxon (Saga University).  

Interaction is considered to be facilitative of L2 learning (Long, 1996) and a key role is played by corrective feedback in the processes of noticing (Schmidt, 1990) that mark the beginnings of interaction-driven learning. The importance of attention in the processes of noticing suggest noticing is mediated by individual differences in cognitive capacities, and previous research indicates that global working memory capacities may play a role.

This research extends this viewpoint by focusing on the microprocesses of working memory supervision, namely executive function (Miyake et al., 2000).   The role of three executive functions, namely shifting, updating, and inhibition during interactionally situated noticing were examined using two tests to measure each of these microprocesses. Three approaches to the measurement of noticing were used, namely stimulated recall, modified output, and a methodologically innovative approach which used a novel measure which combined these two data sets into a single variable where noticing was considered to have occurred when either participants reported noticing or noticing the gap during stimulated recalls, and/or produced modified output in the light of feedback. Modified output data was further informed by response time data (Moxon, 2020), where unusually longer response times were considered indicative of noticing regardless of uptake type and shorter response times were coded no noticing, again regardless of uptake type.  The target structure investigated was past counterfactual conditional antecedent and main clause formulation.   Simple linear regression analyses indicated that noticing as measured by the response time measure in if clause-focused episodes was predicted by shifting capacities, providing some support for the research hypothesis that greater cognitive flexibility in shifting between meaning and form-focused task sets results in a higher likelihood of noticing corrective feedback. In contrast, updating and inhibition were not found to be significantly associated with any of the noticing measures. 

References

Long, M. H. (1996). The role of the linguistic environment in second language acquisition. In W. C. Ritchie & T. K. Bhatia (Eds.), Handbook of Second Language Acquisition (pp. 413–468).

Miyake, A., Friedman, N. P., Emerson, M. J., Witzki, A. H., Howerter, A., & Wager, T. D.(2000). The unity and diversity of executive functions and their contributions to complex “frontal lobe” tasks: A latent variable analysis. Cognitive Psychology, 41(1), 49–100.

Moxon, J. (2020). The efficacy of uptake response times as a measure of noticing of corrective feedback during oral interaction feedback episodes. Kumamoto University Studies in Social and Cultural Sciences, 18, 75–98.

Schmidt, R. (1990). The role of consciousness in second language learning. Applied Linguistics, 11(2), 129–158. 

Processing reduced speech in the L1 and L2: A combined eye-tracking and ERP study

Kimberley Mulder (Amsterdam Center for Language and Communication, University of Amsterdam), Sophie Brand (Radboud Teacher Academy, Nijmegen), Lou Boves (Center for Language Studies, Radboud University Nijmegen) and Mirjam Ernestus (Center for Language Studies, Radboud University Nijmegen).  

In everyday speech, words are often reduced. Native listeners generally understand reduced forms effortlessly (e.g., Ernestus, Baayen & Schreuder, 2002), but learners can have problems understanding reduced forms (Ernestus, Dikmans & Giezenaar, 2017; Nouveau, 2012; Wong et al., 2015). 

We investigated the effect of reduction in a passive listening visual world task with French natives and Dutch learners of French. Participants listened to French sentences containing either a reduced or full form of a target noun with a schwa in the first syllable (e.g., /rkɛ̃/ vs. /rəkɛ̃/ for requin ‘shark’). The targets were presented in the middle of sentences and were not predictable from the preceding context. The sentences ended with a noun phrase that was semantically related to the target. One second before the start of the sentences four line drawings were shown on the screen, one of which corresponded to the target word. At the end of the sentences a photograph of a scene was shown, and participants had to decide whether the scene depicted the content of the sentence. Eye movements and EEG were recorded simultaneously throughout the experiment. We used eye fixations to define alternative time-lock moments -in addition to the onset of the target words- for the analysis of the EEG signals around the target words. In addition, we developed novel combinations of fixations and saccades as additional predictors in Linear Mixed-Effects Models of the EEG signals. We found a stronger effect of reduction on phonetic processing and semantic integration in learners than in natives, but the effects are different from the N100/N400 and P600 effects found in previous research. Interestingly, we found significant effects of the eye movements and fixations in the one second preview of the four line drawings, as well as in a time window covering the target words. Also, the LME models with time-lock defined by a fixation on the target picture were different from the models with time-lock on target word onset. We model how the use of visual information affects the auditory processing. 

References

Ernestus, M., Baayen, H., & Schreuder, R. (2002). The recognition of reduced word forms. Brain and language, 81, 162 ̶ 173.

Ernestus, M., Dikmans, M., & Giezenaar, G. (2017). Advanced second language learners experience difficulties processing reduced word pronunciation variants. Dutch Journal of Applied Linguistics, 6(1), 1-20.

Nouveau, D. (2012). Limites perceptives de l'e caduc chez des apprenants néérlandophones. Revue Canadienne de Linguistique Appliquée, 15 , 60-78.

Wong, S. W. L., Mok, P. P. K., Chung, K. K.-H., Leung, V. W. H., Bishop, D. V. M., & Chow, B. W.-Y. (2015). Perception of native English reduced forms in Chinese learners: Its role in listening comprehension and its phonological correlates. TESOL QUARTERLY, 51, 7-31. 

i-lex: an improved method of assessing L2 learner ability to see connections between words?

Ian Munby (Hokkai Gakuen University).   

Knowing a word’s associations is considered an aspect of word knowledge (Nation, 2001). It follows that L2 learner ability to see connections between words may improve with gains in vocabulary knowledge and proficiency. However, WA research in L2 has not produced a conclusive link between WAT performance and proficiency (Wolter, 2002) and existing WATs suffer from methodological weaknesses. For example, receptive WATs allow testees to guess word associations through elimination processes without knowing why the words are associated. With productive WATs, scoring responses with word association norms is problematic because results depend on the cue words and norms lists used (Schmitt, 1998). The objective of this study is to circumvent these problems by developing a new WAT inspired by Meara (1994). He suggests presenting learners with the three most common associates of a cue word and asking them to supply the missing word. Following this format, a new WAT called i-lex (Munby, 2017) was developed using sets of three cue words (CWs) chosen from the five most common associates to 50 target words (TWs) listed in the Edinburgh Associative Thesaurus, or EAT (Kiss et al, 1973). 

In the first study (i-lex v1), on average, a group of 25 native speakers outperformed an experimental group comprising 98 Japanese learners of English. An unpaired t-test between the mean scores of the two groups yielded a significant difference of t= 11.153, p < 0.0001. Further, to probe a possible link between i-lex performance and vocabulary knowledge, non-native i-lex scores were compared with a kanji translation test (Webb, 2008). Pearson correlations among these scores are r=.729 (1-sided p value, significant at p < 0.01). In a follow-up study (i-lex v2), i-lex scores of 164 Japanese L2 users of English were compared with the same kanji translation test and an additional measure of vocabulary knowledge: Levels 1-5 of the New VLT (McLean and Kramer, 2015). This yielded significant positive correlations of r=.729 and r=.827 (p < 0.01) respectively, suggesting that the ability of this group to see links between highly frequent English words is related to their vocabulary knowledge. A comparison of i-lex v2 initial scores and retest scores after 2 weeks yielded r= .871 (p< 0.01) indicating that the results are reliable. 

References

Kiss, G.R., Armstrong, C., & Milroy, R. (1973). An Associative Thesaurus of English. EP Microfilms, Wakefield.

McLean, S., & Kramer, B. (2015). The creation of a new vocabulary levels test. Shiken, 19(2), 1–11. 

Meara, P.M. (1994) Word associations in Spanish

Munby, I. D. (2017). I-lex v1 and v2: An Improved Method of Assessing L2 Learner Ability to See Connections between Words? Vocabulary Learning and Instruction, 6(1), 75-90.

Nation, I.S.P. (2001). Learning vocabulary in another language. Cambridge: Cambridge University Press. Schmitt, N. (1998). Quantifying word association responses: What is native-like? System, 26, 389-401.

Webb, S. (2008). Receptive And Productive Vocabulary Sizes Of L2 Learners. Studies in Second Language Acquisition, 30, 01,79-95.  

Crosslinguistic influence and proficiency in L2 and L3 knowledge of aspect in Japanese

Akbar Nadjar Hendra (University of York), Leah Roberts (University of York) and Heather Marsden (University of York).  

This study investigates prior-language influence in the acquisition of aspectual interpretation. We compare L1-Indonesian L2-English L3-Japanese speakers with L1-English L2-Japanese. Aspect is grammaticized in Japanese and English, whereas it is marked lexically in Indonesian. Previous research (Roberts & Liszka, 2013, 2019) showed that acquisition of grammaticized aspect in L2 was facilitated by grammaticized aspect in the L1. Further, Gabriele (2009) showed that L1-English L2-Japanese speakers could acquire interpretive properties of Japanese aspect that differ from English. The present study expands the focus to L3 speakers, investigating the effect of an L1 with non-grammaticized aspect and L2 with grammaticized aspect on L3 acquisition of aspect. 

The study focuses on interpretation of the Japanese “-te iru” verbal morphology. The canonical interpretation is the progressive (e.g., “tabe” [eat] + “te iru” = is eating). However, with achievement verbs “-te iru” has a resultative interpretation, unlike English “be+ing” (1). In further contrast to English, “-te iru” is incompatible with a futurate meaning, and hence infelicitous with futurate adverbs (2) where the simple non-past is required. Indonesian expresses the different aspectual meanings with distinct pre-verbal adverbials. 

1. Hikouki-ga kuukoo-ni tsui-te iru Plane-NOM airport-at arrive-TE IRU “The plane has arrived (*is arriving) at the airport.” 

2. Hikouki-ga sugu-ni shuppatsu suru / *shi-te iru Plane-NOM soon departure do / do-TE IRU “The plane is departing soon.” 

If L2/L3 acquisition of Japanese “-te iru” is influenced by English, then the semantics of “be+ing” may (i) obstruct acquisition of the resultative interpretation, and (ii) lead to non-target-like acceptance of futurate interpretations. However, if influence arises from Indonesian, while acquisition of the licit meanings of “-te iru” could be hindered, a futurate interpretation should not be posited. 

Twenty-nine native Japanese, 20 L1-English L2-Japanese and 22 L1-Indonesian L2-English L3-Japanese speakers participated in the experiment. The L2/L3 speakers’ Japanese proficiency was intermediate–advanced, with similar proficiency test score distributions in each group. Participants completed two tests of knowledge of “-te iru”. An adapted replication of Gabriele’s (2009) sentence compatibility task included accomplishment verbs in the “-te iru” form following contexts that depicted completed events (i.e., grammatical resultative), or ongoing events (ungrammatical). An acceptability judgement task (AJT) compared grammatical resultative “-te iru” with ungrammatical “-te iru” after futurate adverbs. 

Ordinal mixed-effects models of the ratings for each task showed that both non-native groups differentiated between grammatical and ungrammatical interpretations, but the magnitude of differentiation was significantly smaller than in the native group (p<.001). Notably, the L2/L3 ratings on ungrammatical conditions were higher than in the native group. Follow-up models on the L2/L3 groups with proficiency score as a predictor showed a weak interaction in the AJT of proficiency with condition (futurate v. resultative) and group (b=–0.065; p=.076), due to lower ratings for the futurate in the L3 group than in the L2 group, as proficiency increased. We argue, drawing on models of L3 acquisition, that this relatively more target-like behaviour on the futurate by the higher proficiency L3 speakers suggests an effect of L1 Indonesian influence during L3 development. 

Insights into the Appropriate Use of Machine Translation in Foreign Language Education

Mikie Nishiyama (Postgraduate School, Tokyo Healthcare University) and Noriko Matsuda (Kindai University).  

Because of the shift in many university classes to an online format due to the COVID pandemic, computer-mediated online educational environments will continue to be normal for EFL (English as a foreign language) learners. Under such circumstances, it might be better to incorporate the use of machine translation (MT) into foreign language (FL) education at universities. An overview of previous research shows that there are still few studies involving questionnaire surveys of EFL learners regarding the use of MT in FL learning (Briggs, 2018), and that the educational effects of using MT as a supporting tool for learners in FL learning require further investigation (Lee, 2020). Moreover, learners’ attitudes to the use of MT are also unknown, and it is necessary to find out what they think about its use in language study (Gally, 2019). In order to propose a method to use MT in FL education, it is necessary to examine how learners use MT, how learners can effectively use MT for FL learning, and how instructors involved in FL education should appropriately respond to learners’ use of MT.

In light of the above, the present study conducted a questionnaire survey of 245 Japanese EFL learners regarding the use of MT in FL learning, such as translation sites and translation applications, to investigate: (1) how they use MT, (2) their awareness of language forms between the target language and their first language when using MT, and (3) their perceptions toward using MT. The participants’ responses to the questionnaire survey were subjected to exploratory factor analysis, and four factors were identified: 1. first language use pre-edit, 2. English-Japanese comparison post-edit, 3. awareness, and 4. trust in MT. Also, based on the responses to the questionnaire items about (1), EFL learners’ MT uses were classified into four clusters, which we refer to as the pre-edit only cluster, post-edit only cluster, both pre-edit and post-edit cluster, and neither pre-edit nor post-edit cluster. This study analyzes the relations between the four clusters of learners’ use of MT and the questionnaire items of (2) and (3). Based on the results of the analysis, this study suggests the usefulness of using MT as a supporting tool in FL learning, and the perceptions that may be related to learners’ use of MT. The appropriate use of MT in FL education is discussed. 

References

Briggs, N. (2018). Neural machine translation tools in the language learning classroom: Students’ use, perceptions, and analyses. The JALT CALL Journal, 14(1), 3–24. 

Gally, T. (2019). The implication of machine translation for English education in Japan. Language Teacher Education, 6(2), 1-14.  Lee, S.-M. (2020). The impact of using machine translation on EFL students' writing. Computer Assisted Language Learning, 33(3), 157–175. 

A core metadata schema for L2 data

Magali Paquot (Université catholique de Louvain), Alexander König (CLARIN ERIC), Egon W. Stemle (Eurac Research) and Jennifer-Carmen Frey (Eurac Research).  

The main objective of this presentation is to introduce a core metadata schema for L2 production data, more particularly learner corpora, which is the result of extensive collaboration between learner corpus compilers at two European institutions and a research data infrastructure expert and member of CLARIN's metadata taskforce. 

The project stems from the recognition that one area that would benefit significantly from standardization is L2 data description, which includes metadata at the level of the dataset as a whole and metadata used to describe the individual learners and task types/registers the corpus is meant to represent. There are a number of reasons why this is important. First, standardized and well-structured metadata increases the findability and usability of existing learner corpora. Second, it should enhance the comparability of datasets and comparability of L2 studies, provided researchers agree on a common set of definitions. Extensive metadata that follow - at best - a standardized vocabulary, and have a strong focus on findability, accessibility, interoperability, and reusability (FAIR) are an essential aspect of FAIR research data (Wilkinson et al. 2016). 

In continuation of Granger & Paquot (2017), our proposed metadata schema is divided into a number of different sections for Corpus metadata (itself divided into Administrative metadata (e.g. authors or license) and Corpus design metadata (e.g. date and place of collection or type of task)), Text metadata (fine-grained per-text information), Learner metadata (details about the learners, e.g. age, languages spoken), Annotation metadata (e.g. details about manual or automatic annotation), Annotator metadata (e.g. professional and language background), Transcriber metadata (e.g. native language or language repertoire) and Task metadata (e.g. instructions, time constraints). While basic information about learners (authors) and language samples (texts) are typically found as part of metadata associated with a learner corpus, other aspects such as those related to the annotation or transcription procedure or the specificities of a task are often found elsewhere (e.g. corpus manual) or are just absent from currently available learner corpora. Our proposal is to provide a systematic description of all these aspects as part of core metadata. 

A beta version of the core metadata schema was tested on a number of learner corpora representing a variety of learners and language samples. After a first presentation to the community at a domain-specific conference we are now in the process of revising our initial proposal based on the comments received and will release a stable version of the core metadata schema in Spring 2023. 

References

Granger, S. & Paquot, M. (2017). Towards standardization of metadata for L2 corpora. Invited talk at the CLARIN workshop on Interoperability of Second Language Resources and Tools, 6-8 December 2017, University of Gothenburg, Sweden. 

Wilkinson, M. D. et al. (2016). The FAIR Guiding Principles for scientific data management and stewardship. Scientific data, 3(1), 1-9.  

Autism and Cognition in Bidialectalism

Kakia Petinou (Cyprus University of Technology) and Kyriakos Antoniou (Cyprus University of Technology).  

Speakers of two languages (bilinguals) or dialects (bidialectals) often exhibit smaller vocabularies in each language separately compared to monolinguals. In contrast, some (controversial) evidence suggests that bilinguals enjoy benefits in aspects of non-verbal cognition (e.g., executive control) [1]. Moreover, Autism Spectrum Disorder (ASD) is often linked to language and other intellectual disabilities, with language being more adversely impacted [2]. Finally, work with ASD bilinguals has generally shown that bilingualism does not have any further detrimental effects on ASD individuals’ cognition, with few studies even reporting positive bilingual cognitive effects [2].

Here, we focus on bidialectalism—on bidialectal experiences such as second-dialect proficiency, use and degree of dialect switching—and examine how its interaction with autistic traits affects fluid intelligence and second-dialect vocabulary in neurotypical young adults.  Sixty-two bidialectals who spoke Cypriot Greek (native dialect) and Standard Greek (SG) as second dialect (mean age=20.5, SD=5.3) took the Autism-Spectrum (AQ) and Systemizing Quotient (SQ) [4] for autistic traits, a Language Background and Dialect Switching Questionnaire, a SG vocabulary test, the WASI [5] and Standard Progressive Matrices (SPM) intelligence tests [6]. We formed composite scores for related variables by averaging relevant z-transformed measures (in parentheses): Autism (AQ, SQ total) and IQ (WASI, SPM).  An analysis on SG vocabulary, with Autism, SG use, Switching, the Autism by SG use and Autism by Switching interactions as predictors showed only a negative effect of Autism (F(1, 42)=10.96, p<.01). A similar analysis on IQ, that additionally included SG vocabulary (proficiency) and the Autism by SG vocabulary interaction as predictors, revealed a significant Autism by Switching interaction (F(1, 39)=5.73, p<.05). This indicates that (1) high autistic traits have a negative effect on IQ only at low levels of switching (Figure 1) and/or that (2) switching has a positive effect on IQ only at high autistic traits (Figure 2). The latter interpretation may be in line with claims that bilingualism has positive effects only in individuals who are not at the peak of cognition [1].  While this study is ongoing, our results suggest that bidialectalism does not interact with autistic traits in affecting verbal cognition. We discuss whether the interactive effect of switching and autistic traits on fluid intelligence represents a bilingual advantage. 

References

[1] Bialystok, E. (2017). The bilingual adaptation: How minds accommodate experience. Psychological bulletin, 143, 233.

[2] Peristeri, E., Silleresi, S., & Tsimpli, I. M. (2022). Bilingualism effects on cognition in autistic children are not all-or-nothing: The role of socioeconomic status in intellectual skills in bilingual autistic children. Autism, 26, 2084-2097.

[3] Wheelwright, S., Baron-Cohen, S., Goldenfeld, N., Delaney, J., Fine, D., Smith, R., ... & Wakabayashi, A. (2006). Predicting autism spectrum quotient (AQ) from the systemizing quotient-revised (SQ-R) and empathy quotient. Brain research, 1079, 47-56.

[4] Wechsler, D. (1999). Wechsler Abbreviated Scale of Intelligence. San Antonio, TX: The Psychological Corporation.

[5] Raven, J., Raven, J. C., & Court, J. H. (2000). Manual for Raven’s Progressive Matrices and Vocabulary Scales. Section 3: The Standard Progressive Matrices, Including the Parallel and Plus Versions. San Antonio, TX: Harcourt Assessment. 

Tracking the development in Japanese EFL learners’ alignment activity and topic management in study abroad, virtual exchange, and language classroom

Sajjad Pouromid (Kansai University).  

Interactional competence (IC), defined as the contingent and context-specific use of interactional practices to achieve joint actions by interlocutors in social interaction (Pekarek Doehler, 2019), has attracted a growing body of research. In the realm of second language acquisition, research has drawn on ethnomethodology and conversation analysis to elucidate how L2 learners develop the competencies to participate in social activities. There are studies with both longitudinal and cross-sectional designs with the aims of tracking developments in IC or documenting its facets. Such studies often focus on specific actions and the methods L2 speakers employ to accomplish those actions. Some aspects that have already been explored include turn-taking practices (Cekaite, 2007), opening tasks (Hellermann, 2008), telling stories and responding to them (Pekarek Doehler & Berger, 2018), as well as repairing troubles (Hellermann, 2011). Learning a foreign language is not a monolithic experience, however, as L2 is learnt in various learning environments.

Against this backdrop, this presentation reports on the preliminary results of a longitudinal conversation analytic study comparing the development of IC in three groups of Japanese EFL learners drawing on analogous data collected at regular intervals from synchronous online discussions among members of each group. The first group comprised three participants studying abroad in the US and Canada over a one-year span, and there were three participants in the second group engaged in a language learning virtual exchange (VE) program between their home university and a partner university from the US. The third group also included three participants taking a general English course at their university without a study abroad or a VE component. The analysis focuses on how these novice L2 learners developed their linguistic repertoire and expanded their interactional inventories to perform conversational alignment activity and topic management. The findings suggests that the participants studying abroad in an English-speaking context performed alignment activities earlier than their peers in the other two groups. These participants were able to produce locally contingent turns and demonstrate their active listenership by displaying their understanding of the previous turns and orienting to them. They were also able to manage the emerging topics in their interactions and expand them to sustain the flow of the conversation. 

References

Cekaite, A. (2007). A child’s development of IC in a Swedish L2 classroom. The Modern Language Journal, 91(1), 45–62. 

Hellermann, J. (2008). Social actions for classroom language learning. Clevedon: Multilingual Matters Ltd. 

Hellermann, J. (2011). Members’ methods, members’ competencies: looking for evidence of language learning in longitudinal investigations of other-initiated repair. In Joan K. Hall, John Hellermann & Simona Pekarek Doehler (Eds.), L2 IC and Development (pp. 147–172). Clevedon: Multilingual Matters. 

Pekarek Doehler, S., & Berger, E. (2018). L2 IC as increased ability for context-sensitive conduct: A longitudinal study of story-openings. Applied Linguistics, 39(4), 555-578. doi:10.1093/applin/amw021 

Pekarek Doehler, S. (2019). On the nature and the development of L2 interactional competence: State of the art and implications for praxis. In R. M. Salaberry & S. Kunitz (Eds.), Teaching and testing L2 IC (pp. 25-59). New York: Routledge. 

Adult learner motivation to learn Chinese in second and foreign language contexts

Jie Rao (PhD candidate at University of York) and Bimali Indrarathne (Associate Professor at University of York).  

The last two decades have witnessed an increasing interest among adults in many countries to learn Chinese as a second/foreign language. A few studies have already examined adult learner motivation to learn Chinese in different language contexts, for example, in the USA (e.g., Wen, 2022), the UK (e.g., Mayumi & Zheng, 2021), and in China (e.g., Li & Zhang, 2021). To investigate whether and how L2 motivation of Chinese as a second and foreign language differs in second and foreign language contexts, comparative study needs to be conducted with L2 learners from similar backgrounds, which enables us to better understand the influence of language context on L2 motivation. Therefore, this study aims to investigate the differences in the motivation of adult learners of Chinese in a foreign and second language context by using Dörnyei (2005, 2009)’s Motivational Self System which includes the Ideal L2 self, Ought to L2 self and L2 learning experience as its theoretical framework.

Qualitative method was adopted in this study with semi-structured interviews. Ten participants were recruited from each context by convenient sampling (20 in total). They belonged to the UK, the USA, Mexico, Colombia, Zambia, Nigeria, Zimbabwe, and Pakistan. The interviews were approximately 50 minutes long. The questions were designed based on the three facets of Dörnyei’s L2MSS (ideal L2 self, Ought-to L2 self and L2 learning experience), on which the coding themes were formed as well. Nvivo (software) was used in coding and analyzing the interview. The results revealed that the learning experience is the most powerful factor to boost learners’ motivation in both contexts. For example, attitude towards the target language learning experience, teacher influence on motivation, and interaction with target communities. Both groups showed similar levels of Ideal L2 self, that is, interest in the language and culture, etc. A disparity between the two groups is that participants from the Chinese context posed a much stronger Ought-to L2 Self than those from the UK context. This can be attributed to their higher level of anxiety in speaking to native speakers of Chinese and stronger instrumental motivation such as seeking for better job opportunities in China.

References

Dörnyei, Z. (2005). The psychology of the language learner: Individual differences in second language acquisition. Mahwah, NJ: Lawrence Erlbaum Associates.

Dörnyei, Z. (2009). The L2 motivational self system. In Z. Dörnyei, & E. Ushioda (Eds.), Motivation, language identity and the L2 self (pp. 9–42). Bristol: Multilingual Matters.

Li, M., & Zhang, L. (2021). Tibetan CSL learners’ L2 motivational self system and L2 achievement. System, 97, 102436.

Mayumi, K. & Zheng, Y. (2021). Becoming a speaker of multiple languages: an investigation into UK university students’ motivation for learning Chinese. The Language Learning Journal. 1-15. 

Wen, X. (2022). Chinese language learning motivation: a study of individual-contextual interactions. Journal of Multilingual and Multicultural Development, 1-17.  

Cross-linguistic pervasiveness of Agent-first in passive contexts and effects in L2 grammars

Isabel Repiso (Salzburg University) and Cyrille Granget (Université Toulouse - Jean Jaurès). 

One of the principles governing utterance structure in initial stages, i.e. the so-called Basic Variety, is that the NP-referent with highest control comes first (Klein & Perdue 1997: 315). This semantic principle has not been tested in the grammar of more advanced L2 learners that have benefited from guided instruction. In the aim of addressing this gap, our study analyses utterance structure in productions where agent-first and patient-first are two options. Crucially, French and Spanish share a common repertoire of agent-first constructions, such as SVO with causative verbs (1a) and causative verbal periphrasis (1b), as well as a common repertoire of patient-first constructions, such as passives with ‘être’/‘ser’ (2a), and verbal periphrasis ‘se faire’/‘hacerse’ (2b). The former constructions mark the instantiation of the patient as object, whereas the latter mark it as subject. 

1. a. Le bully l’a agressé / El bully lo agredió [The bully attacked him]     b. Le garçon l’a fait trébucher / El chico lo hizo tropezar [The guy made him tripped]  2. a. Il a été agressé / Él fue agredido [He was assaulted]       b. Il s’est fait gifler / Él se hizo abofetear [He got slapped] 

We designed a visual stimulus in which 50% of the vignettes represented a human figure as agent, and 50% as patient with another animated entity; the latter were our target items. Participants were asked to retell what happened to the human figure. First, we carried out an offline-retelling task targeting the elicitation of intra-typological variation data across L1 French vs L1 Spanish (84 responses obtained from 21 speakers per group). Then, we replicated the task in L2 French with a group of 20 Spanish-speaking learners. Our L1 results show significant differences in the syntactic instantiation of patients: they are saliently marked as syntactic subjects in French and as direct objects in Spanish (e.g., Il s’est fait agresser vs. El bully lo golpeó). This result suggests that constructions in Spanish are stronger governed by the semantic principle Agent-first, even in contexts where a form of passive construction could be expected. Unlikely L1 grammars, the degree to which one argument “controls” the situation was not so clear-cut expressed in L2 French grammars. Learner Varieties exhibit greater frequencies of stative verbs that give to read the syntactic subject as a sort of middle voice’s experiencer (e.g., Il a peur du chien). Beyond this developmental feature, L2 grammars prove to be Agent-first rooted. In all, our results suggest that L2 learners do not efficiently overcome the Agent-first pervasiveness of their L1 in passive contexts, suggesting that the Agent-first principle is not only linked to the learner’s environment and competence but also to the typological differences between the native and the target language. 

References

Klein, W., Perdue, C. (1997). The Basic Variety. Or: Couldn’t natural language be much simpler? Second Language Research 13, 301-347.  

Learning syntactic variation in L2 sentence processing: The role of prediction error

Duygu Şafak (TU Braunschweig) and Holger Hopp (TU Braunschweig).  

Learning and using syntactic variation gives rise to several challenges, as illustrated in the English dative alternation, since learners need to acquire the syntactic options available in an L2 and the different constraints on optionality. In English, most ditransitive verbs can take both double-object dative (DO) and prepositional-object dative (PO). However, this optionality is constrained by gradient selectional restrictions on the type of their arguments: DO-bias verbs, e.g., pay, tend to occur more frequently with DO, while PO-bias verbs, e.g., send, show a probabilistic tendency to prefer PO. Research on L2 predictive processing indicates that adult L2 learners are sensitive to these gradient verb constraints during real-time comprehension [1-2], which suggests that they can extract distributional probabilities of complementation from the L2 input. 

In this study, we explore how L2 learners learn syntactic variation by testing whether and how L2 learners adapt their syntactic predictions when encountering unexpected input that signals shifts in syntactic variation, e.g., when DO-bias verbs occur with the PO structure rather than the DO structure. To this end, we tested adult L1 German intermediate to advanced L2 learners of English (n=48) in a priming experiment using visual world eye-tracking. Participants first read aloud written prime sentences crossing the factors verb bias (DO-bias vs. PO-bias) and structure type (DO vs. PO); see Figure 1. Subsequently, they listened to spoken target sentences while viewing visual displays with an agent referent (the tailor), a recipient referent (the model), and a theme referent (the dress) (Figure 1).  The results of cluster-based permutation analyses showed PO-priming effects: L2 learners generated more expectations for the PO structure when they heard the target verb after PO (vs. DO) primes, as evidenced by a significantly lower proportion of looks to the recipient than to the theme in Figure 2a. The results also revealed surprisal effects of verb bias on PO-priming effects, such that priming effects were larger when the prime structure did not match the bias of the prime verb, i.e., after PO primes with DO-bias verbs than with PO-bias verbs (Figure 2b). However, there were no strong priming effects for the DO structure (Figure 3). 

The overall pattern of results demonstrates that, like for L1 learners and users [3], prediction error drives implicit learning in an L2. We discuss the findings in the context of error-based implicit learning models [4]. 

References

[1] Şafak, & Hopp. (in press). Cross-linguistic differences in predicting L2 sentence structure: The use of categorical and gradient verb constraints. SSLA.

[2] Wolk, Wolfer, Baumann, Hemforth, & Konieczny. (2011). Acquiring English dative verbs: proficiency effects in German L2 learners. In Carlson, Hölscher, & Shipley (Eds.), Proceedings of the 33rd Annual Conference of the Cognitive Science Society.

[3] Chen, Wang, & Hartsuiker. (2022). Error-based structure prediction in language comprehension: Evidence from verb bias effects in a visual-world structural priming paradigm for Mandarin Chinese. J Exp Psychol Learn Mem Cogn, 48, 60-71. 

[4] Chang, Dell, & Bock. (2006). Becoming syntactic. Psychological Review, 113, 234-272. 

Does L1 orthographic depth influence L2 orthographic processing of inflected words?

Vera Serrau (Univerisité Toulouse Jean Jaurès), Cecilia Gunnarsson-Largy (Université Toulouse-Jean Jaurès) and Pierre Largy (Université Toulouse-Jean Jaurès). 

Orthographic depth, i.e. the extent to which an orthography deviates from the “one-grapheme-to-one-phoneme” principle, influences orthographic processing in expert readers/writers (Frost, 1994). In deep orthographies like English or French, L1 writers (Deep_L1) tend to access previously-acquired words as whole visual forms. Conversely, in shallow orthographies like Italian or Spanish, L1 writers (Shallow_L1) tend to sequentially parse sub-lexical units like phonemes and graphemes. Recent research shows that Shallow_L1 “upgrade” the sublexical route to morphological parsing when dealing with complex words (Görgen et al., 2021). L1 orthographic depth seems likewise to influence L2 orthographic/phonological processing. Compared to Deep_L1, Shallow_L1 show a higher L2 spelling performance (van Daal & Wass, 2017) but more frequent L2 orthography-based phonological errors (e.g. silent letters erroneously-pronounced in L2 words: Bassetti & Atkinson, 2015). This suggests that Shallow_L1 transfer their reliance on orthography to the L2.

Our previous study (Author, 2023) on orthographic processing of homophonic vs. heterophonic inflected verbs in L2 French (mange/mangent vs. boit/boivent) revealed that Shallow_L1 having a richly-inflected L1 (Italian/Spanish) rely more on L2 orthography-oriented morphological processing (= same performance for homophonic and heterophonic inflected words), whilst Deep_L1 having a poorly-inflected L1 (English) rely more on L2 phonology-oriented lexical processing (= lower performance while dealing with homophonic inflected words).  This study investigates whether, during L2 morphographic processing (i.e. processing of written inflected word), L1-to-L2 transfer effects attributable to L1 orthographic depth emerge also when comparing L2 writers whose L1s have a similar degree of inflectional richness.

We observe two groups of 60 expert L2 English writers (C1) whose L1 is Italian (IT-L1: Shallow_L1) vs. French (FR-L1: Deep_L1), two richly-inflected languages, dealing with L2 English morphography in a task of written-recall of dictated sentences. We focus on 36 homophonic (passed/past; won/one) vs. heterophonic (based/best; bought/boat) English word-pairs, containing regular (passed; based) vs. irregular (won; bought) inflected verbs. Controlling word-pairs audibility enables us to observe to what extent the compared groups rely on phonological cues during L2 morphographic processing, whereas controlling verb regularity enables us to observe how the compared groups deal with morphologically-decomposable words (effective sub-lexical processing) vs. morphologically-indecomposable words (ineffective sub-lexical processing).  We generally expect L1 orthographic depth to influence L2 morphographic performance. More precisely, IT-L1 should be less affected by homophony than FR-L1, since Shallow_L1 tend to rely less on phonological cues. IT-L1 should likewise outperform FR-L1 when dealing with L2 word-pairs including regular verbs. Indeed, processing regularly-inflected words sublexically should help Shallow_L1 distinguish the verb (pass-ed) from its uninflected homophonic/heterophonic counterpart (past/pest). Conversely, FR-L1 should outperform IT-L1 when dealing with L2 heterophonic word-pairs including irregular verbs, since the distinction between the verb (bought) and its heterophonic counterpart (boat) should benefit from the reliance on phonological cues (dominant in Deep_L1), and not from the reliance on sub-lexical processing (dominant in Shallow_L1).  The data collection was conducted via an ad hoc-built privacy-preserving keystroke-logging website (OnLog), and statistical analyses are currently being performed (results in progress).  

A longitudinal study of preschool learners’ L1-L2 vowel production

Šárka Šimáčková (Palacky University Olomouc), Václav Jonáš Podlipský (Palacký University Olomouc) and Monika Kučerová (Palacký University Olomouc).  

Bilinguals, including L2 learners, represent speech sounds in one shared phonological system [1] which changes dynamically with experience [2]. Interactions of L1~L2 sound categories are bidirectional [3]. L1 sounds can drift toward L2 sounds because of immersion [2] or even few-week-long classroom learning [4]. They can also move away from similar L2 categories [5]. Most cross-language effects are reported for late(r) learners; young simultaneous bilinguals in bilingual communities keep equivalent sound categories separate [6]. Little is known about preschool classroom learners; longitudinal studies are also scarce. 

This longitudinal study focused on L1&L2 vowel production by 7 Moravian-Czech preschoolers (aged 3;9-5;8) attending weekly EFL classes (45-minute exposure to SSBE) for at least 10 months. We asked: (1) Are the learners’ L1&L2 vowels separated acoustically? (2) Did production of L1 and/or L2 vowels change over time? During 3 months, the children attended 8 recording sessions, 2 in Czech (9 weeks apart) and 6 in English. They produced mono-/di-syllabic words in a picture-naming task. The target words (38 English, 24 Czech), selected according to familiarity and imageablity, included the RP vowels /i,ɪ,ɛ,æ,ʌ/ and Moravian-Czech short /i,ɛ,a/; the remaining English and Czech monophthongs occurred in fillers. 

Normalised vowel height, F1-F0, and retraction, F2-F0, (all in ERB) were modelled by two linear mixed-effects models with Vowel and Time and their interaction as the fixed effects (/æ/ in Time 1 as intercept), and Speaker (varying intercepts and slopes for Time) and Word (varying intercepts) as the random effects. In Time 1, English-/æ/ (typically merged with Czech-/ɛ/ in late(r) learners [7]) differed reliably from the nearest categories: both in height and retraction, it was in-between English/Czech-/ɛ/’s on the one hand, and Czech-/a/ and English-/ʌ/ on the other. Reliable shifts between Times were found in height for Czech-/ɪ/, which raised (unlike English-/ɪ/), for Czech-/ɛ/, which lowered (and started overlapping with English-/æ/), and for English-/ʌ/, which raised (and became more centralised, unlike Czech-/a/). 

In summary, we found that a difficult L2 category, /æ/, was surprisingly well differentiated from the neighbouring vowels. The longitudinal changes suggest that for preschoolers even limited input can lead to measurable phonetic shifts to accommodate L1&L2 categories in the shared vowel space. 

References

[1] Flege & Bohn (2021). The revised speech learning model. In R. Wayland (Ed.), Second language speech learning: Theoretical and empirical progress, 3-83. CUP.

[2] Sancier & Fowler (1997). Gestural drift in a bilingual speaker of Brazilian Portuguese and English. J. Phonetics, 25(4), 421-436.

[3] Kartushina et al. (2016). How and when does the second language influence the production of native speech sounds: A literature review. Language Learning, 66(S2), 155-186.

[4] Chang (2012). Rapid and multifaceted effects of second-language learning on first-language speech production. J. Phonetics, 40(2), 249-268.

[5] Flege (2007). Language contact in bilingualism: Phonetic system interactions. Laboratory Phonology, 9(353-381).

[6] Guion (2003). The vowel systems of Quichua-Spanish bilinguals. Phonetica, 60(2), 98-128.

[7] Šimáčková & Podlipský (2018). Production accuracy of L2 vowels: Phonological parsimony and phonetic flexibility. Research in Language, 16(2), 169-191. 

The Instructed Learning of Form–Function Mappings of L2 English Generic NPs

Neal Snape (Gunma Prefectural Women's University), Helen Zhao (The University of Melbourne) and Menghan Wang (The University of Melbourne). 

An interest in providing instruction to L2 learners about article semantics has been approached recently by adopting a generative approach to SLA (GenSLA). Our study, like Umeda et al. (2019), provides instruction on English genericity to upper-intermediate and advanced L1-Japanese L2-English learners. However, instruction was designed and conducted as an experimentalized online tutor (Zhao & MacWhinney, 2018) that implements random assignment, practice tasks, and immediate feedback. Participants were randomly divided into a treatment group (n = 20) and a control group (n = 17). Participants’ TOEIC scores were used as the measure of L2 proficiency. A pre-test, training sessions and two post-tests were administered that measured knowledge of English generics based on Krifka et al.’s (1995) classification of kind NP-level and sentence-level generic sentences. Only the definite singular and bare plural can be used for NP-level generics such as a natural kind with a kind predicate like the dodo bird/dodo birds is/are extinct. The indefinite singular and bare plural are acceptable for sentence-level generics, e.g., an orange/oranges has/have lots of vitamin C. The treatment group were trained for generic usage through an online acceptability judgement task with lots of practice trials and metalinguistic feedback (in Japanese). The control group received a comparable amount of time of online English preposition training. The study was programmed in PsyToolKit by a research programmer. Linear mixed effects models were built to analyse group differences in treatment effects across time and moderated by L2 proficiency.

The results from both groups are provided in figures 1-8 below. For the NP-level generics, the treatment group showed a significant growth in the use of the definite singular and bare plurals from pre-test to immediate post-test. Similarly, they significantly improved on the use of the indefinite singular and bare plurals for the sentence-level generics. All improvements were retained in the delayed post-test (two weeks apart). The control group’s performance did not differ over time. L2 proficiency was found to influence treatment effects on the more difficult generic forms (i.e., definite singular for NP-level generics and indefinite singular for sentence-level generics).  In GenSLA, some properties of language are deducible through exposure to primary linguistic data (PLD), as discussed in Rothman (2008). We believe that due to a lack of generic NPs in PLD, ongoing exposure to form-function mappings can help learners to restructure their interlanguage grammars with explicit knowledge becoming implicit knowledge over time. 

References 

Krifka, M. Pelletier et al. (1995). Genericity: an introduction. In G. Carlson & F. Pelletier (Eds.), The generic book (pp.1-125). Chicago: University of Chicago Press.

Rothman, J. (2008). Aspect selection in adult L2 Spanish and the Competing Systems Hypothesis: When pedagogical and linguistic rules conflict. Languages in Contrast, 8, 74-106. 

Umeda, M. et al. (2019). The long-term effect of explicit instruction on learners’ knowledge on English articles. Language Teaching Research, Special Issue: Grammatical Meaning and the Second Language Classroom, 23, 179-199. 

Zhao, H., & MacWhinney, B. (2018). The instructed learning of form–function mappings in the English article system. The Modern Language Journal, 102, 99-119.  

Exploring receptivity to adjectival scales in L2 implicature derivation

Glenn Starr (University of Wisconsin-Milwaukee) and Emilie Destruel (University of Iowa).  

Recent forays in experimental scalar implicature research show that native speaking participants derive implicatures far more often in scenarios where the stronger term on an adjectival scale represents a bounded endpoint (Van Tiel et al. 2016). With the statement “This math problem is difficult”, for example, a scale containing the words “difficult” and “impossible” is generated. The term “impossible” denotes a maximal value of the property in question, a feature embedded into the lexico-semantics of the word itself (Kennedy 2007). This promotes a fixed interpretation of the term that can be arrived at independent of context, thus making it clearly distinguishable from its weaker counterpart on the scale. Therefore, when a participant is asked, "Would you conclude that the problem is difficult but not impossible?" the answer will likely be “yes”. However, with the statement “This student is intelligent”, a hearer is much less likely like to conclude that the speaker intended to mean "the student is intelligent but not brilliant". This is because the scale < intelligent, brilliant > contains relative adjectives that better describe ranges rather than endpoints. Delineating a boundary between the two terms necessitates reference to a contextual domain in relation to the object(s) being described. Researchers argue that more effort is required to determine whether implicature calculation is necessary under these conditions because hearers have difficulty (1) identifying a borderline and (2) establishing clear standards which might warrant the derivation (Frazier et al. 2008). 

Some research suggests that language learners do not process certain constructions as efficiently as native speakers due to increased processing costs in the L2. However, researchers have discovered that L2 learners can compensate for this with increased sensitivity to surface-level lexico-semantic cues during sentence processing (Clahsen & Felser 2018). Since the stronger terms on bounded or partially bounded scales contain built-in endpoints which facilitate the demarcation of boundaries between scalemates, this study investigates L2 receptivity to these properties. Native speakers and L2 participants will complete an inference task and decide whether the speaker meant to negate the stronger term. Target bounded/unbounded scales contain pairs of gradable adjectives only, an understudied part of speech in the L2 scalar implicature literature. We also consider whether this phenomenon applies cross-linguistically by testing both L1-French L2-English as well as L1-English L2-French learners. Results will further our understanding of scalar diversity in L2 scalar implicature processing. The experiments are currently underway, and we are eager to present our findings in August. 

References

Clahsen, H., & Felser, C. (2018). Some notes on the shallow structure hypothesis. Studies in Second Language Acquisition, 40(3), 693-706. 

Frazier, L., Clifton C., and Stolterfoht, B. (2008). Scale structure: processing minimum standard and maximum standard scalar adjectives. Cognition 106, 299–324. 

Kennedy, C. (2007). Vagueness and grammar: The semantics of relative and absolute gradable adjectives. Linguistics and Philosophy, 30(1), 1–45. 

Van Tiel, B., Van Miltenburg, E., Zevakhina, N., & Geurts, B. (2016). Scalar Diversity. Journal of Semantics, 33(1), 137–175. 

The German Summary corpus (GerSumCo): A new resource for contrastive research into L2 German of advanced writers

Helena Wedig (University of Antwerp), Carola Strobl (University of Antwerp), Jim J.J. Ureel (University of Antwerp) and Tanja Mortelmans (University of Antwerp).

The German Summary corpus (GerSumCo) is a new corpus for contrastive research into German as a second (L2) vs. first language (L1). GerSumCo was created to investigate cohesion in academic L2 German writing produced by advanced learners. There are several corpora for the contrastive investigation of German learner language available, targeting diverse acquisition levels, text types and L1 backgrounds (e.g., KOLAS: Knorr & Andersen, 2017; Falko: Lüdeling et al., 2008). However, whereas summary writing is an interesting genre for the analysis of cohesion (as seen in Walter, 2007), the only existing corpus of summaries to date is the Falko summary subcorpus (Lüdeling et al., 2008). Preliminary analyses of the Falko summary L2 subcorpus revealed a high degree of patchwriting, i.e., students copy-pasting larger chunks of text from the original text. Since this creates a bias in the data, we decided to compile a new summary corpus.

The specificity of our corpus is twofold: First, students created summaries from two different source texts, i.e., they needed to create their own coherent flow, which diminishes the problem of patchwriting. Second, all summaries were written based on the same source texts and under comparable conditions: All students had to write a summary of two popular scientific texts about a topic related to language variation in contemporary German (e.g., Kiezdeutsch, Mundartdebatte in der Schweiz).  To date, GerSumCo consists of 89 summaries which were written by 42 L2 German students with diverse L1s and 47 L1 German students, with the corpus still growing. The texts were collected at several German Universities during the academic year of 2022-23. For a research project aimed at investigating cohesive strategies deployed by L1 and L2 German writers, the corpus was pre-processed and general linguistic information was added automatically (e.g. part-of-speech). The first analysis of the corpus focuses on connectives as a well-researched cohesive device in learner language. Manual corrections of an automatic pre-annotation via DimLex (Scheffler & Stede 2016; Stede 2002) were conducted by three trained annotators using guidelines based on the PDTB-3 scheme (Webber et al. 2019). The poster will introduce the corpus to the research community and first results of the contrastive analysis of connectives. 

References 

Knorr, D., & Andresen, M. (2017). Commented Learner Corpus Academic Writing (KoLaS). Hamburger Zentrum für Sprachkorpora. 

Lüdeling, A., Doolittle, S., Hirschmann, H., Schmidt, K. & Walter, M. (2008). Das Lernerkorpus Falko. Deutsch als Fremdsprache, 2, 67–73. 

Scheffler, T., & Stede, M. (2016). Adding semantic relations to a large-coverage connective lexicon of German. In Proceedings of the Tenth International Conference on Language Resources and Evaluation (LREC 16) (pp. 1008–1013). ELRA. 

Stede, M. (2002). DiMLex: A lexical approach to discourse markers. In A. Lenci & V. Di Tomaso (Eds.), Exploring the lexicon: Theory and computation (pp. 1–15). Edizioni dell'Orso. 

Walter, M. (2006): Hier wird die Wahl schwer, aber entscheidend. Konnektorenkontraste im Deutschen. Österreichisches Jahrbuch Deutsch als Fremdsprache 2006. 

Webber, B., Prasad, R., Lee, A., & Joshi, A. (2019). The Penn Discourse Treebank 3.0 Annotation Manual.  University of Pennsylvania. 

Multilingual experience results in early noticing and resolution of translation ambiguity in vocabulary learning

Lari-Valtteri Suhonen (University of Borås).   

One difficulty that arises when learning new words in a new language is whether the meaning of new lexical items aligns with pre-existing conceptual categorizations/representations. Previous research (e.g., Eddington & Tokowicz, 2013) refers to the lack of such alignment across languages as translation ambiguity. Jiang (2002) has proposed that learners are largely dependent on metalinguistic awareness in translation ambiguity resolution.   If pre-existing metalinguistic awareness is a factor in learning and using translation-ambiguous items, it should be the case that experienced language learners (i.e., multilinguals) will outperform language learners with less experience. This would be expected to be the case since they – as a part of their previous successful acquisition of additional languages – have experience with the phenomenon. While experienced language learners have been found to be overall better at vocabulary learning than inexperienced language learners (Kausihanskaya & Marian, 2009; Papagno & Vallar, 1995; Van Hell & Mahn, 1997), the present study aims to investigate whether there is an additional effect for specifically translation-ambiguous items. 

This study compared language learners with little experience in foreign language learning (L1 English, N = 30) and experienced language learners (L1 Swedish and L2 English, N = 30). The participants learned both translation ambiguous and non-translation ambiguous items using virtual flashcards with a picture and the target item. During learning, time spent on each virtual flashcard during learning was recorded as were response time and accuracy during assessment.  The experienced language learners were indeed overall faster during learning (est. -617ms, p = < 0.01), corroborating previous research. The difference between the groups decreased over the course of learning (est. 55ms, p = < 0.01). For translation ambiguous items, no significant differences between the groups were found in terms of accuracy. Furthermore, both groups were affected by translation ambiguity during assessment in that they used significantly more time on translation ambiguous items (est. 834ms, p = < 0.01). However, the experienced learners were comparatively less affected by translation ambiguity (est. -513ms, p = < 0.01). It can be hypothesized that this is due to their pre-existing awareness of the phenomena as well as how to deal with it. 

References

Eddington, C., & Tokowicz, N. (2013). Examining English-German translation ambiguity using primed translation recognition. Bilingualism: Language and Cognition, 16, 442–457. 

Jiang, N. (2002). Form-meaning mapping in vocabulary acquisition in a second language. Studies in Second Language Acquisition, 24(4), 617–637. 

Kaushanskaya, M., & Marian, V. (2009). The bilingual advantage in novel word learning. Psychonomic Bulletin & Review, 16, 705–710. 

Papagno, C., & Vallar, G. (1995). Verbal short-term memory and vocabulary learning in polyglots. Quarterly Journal of Experimental Psychology, 48A, 98–107. 

van Hell, J. G., & Mahn, A. C. (1997). Keyword mnemonics versus rote rehearsal: Learning concrete and abstract foreign words by experienced and in experienced learners. Language Learning, 47, 507–546. 

Acquisition of the “that”-trace effect by Japanese learners of English: Examination of the adverb effect and its implications for the theory of the anti-locality

Kasumi Takahashi (Graduate Student, University of Tsukuba) and Yuichi Ono (Faculty of Humanities and Social Sciences).  

This paper discusses the L2 learner’s acquisition of the "that"-trace effect, in which the ungrammaticality originates from the sequence of “that+trace” like “*Whoi do you think [that ti met John]?” The well-known syntactic constraint responsible for this effect is “the anti-locality constraint (ALC)” on wh-movement, which prohibits movement that is too short (e.g., Brillman & Hirsch, 2016; Erlewine, 2020). This constraint further explains the so-called “adverb effect,” which is important evidence. The unacceptability is ameliorated by the intervention of adverbials between "that" and trace, as in “Whoi do you think [ _ that fortunately [ ti met John]]?” (Schippers, 2020). 

ALC is assumed to be a universal constraint. Thus, examining L2 acquisition of "that"-trace effect might investigate the general issue of accessibility to universal knowledge. However, few studies in SLA have examined sensitivity to the adverb effect on the "that"-trace construction. A recent study by Kim and Goodall (2022) examined Korean and Spanish learners of English and concluded that ALC does not work in L2. However, neither the adverb effect nor presence/absence of "that" was not examined in their work, which is an academic gap in our research. Second, previous studies have not considered individual differences in acceptability (Cowart & McDaniel, 2021). Accordingly, our research questions are: (1) Are Japanese intermediate learners of English (JLEs) sensitive to the adverb effect and the presence/absence of "that"? (2) Are there individual differences in sensitivity to ALC observed by JLEs? 

Nineteen Japanese university students with intermediate proficiency levels (CEFR: B2-C1) participated in our study. The factors and conditions were set as follows: [±that (presence/absence)], [±adverbial (absence/presence(word/phrase/clause))], and [+adv_function (sentential/temporal)]. Four tokens were created per condition (n=112 items in total), along with filler items (n=96). 

Main findings are as follows: (1) Main effect of [±that] condition was significant (F(1,18)=8.650, p= .009**), and [–that] was rated higher for both [+adverbial] and [–adverbial], implying the sensitivity to [±that]. (2) Regarding the [+that] condition, there is no significant difference in the acceptability of [±adverb] (F(1,18)=0.0451, p= .834), indicating that no adverb effect was observed in this experiment. After cluster analysis for individual differences into two groups, we found a tendency that the upper-rating group (n=11) prefers [+that] systematically, contrary to the prediction of ALC, while the-lower-rating groups (n=8) prefers [–that] systematically, in accordance with the prediction of ALC. Thus, we conclude that two separate groups gave a systematic complementary judgement in our intermediate level Japanese learners of English. 

References

Brillman, R., & Hirsch, A. (2016). An Anti-Locality Account of English Subject/Non-Subject Asymmetries, CLS 50.

Cowart, W., & McDaniel, D. (2021). The that-trace effect. Goodall G (ed.) The Cambridge handbook of experimental syntax. Cambridge: Cambridge University Press, pp. 258–77.

Erlewine, M. Y. (2020). Anti-locality and subject extraction. Glossa: A Journal of General Linguistics 5: 84. 

Kim, B., & Goodall, G. (2022). The source of the that-trace effect: New evidence from L2 English. Second Language Research. 1-24.

Schippers, A. (2020). COMP-trace revisited: an indirect dependency analysis. ResearchGate. 1-24.  

Indirect effect of orthographic form on phonetic realisation in L2 German: A Corpus study of inflectional endings in spontaneous speech

Megumi Terada (MA Student).  

Recent studies have demonstrated the effect of orthography on L2 speech production (Hayes-Harb & Barrios, 2021). However, most studies have been conducted in experimental settings and little is known about spontaneous L2 speech production, with the exception of Young-Scholten and Langer (2015). To address this issue, a corpus study was conducted on the German schwa in inflectional endings; despite its orthographic representation in the canonical form, it is known that schwa is mainly not realised by German L1 speakers, especially in the final position of the verb (Kohler & Rodgers, 2001).

This study tested if learners of German realise more schwa in inflectional endings than native speakers, as the schwa is canonically presented in the written form. The analysis was based on free dialogues in the CoNNAR corpus which is fully transcribed and aligned with the audio (including 40 dialogues between L1-L1 speakers and L1-L2 speakers of German). The whole dataset consists of 26597 tokens whereas L2 German speakers with L1 English (n = 4) were compared to L1 German speakers (n = 4). The analysis included two types of inflectional endings with schwa: <e> as a marker for the first person singular present indicative and <en> for the first, second, and third person plural present indicative. In the annotation, the realisation of inflectional endings was analysed acoustically and divided into two categories: whether the schwa was realised in inflectional endings or not. In order to avoid arbitrariness in the annotation process, the following criteria were established: 1) if periodic movements were observed in the waveform and 2) if there were visible formant movements and/or glottalisation in the sonagram when the schwa was preceded or followed by a vowel. Where this was the case, the schwa was annotated as 'present' and in other cases as 'absent'. A total of 1252 cases were analysed using generalised linear mixed effects models. The results suggest that L2 speakers produce significantly more schwa in inflectional endings than L1 speakers, with a stronger main effect of the inflectional ending <en>. The results of this study suggest that orthography may also indirectly influence spontaneous L2 speech production. 

References

Hayes-Harb, R., & Barrios, S. (2021). The influence of orthography in second language phonological acquisition. Language Teaching, 54(3), 297–326. 

Kohler, K. J., & Rodgers, J. (2001). Schwa deletion in German read and spontaneous speech. Spontaneous German Speech: Symbolic Structures and Gestural Dynamics, 97–123.  (retrieved: 27.01.2023)

Young-Scholten, M., & Langer, M. (2015). The role of orthographic input in second language German: Evidence from naturalistic adult learners’ production. Applied Psycholinguistics, 36(1), 93–114.   

Linguistic and interactional development of interrogatives in French L2: proficiency or exposure?

Anita Thomas (University of Fribourg) and France Rousset (Institute of Multilingualism, Fribourg).  

Research on the development of interrogatives in L2 spoken French shows that classroom learners at beginner levels tend to overuse formal variants (i.e. subject-verb inversion) as well as the question marker “est-ce que” ‘is it that’ (Dewaele, 1999; Myles et al., 1999). These structures are present but not frequent in native spoken French (Gadet, 1997).  A recent study by Donaldson (2016) showed that the production of interrogatives by advanced learners of French was very similar to what is found in native speakers, suggesting that the level of proficiency has an effect on the type of interrogatives produced by L2 learners.

The aim of the present study is to examine the production of interrogatives from a linguistic and interactional point of view in learners at (low) intermediate level of French who are mainly exposed to French spoken language. It is based on the longitudinal data from seven adult migrants (L1 Tigrinya) who are training in manual professions. They were audio-recorded for ten minutes during free peer-to-peer interactions, four times over one year. The results are based on the manual coding of 320 questions. They show no use of subject-verb inversion and only eight tokens of “est-ce que”. The learners all produced different variants of interrogatives, verbless questions (30%), subject-verb questions (30%) and WH-questions (30%). Moreover, from an interactional point of view, the question-answer sequences included, from the first recording, elements considered central to active listenership (Salaberry & Kunitz, 2019). The results suggest that the linguistic variants and interactional aspects of interrogatives in L2 French is less a factor of language proficiency than of the type of language input. The results are discussed with respect to the role of written and oral language norms in L2 learning. 

References

Dewaele, J.-M. (1999). Word order variation in French interrogative structures. ITL - International Journal of Applied Linguistics, 125(1), 161–180.

Donaldson, B. (2016). Aspects of interrogative use in near-native French: Form, function, and register. Linguistic Approaches to Bilingualism, 6(4), 467–503.

Gadet, F. (1997). Le français ordinaire. (2nd ed.).

Armand Colin. Myles, F., Mitchell, R., & Hooper, J. (1999). Interrogative Chunks in French L2: A Basis for Creative Construction? Studies in Second Language Acquisition, 21(1), 49–80. 

Salaberry, M. R., & Kunitz, S. (Eds.). (2019). Teaching and Testing L2 Interactional Competence: Bridging Theory and Practice. Routledge.  

Explaining academic achievement among international students in HE in the UK: the role of creative coping strategies

Jeanine Treffers-Daller (Emeritus Professor) and Anne Vicary (Honorary Fellow).  

Many international students in Higher Education in the UK struggle to understand the compulsory texts for their course and obtain lower scores for their modules than their monolingual peers. A key reason for this achievement gap is the fact that international students have lower language and literacy skills (Hu and Trenkic, 2021; Trenkic and Warmington, 2018). Here we explore the relationship between vocabulary knowledge, reading behaviour and academic achievement in terms of 30 international students on a postgraduate Law module. They took the 20k version of the Vocabulary Size test (Coxhead, Nation and Slim 2015) and provided summative assignment grades for their Law module and their IELTS scores (overall scores and separate scores for reading). A subgroup of 14 students also took part in qualitative interviews to explore their reading behaviour.  Students’ mean IELTS reading sub-scores were relatively low (6.63), and there were 15 students with reading scores of 6.0 or below. The VST scores appeared to indicate that students had a vocabulary of 10,000 words on average. However, students’ actual vocabulary sizes were probably lower than those indicated by the VST; the poorer sampling of items for the 20k version by comparison with the 14k version may have led to inflated vocabulary size values (Gyllstad, Vilkaitė,  and Schmitt, 2015). The fact that our students’ mean vocabulary scores were higher than those in Trenkic and Warmington (2018), who used the 14k version, while our students’ mean reading scores were lower also suggest that the VST scores in the current sample were inflated. Nevertheless, we found mid strength correlations between the IELTS reading sub-score and the VST (rs = 0.594, p = .002). Contrary to expectations, the module score did not correlate with either the VST or the IELTS reading subscore. Put differently, neither students’ vocabulary sizes, nor their reading levels could explain variance in academic achievement. In-depth interviews revealed that the absence of correlations was likely due to students with IELTS scores of 6.0 or lower using ‘creative’ reading strategies to pass assignments, which included the use of Google translate or similar online tools for reading compulsory texts and writing assignments. We conclude by formulating implications for admissions tutors and teachers in HE. 

References

Coxhead, A., P. Nation, and D. Sim (2015). “Creating and Trialling Six Versions of the Vocabulary Size Test.” The Tesolanz Journal 22: 13-27.

Gyllstad, H., L. Vilkaitė,  and N. Schmitt. 2015. “Assessing Vocabulary Size through Multiple-Choice Formats: Issues with Guessing and Sampling Rates.”ITL-International Journal of Applied Linguistics, 166 (2): 278-306.

Hu, R., and D.Trenkic, D. 2021. “The Effects of Coaching and Repeated Test-Taking on Chinese Candidates’ IELTS Scores, their English Proficiency, and Subsequent Academic Achievement.” International Journal of Bilingual Education and Bilingualism 24 (10): 1486-1501.

Trenkic, D., and M. Warmington. 2018. “Language and Literacy Skills of Home and International University Students: How Different are they, and Does it Matter?” Bilingualism 22 (2): 349–365.  

Assessing syntactic complexity in L2 academic writing: teacher judgments, student ratings and complexity indices

Ineke Vedder (University of Amsterdam).  

The contribution focuses on the assessment of syntactic complexity as a predictor of L2-development in academic writing, by holistic ratings of language teachers and L2 learners, compared with the outcomes on overall complexity measures. For assessing syntactic complexity various indices have been proposed (Bulté & Housen, 2012; Kyle, Crossley & Verspoor, 2021). Overall syntactic complexity has typically been assessed with measures focusing on the length of syntactic units (mean length of T-unit, sentence, clause;  number of Words/T-unit) and clausal subordination measures (Clauses/T-unit, Dependent Clauses/Clause). To a lesser extent, indices for coordination and phrasal complexity have been employed, such as Bardovi Harlig’s Coordination Index (1991) or number of complex noun phrases per T-unit (Biber et al., 2011).   The question arises how far the results obtained with these complexity indices may also be relevant for language pedagogy, and to what extent they correlate with perspectives on syntactic complexity of stakeholders in the L2 classroom. Few studies, so far, have taken a bottom-up approach by analysing judgments of teachers (Kuiken & Vedder, 2019) and L2 learners in the process of acquiring the target language, or by comparing holistic human ratings with the results yielded by automated complexity tools (Granfeldt & Ågren, 2014). Another under-studied area of research is the potential influence of raters’ native or non-native speaker status, which might lead to L1-based preferences for particular syntactic L2 features and a major leniency/severity when evaluating syntactic complexity in L2  (Bogorevich, 2018, 2019; Duym et al., 2018). The proposed study aims to fill this gap.

The following research questions have been formulated:  1) How do teacher  judgments of syntactic complexity converge with ratings by L2 learners? 2) How far do judgments of syntactic complexity by native and non-native raters correlate?  3) To what extent are teacher and student ratings of syntactic complexity correlated with scores obtained by of overall indices of syntactic complexity? A group of 16 expert raters, language teachers with Italian as their L1, were asked to rate on a six-point Likert scale the syntactic complexity of six argumentative texts written by intermediate Dutch L2 learners of Italian (level A2-B1). Teacher judgments, score justifications and suggestions for feedback were confronted with the ratings and feedback of 28 non-expert L2 raters (university students of Italian, level B2-C1), in two different linguistic contexts (Finland, Hungary), and with the judgments of 20 native Italian university students. All holistic human ratings were then compared with the outcomes on three indices of overall complexity (Clauses/T-unit, Dependent Clauses/Clause, Words/T-unit).  The results showed that scores of teachers and students highly agreed, although teacher ratings appeared to be motivated by accuracy rather than by reasons concerning syntactic complexity. Students’ feedback turned out to be more encouraging compared to feedback provided by teachers. Teacher and student ratings corresponded most with Words/T-unit. No impact of (non) nativeness and linguistic background was found.  In the paper the background, methodology and the implications of the study for complexity research are discussed, together with the implications for classroom practice and teacher training.  

Relationships between bilingual exposure at daycare and vocabulary growth in a linguistically diverse group of two- to four-year-olds

Josje Verhagen (University of Amsterdam), Jan Boom (Utrecht University), Suzanne Aalberse (University of Amsterdam), Darlene Keydeniers (Royal Auris Group), Folkert Kuiken (University of Amsterdam), Anne-Mieke Thieme (University of Amsterdam) and Sible Andringa (University of Amsterdam).  

Earlier work has shown that the amount of exposure bilingual children receive in a language is associated with their development in this language. However, previous research has typically focused on the home context, looking at exposure by caregivers, and to a lesser extent, siblings (Unsworth, 2013). In many countries, children attend daycare from a young age onward, and an increasing number of children attend bilingual programs where an additional language is spoken next to the majority language (e.g., English in Spain). Few studies to date have examined the effects of bilingual exposure at daycare on children’s language development, but the available studies indicate that bilingual daycare has positive effects on children’s development of the minority language and does not negatively impact their development of the majority language (Thieme et al., 2021). However, it is currently unknown if these effects remain if home language exposure is taken into analysis: earlier studies collapsed children irrespective of home language background or looked at children from majority language only families. The aim of this study was to investigate how bilingual exposure at daycare relates to children’s dual language development when differences in home language exposure are taken into account. 

Participants were 584 two- to four-year-old children who attended Dutch-only or Dutch-English daycare in the Netherlands. In the Dutch-English daycare centers, English as a foreign language was used between 11% and 50% of the time. Children formed a diverse group, as they were exposed to Dutch, English, other languages, or combinations thereof at home. Dutch and English versions of the Peabody Picture Vocabulary Test were administered in four waves that were approximately nine months apart. Estimates of language exposure to Dutch and English at home and at daycare were derived from a parental questionnaire, and weighted for the actual time children spent at home and daycare in a typical week. 

Latent Growth Modeling analyses showed that the amount of English exposure at daycare was significantly and positively related to growth of English receptive and expressive vocabulary. Amount of Dutch exposure at daycare did not show significant relationships with growth of Dutch receptive and expressive vocabulary. The strengths of the relationships between Dutch and English exposure at daycare and vocabulary growth did not differ between children depending on whether they were exposed to Dutch or English at home. 

These results corroborate the effects of bilingual daycare found in earlier research and indicate that these effects remain if home language exposure is taken into account. Thus, our findings indicate that, at least with the current sample and in the current context, bilingual daycare supports dual language development from a young age onward. 

References

Thieme, A. M. M., Hanekamp, K., Andringa, S., Verhagen, J., & Kuiken, F. (2022). The effects of foreign language programmes in early childhood education and care: a systematic review. Language, Culture and Curriculum, 35, 334-351. 

Unsworth, S. (2013). Current issues in multilingual first language acquisition. Annual Review of Applied Linguistics, 33, 21-50.  

Lexical overlap in foreign language speech segmentation in primary-level students

Katie Von Holzen (English Linguistics, TU Braunschweig), Marie Schnieders (English Linguistics, TU Braunschweig), Sophia Wulfert (English Linguistics, TU Braunschweig) and Holger Hopp (English Linguistics, TU Braunschweig).   

Policy makers base decisions about the ideal starting date for foreign language (FL) teaching on FL learning outcome studies at secondary level which often support a later start (Jaekel et al., 2017). However, their results are influenced by the quality of instruction rather than the learning abilities by the students (Baumert et al., 2020). 

To study students’ initial learning abilities, we focus on a key skill: speech segmentation. Adults initially struggle to extract or segment word forms from continuous speech in an FL but succeed when target words partially overlap in form and meaning with their L1 equivalents (i.e. cognate: English: /kraʊn/; German: /kroːnə/; noncognate: English: /skɪn/; German: /haʊ̯t/; Shoemaker & Rast, 2013) or when target words are adjacent to previously familiarized words (Cunillera et al., 2010). To identify the optimal age for starting FL learning at primary-school level, we study the role of lexical overlap with German for the segmentation of English speech among 1st graders early in the school year and among 2nd graders towards the end of the school year before they have received instruction in English. 

In a word recognition study, English-German cognate and noncognate word pairs (n = 160) were embedded into an otherwise identical English sentence frame (She reduced her crown/skin mursk to poverty). Within each sentence frame cognate and noncognate words were followed by a noncognate pseudoword (i.e. mursk). Students listened to the sentence followed by an isolated probe-word and indicated via button press whether they heard the probe word in the sentence. 

In the talk, we will report results from 48 1st graders and 48 2nd graders to be collected by July. So far, we have analyzed results from 27 first-graders. A general-linear mixed model revealed no difference in target accuracy for cognate and non-cognate words (p = 0.83), nor was there a difference in d’prime between them (p = .97; Figure 1). There was also no difference in target accuracy for cognate and non-cognate following words (i.e. mursk, p = 0.53) or in d’prime between them, (p = .97; Figure 2). These preliminary results provide suggestive evidence that 1st grade students do not yet transfer knowledge from their L1 to segment words in an unknown FL. In our final, fully powered sample, we will also examine individual differences in L1 knowledge on the segmentation of FL speech. Results will show whether age (grade level) or L1 knowledge predict initial FL learning skills. 

References 

Baumert, J., et. al. (2020). The Long-Term Proficiency of Early, Middle, and Late Starters Learning English as a Foreign Language at School: A Narrative Review and Empirical Study. LangLearn, 70(4), 1091–1135. 

Cunillera, T., et. al. (2010). Words as anchors: Known words facilitate statistical learning. ExpPsych, 57(2), 134–141. 

Jaekel, N., et. al. (2017). From Early Starters to Late Finishers? A Longitudinal Study of Early Foreign Language Learning in School. LangLearn, 67(3), 631–664.  

Shoemaker, E., & Rast, R. (2013). Extracting words from the speech stream at first exposure. SecLangRes, 29(2), 165–183.  

French Immersion vs. Core French L2 Accentedness: Proficiency Scores and Native Speaker Ratings

Hilary Walton (University of Toronto).  

This study evaluates whether there are differences in the second language (L2) speech of students enrolled in French immersion and core French programs by comparing learners’ oral proficiency as evaluated by native speaker accentedness ratings and their overall French language ability as evaluated by a validated French proficiency test. French immersion and core French programs are the primary French-as-a-second language (FSL) programs in Canada. While students in both programs acquire French in formal classroom contexts in English-speaking communities, the amount of French instruction and the subjects that are taught in French differ between programs (Canadian Parents for French, 2017). Notably, French immersion speakers have been shown to have a particular non-native French accent, distinct from that of core French speakers (Poljak, 2015); however, the structures that distinguish the L2 speech of these groups remain virtually unexplored. Thus, we seek to distinguish whether there are, indeed, differences in the overall accentedness of these learner groups and if so, to determine whether such differences can be attributed to L2 proficiency. 

This study has two primary research questions (1) Are there differences in the accentedness of the L2 French speech of French immersion and core French speakers? and (2) If so, are such differences due to program- or proficiency-related factors? To assess these questions, this study examined the French oral production of 58 Grade 12 FSL students enrolled in either a French immersion (n = 28) or a core French program (n = 30). Participants first read the French passage “Mes parents m’énervent” aloud. Next, participants completed the University of Toronto Test of French as an overall measure of language ability. Lastly, recordings of the participants’ passage reading underwent an accentedness rating task which was evaluated by six native speaker judges along a 5-point Likert scale from 1 (Definitely non-native; very strong foreign accent) to 5 (Definitely native; no foreign accent).   Preliminary results indicate that native speakers do, indeed, score the accentedness of French immersion and core French speakers differently. Overall, French immersion speakers were given a lower accentedness rating, indicating that native speakers perceived this group as sounding more native-like. This group also had higher levels of overall French proficiency than the core French group. However, when evaluating the recordings of proficiency-matched participants, native speakers seem to associate core French learners with a lower accentedness rating, indicating that they have a more native-like accent than their French immersion counterparts. Further analyses will confirm whether between-group accentedness differences between the French immersion and core French program-level and proficiency-matched speaker groups exist. This study evaluates whether differences in the L2 speech of French classroom learners are program- or proficiency-based. The significance of the results will contribute to the field of L2 speech learning and could highlight the influence of FSL programs on the L2 acquisition of French speech. 

References

Canadian Parents for French. (2017). Category placement/hours of instruction/French language experience.

Poljak, L. (2015). A search for “Immersionese”: Identifying French immersion accents in BC. Masters Thesis. Simon Fraser University, British Columbia.  

Multilingualism negatively predicts prosocial behaviour in young adults, mediated by empathic concern

Yifan Wang (University of Birmingham) and Andrea Krott (University of Birmingham).  

Multilingual speakers demonstrate increased empathy [1] compared to monolinguals (but see [2]). Also, empathy is closely related to prosocial behaviour [3]. The present study tested whether multilingualism, especially frequency of language use and language proficiency, promotes prosocial behaviour and, if so, whether this is mediated by empathy.  126 monolinguals, 126 bilinguals, and 50 participants knowing three to five languages filled in the Language and Social Background Questionnaire [4], the Prosocialness Scale [5] and the Interpersonal Reactivity Index [3] (measuring empathy), plus the Culture Orientation Scale [6] and a Socioeconomic Status (SES) questionnaire to control for confounds. We measured multilingualism as self-reported language usage frequency and proficiency across languages. 

We first established whether multilingualism promoted empathy in our sample. We conducted mixed-effect model analyses for empathy and its four subscales (Perspective taking, Fantasy Score, Empathic Concern, Personal Distress), with the two multilingualism measures as predictors in separate models; and age, SES and cultural orientation as covariates. Both multilingualism measures were significant negative predictors for Empathic Concern (next to collectivism and individualism) and Personal Distress (next to SES).  We then tested whether multilingualism promoted prosocial behaviour and, if so, whether it is mediated by Empathic Concern or Personal Distress. Both multilingualism measures were negatively related to prosocial behaviour, and Empathic Concern (but not Personal Distress) was positively related to prosocial behaviour. The mediation analysis showed only one and only an indirect relationship of multilingualism with prosocial behaviour: Language usage frequency predicted prosocial behaviour indirectly via Empathic Concern.  These results suggest that being multilingual, or more precisely, high language usage frequency, does not lead to higher but lower prosocial behaviour than being monolingual, and that this is due to lower empathic concern. The negative relationship with empathy stands in contrast to previous findings [1] and might be due to multilinguals being more emotional stable than monolinguals and being less impacted by emotional distraction [7]. Furthermore, multilinguals scoring lower in personal distress confirmed that bilinguals are less anxious than monolinguals [8].

References

[1]Dewaele,J.-M.,&Wei,L(2012). Multilingualism, Empathy and Multicompetence. International Journal of Multilingualism, 9(4),352-366. 

[2]Ghoreyshi Rad,F.,Ahmadi, E.,&Gorbani,F.(2020). Comparison of Social Cognition and Executive Functions of Motivation, Inhibitory Control, and Empathy in Bilingual and Monolingual Individuals. International Journal of Psychology, 14(1),59–82.

[3]Davis,M.H.(1983). Measuring individual differences in empathy: Evidence for a multidimensional approach. Journal of Personality and Social Psychology, 44(1),113–126. 

[4]Anderson,J.A.E.,Mak,L.,Keyvani Chahi,A.,&Bialystok,E.(2018). The language and social background questionnaire: Assessing degree of bilingualism in a diverse population. Behavior Research Methods, 50(1),250–263.

[5]Caprara,G.V.,Steca,P.,Zelli,A.,&Capanna,C.(2005). A New Scale for Measuring Adults' Prosocialness. European Journal of Psychological Assessment, 21(2),77–89. 

[6]Triandis,H.C.&Gelfland,M.J.(1998). Converging measurement of horizontal and vertical individualism and collectivism. Journal of Personality and Social Psychology, 74,118-128.

[7]Barker,R.M.,&Bialystok,E.(2019). Processing differences between monolingual and bilingual young adults on an emotion n-back task. Brain and cognition, 134,29–43.

[8]Dewaele,J.-M.,&Wei,L.(2013). Is multilingualism linked to a higher tolerance of ambiguity?. Bilingualism: Language and Cognition, 16(1),231–240. 

Processing cognates in idiomatic expressions: a real ‘nut to crack’

Kristina Weissbecker (University of Kassel).  

Cognates make up a significant amount of a bilingual’s vocabulary, and many psycholinguistic studies have attested, they provide processing advantages as opposed to non-cognates (e.g., Costa et al. 2000, Dijkstra, 2005, Sánchez-Casas and García-Albea, 2005). Cognates are also part of numerous idiomatic expressions. Idioms, in turn, present an intricate case for study concerning their (non-)compositionality and the question if they require their own entry in the mental lexicon (see, e.g., Gibbs 1993, Jackendoff 2007). Also, it is yet an open question whether the literal meanings of the corresponding words in an idiom are accessed during processing. 

This study deals with the processing of cognates within idiomatic expressions. The main research question is: Can bilingual speakers access a cognate’s literal meaning and translation equivalent when it is inside an idiom? In other words, can we also find cognate facilitation in the processing of idioms? To investigate this issue, we are conducting a self-paced reading experiment (SPR) paired with a subsequent lexical decision task (LD). The experimental items consist of English sentences including idiomatic VPs with direct objects as target words. The dependent variable is the cognate status of the target word (cognate vs. non-cognate). The stimuli in the LD are German translations of the targets from the SPR. Fillers contain non-idiomatic phrases in the SPR and pseudo-words in the LD. The measured variables are reaction times (RTs) and accuracy. If speakers can access the literal meaning of the words in an idiom, then faster RTs and a higher accuracy should occur in the LD for experimental items than for fillers. If speakers can access a cognate’s literal meaning in an idiom, a processing advantage should occur in the LD as opposed to items with non-cognates. 

We expect a processing effect for cognate items which would indicate strong connections of cognate equivalents in the lexicon, possibly even a joint storage within one entry. Alternatively, absence of cognate facilitation would imply a separation of cognate translations (similar to non-cognates). Absence of significant differences between experimental items and fillers would mean that accessing the literal meaning of the words within an idiom is not likely. If this is the case, then idiomatic expressions presumably require separate storage in the lexicon. 

References

Costa, A., Caramazza, A., and Sebastian-Galles, N. (2000). The Cognate Facilitation Effect: Implications for Models of Lexical Access. Journal of Experimental Psychology: Leaning, Memory, and Cognition 26.5 (pp. 1283-96). 

Gibbs, R.W. (1993). Why idioms are not dead metaphors. In C. Cacciari & P. Tabossi (Eds.), Idioms. Processing, structure and interpretation (pp. 57-78). Psychology Press.

Dijkstra, T. (2005). Bilingual visual word recognition and lexical access. In J. F. Kroll & A. M. B. de Groot (Eds.), Handbook of bilingualism. Psycholinguistic approaches (pp. 179-201). Oxford UP.

Jackendoff, R. (2007). Language, Consciousness, Culture. MIT Press. 

Sánchez-Casas, R. & García-Albea, J. E. (2005). The representation of cognate and noncognate words in bilingual memory. In: J. F. Kroll & A. M. B. de Groot (Eds.), Handbook of Bilingualism (pp. 226-250). Oxford UP. 

Local and long-distance classifier-noun agreement in L2 Chinese sentence processing

Fei Yuan (University of Cambridge) and Boping Yuan (Shanghai Jiao Tong University/University of Cambridge).  

Background

Second Language (L2) studies of agreement have yielded rich insights into gender and number agreement in L2 grammars. However, relatively few L2 studies have examined classifier-noun (dis)agreement. In the current study, we delve into the processing of classifier-noun (dis)agreement in L2 Chinese among learners from different first language (L1) backgrounds (English: a non-classifier language; Korean: a classifier language). Specifically, we compare the participants’ processing of local and long-distance agreement to explore the distance effect. We also examine the cue-based retrieval model in explaining our findings (Lewis & Vasishth, 2013; Cunnings, 2017). Our goals are to examine L1/L2 differences in processing classifier-noun (dis)agreement and to assess the impact of L1 transfer on processing. 

Methodology

A self-paced reading (SPR) task and an acceptability judgment task (AJT) were administered in this study. Chinese native speakers (NSs) (n=28) were recruited as a control group, with L1-English learners of L2 Chinese (EN group) (n=54) and proficiency-matched L1-Korean learners of L2 Chinese (KR group) (n=51) serving as two experimental groups. The same items were used in both tasks (see example stimuli on the next page). Each type was represented by six tokens, with 77% of the test items being filler items. The data were analysed using R software, and linear mixed-effects models were constructed to compare the types. 

Results

In the AJT, all three groups rated Types 1 and 2 as acceptable (z-scores>0), but judged Type 3, the disagreement condition, as unacceptable (z-scores<0). This suggests that both native speakers and L2 learners could detect classifier-noun disagreement in an off-line setting. In the SPR (see Figures 1-3), the NS group demonstrated sensitivity to the local disagreement, reflected in significantly and marginally shorter RTs of Types 1 and 3 in Region 6 (Types 1 vs 2: p=0.05; Types 2 vs 3: p=0.089). The RT of Region 9 in Type 3 was significantly longer than those in Types 1 and 2 (Types 1 vs 3: p=0.005; Types 2 vs 3: p=0.002). However, the EN group did not exhibit any effect in their SPR results. In the KR group, the RT of Region 6 in Type 2 was marginally shorter than those in Types 1 and 3 (Types 1 vs 2: p=0.087; Types 2 vs 3: p=0.091), similar to the sensitivity observed in the NS group to local disagreement. However, no effect was observed for the remainder of the regions in the KR group, indicating that the KR group was not sensitive to long-distance disagreement during real-time processing. 

Discussion

The L1/L2 differences were found in our study, indexed by the NS group’s sensitivity and the L2 groups’ insensitivity to the long-distance agreement. This phenomenon is likely a result of L2 learners’ inability to retrieve cues from working memory in non-local disagreement. L1 transfer was also observed, as L1-Korean learners, but not L1-English learners, behaved native-like and were sensitive to local disagreement. This is attributed to a facilitative influence of their L1 Korean, which is a classifier language as well. 

Relativized Minimality in L2 revisited: (non-)effects of L1 and tense on processing of object relative clauses

Vera Yunxiao Xia (University of Alberta), Natália Brambatti Guzzo (St. Mary's University) and Lydia White (McGill University).   

Object relative clauses (ORCs) are known to cause greater problems than subject relatives (SRCs) in acquisition and processing (Adani et al. 2010). Friedmann et al. (2009) advance a Relativized Minimality (RM) account: the subject of the ORC intervenes between the head noun and its source position, potentially making interpretation hard. The intervention effect is reduced if the relative head and the intervenor mismatch in features such as number. See (1). 

(1)  I know the musicians (P) who the waiter (S) likes _. 

For L2 acquisition, Authors (2022) report that, contrary to RM predictions, L2ers were faster in the case of matched items (SS vs. SP) in English. Authors speculated that this might be due to the L1 (Mandarin) not having number marking. A second issue is that, in the previous literature, there are always two cues to number: plural inflection on the noun and verbal agreement. In Authors, all verbs were past tense, so the only mismatch involved nominal inflection; compare (1) and (2). 

(2)  I know the musicians who the waiter liked _. 

In this paper, we explore effects of L1 and tense, comparing speakers of Mandarin with Spanish, a language with number marking, and including stimuli with present and past tense verbs. Research questions are: 

i. Do Spanish speakers (unlike Mandarin) respond faster to mismatched items than matched?       ii. Do L2ers respond faster to present tense items than past? 

An experiment was conducted with intermediate level learners of English (L1s Mandarin, n=18, Spanish, n=22), as well as native speaker controls (n=20), using a self-paced reading task. Sentences were presented in segments, followed by a comprehension question. The task included 32 sentences manipulating RC type (subject/object), number on relative head and intervenor (±singular), and tense (±past), plus distractors. 

Results show high accuracy overall for the comprehension questions: Spanish speakers were more accurate on SRCs than ORCs (p = 0.01) and the Mandarin speakers trended in the same direction, consistent with RM. Looking at overall RTs, none of the groups shows a difference between matched and mismatched items, or an effect for tense. The same is true for RTs at the critical region where match or mismatch becomes evident. However, number did make a difference: Mandarin speakers performed significantly faster on PP versus SS clauses (p = 0.04), suggesting that they are in fact sensitive to number. These results are consistent with Contemori and Marinis (2014), who also found no mismatch effect for child and adult English-speakers in a self-paced listening task, and an advantage for plural over singular. Contemori and Marinis propose that mismatch effects are late effects which show up only in offline tasks. Our results are consistent with this proposal. 

References

Adani, F. et al. 2010. Grammatical feature dissimilarities make relative clauses easier. Lingua 120: 2148-2166. 

Contemori, C., & Marinis, T. 2014. The impact of number mismatch and passives on real-time processing of relative clauses. JCL 41: 658-689. 

Friedmann, M. et al. 2009. Relativized relatives: types of intervention in the acquisition of A-bar dependencies. Lingua 119: 67-88. 

The value of sign and print: Language proficiency predicts deaf signers’ occupational prestige and income

Yeqiu Zheng (Erasmus University Rotterdam), Hao Lin (Shanghai international Studies University) and Yan Gu (University of Essex & University College London).  

Language impacts immigrants' earnings and financial wealth concurrently and longitudinally (Zheng, Gu, Backus, & van Soest, 2022). However, no research has investigated deaf signers’ language proficiency and their occupational prestige and income, though deaf signers are disadvantageous in labour market and their language use is of policy interest. 

109 Shanghai Chinese Sign Language (CSL)-Mandarin print deaf signers were paid to fill in an online Mandarin questionnaire about their language history (Li et al., 2020), employment and background information. As Zheng et al. (2022) found that results of language and earnings were generally robust when using a self-assessed language proficiency compared to a language placement test, we followed most past research to use a self-assessment “Please assess your Mandarin/CSL proficiency on a scale of 1-7 (1=completely do not speak, 7=native proficiency)”. 

Deaf signers in our sample had a high employment rate (96.3%). To analyse whether CSL sign and Mandarin print proficiency related to deaf signers’ occupational prestige and income, we coded two binary dependent variables “prestigious_job”  based on categories in Li (2005), and “high_income” (=1 if > 6000 yuan, average net income,  Statistics Shanghai, 2020; otherwise=0). As deaf signers who had a higher CSL proficiency level also had a higher Mandarin print proficiency,  r=0.53,  p<.001, we used Mandarin print and CSL proficiency separately as a predictor for dependent variables rather than putting both in the same model to avoid multicollinearity. Given that age of acquisition (AoA) is a strong predictor for sign proficiency (Mayberry, 2011), for the model of Mandarin print, we controlled for the AoA of CSL as a proxy of CSL language proficiency (same for Mandarin in the model of CSL). We also controlled for demographic factors, type of schools attended, severity of hearing loss, etc. We used the GLM generalizes linear regression in R starting with the full model and then the backwards step function based on AIC values to choose the best fit model. Results showed that a higher education (β=1.22, p<.001) and CSL proficiency level (β=.44, p=.024) predicted a more prestigious job. In addition, CSL proficiency (β=.75, p=.034) and education (β=1.10, p=.045) significantly predicted deaf signers’ probability of having a high income, whereas Mandarin print proficiency was marginally significant (β=.76, p =.049). 

Furthermore, as a sensitivity analysis, we coded the proficiency level (1-7) of CSL or Mandarin print as ‘high’ if  the self-assessed score was 6 or 7, otherwise ‘low’ for the rest. This time we were able to put both CSL and Mandarin proficiency (high/low) in the same model together with other factors. The final model showed that only CSL proficiency level (β=1.11, p=.022) and education (β=1.19, p<.001) predicted whether having a more prestigious job, and a higher CSL proficiency (β=1.88, p=.023) and education level (β=1.06, p=.043) predicted a higher income. 

Our study for the first time showed the relationship between deaf bilinguals’ language proficiency and their occupational prestige and income, with strong implications for policymakers, such as the emphasis on the importance of learning sign language in deaf education. 

L’influence translinguistique dans les productions narratives des francophones en italien L2 et l’utilisation des particules de portée additives “anche” et “ancora”

Maria Rosaria D'Angelo

Notre étude porte sur des productions de récits d’images réalisées par des apprenants francophones de l’italien ayant trois niveaux interlinguistiques différents.
Pour le recueil de données, nous avons utilisé une tâche narrative
 permettant d’analyser le marquage des relations additives (Benazzo & col. 2004). Nous disposons de quinze récits par niveau d’apprenant (débutant, intermédiaire et avancé)
et de deux groups de contrôle italien et francais.

Nous avons analysé les différences et/ou les ressemblances interlinguistiques entre l’italien et le français qui sont variables selon le niveau d’acquisition, sur la mise en œuvre en italien L2 d’adverbes dits « particules de portée» tels que « anche » et « ancora » dont les proprietés, qui jouent un rôle crucial du point de vue communicatif, concernent plusieurs niveaux linguistiques : sémantique, syntaxique et discursif.

« Anche » et « ancora », considérés comme des particules additives créant des relations d’addition, sont abordés habituellement dans le cadre de la sémantique formelle (König 1991). Cependant, des études empiriques sur l’acquisition des phénomènes de portée au sein de l’approche fonctionnelle ont mis en évidence la nécessité de prendre en compte des unités contextuelles plus large que la phrase. Elles analysent ces phénomènes de portée dans des variétés initiées par Dimroth et Klein (1996) et ont contribué à éclairer le fonctionnement de ce type d’items en relation avec le discours dans lequel ils sont produits.

Nos résultats montrent que le français LM influence les productions des apprenants en italien surtout au niveau de la structure syntaxique même si la signification de base de la particule n’est pas affectée.

Nous avons finalement constaté des stratégies interlanguistiques de transfert aussi bien que des stratégies propres au système de l’interlangue dans les domaines de la syntaxe des particules, ce qui amène à des placements anomaux.

On remarque notamment des phenomenes de transfert dans les contexts de portée à distance et a gauche, donc dans des contexts qui crèent des problèmes de distribution, qui produisent évidemment des situations de détresse expressive. Par consequent le transfert semble représenter un support en cas de difficulté, car il exprime la référence instinctive à des solutions connues pour résoudre cette difficulté même au niveau non débutants. 

Bibliographie

Benazzo, S.,Dimroth, C.,Perdue, C.,& Watorek, M. (2004). Le rôle des particules additives dans la construction de la cohésion discursive en langue maternelle et en langue étrangère. Langages 155, pp. 76-105

Dimroth, C., & Klein, W. (1996). Les particules focales dans les variétés d'apprentissage : un cadre d'analyse et quelques exemples. Zeitschrift für Literaturwissenschaft und Linguistik, 104, 73-114.

Konig, E. (1991). The meaning of focus particles: a comparative perspective. London: Routledge. 

Return to the main EuroSLA home page