1,556
Views
16
CrossRef citations to date
0
Altmetric
Articles

A categorisation of the terminological sources of student difficulties when learning chemistry

References

  • Adger, C. T., Snow, C. E., & Christian, D. (2002). What teachers need to know about language. McHenry: CAL.
  • Adúriz-Bravo, A. (2013). A ‘semantic’ view of scientific models for science education. Science & Education, 22(7), 1593–1611.
  • Airey, J. (2012). “I don’t teach language”: The linguistic attitudes of physics lecturers in Sweden. AILA Review, 25, 64–79.
  • Aledo, J. C. (2007). Coupled reactions versus connected reactions: Coupling concepts with terms. Biochemistry and Molecular Biology Education, 35, 85–88.
  • Ali, M., & Ismail, Z. (2006). Comprehension level of non-technical terms in science: Are we ready for science in English? Journal Pendidik Dan Pendidikan, 21, 73–83.
  • Amaral, E. M. R., Mortimer, E. F., & Scott, P. (2014). A conceptual profile of entropy and spontaneity: Characterising modes of thinking and ways of speaking in the classroom. In E. F. Mortimer & C. N. El-Hani (Eds.), Conceptual profiles: A theory of teaching and learning scientific concepts (pp. 201–234). New York: Springer.
  • Amaral, E. M. R., Silva, J. R., & Sabino, J. D. (2018). Analysing processes of conceptualization for students in lessons on substance from the emergence of conceptual profile zones. Chemistry Education Research and Practice, 19, 1010–1028.
  • Anselme, J. P. (1997). Understanding oxidation-reduction in organic chemistry. Journal of Chemical Education, 74(1), 69–72.
  • Archila, P. A. (2014). Are science teachers prepared to promote argumentation? A case study with pre-service teachers in Bogotá city. Asia-Pacific Forum on Science Learning and Teaching, 15(1), 1–21.
  • Arkoudis, S. (2003). Teaching English as a second language in science classes: Incommensurate epistemologies? Language and Education, 17(3), 161–173.
  • Arons, A. (1973). Toward a wider public understanding of science. American Journal of Physics, 41(6), 769–782.
  • Astington, J. W., & Olson, D. R. (1990). Metacognitive and metalinguistic language: Learning to talk about thought. Applied Psychology: An International Review, 39, 77–87.
  • August, D., Branum-Martin, L., Cardenas-Hagan, E., & Francis, D. J. (2009). The impact of an instructional intervention on the science and language learning of middle grade English language learners. Journal of Research on Educational Effectiveness, 2(4), 345–376.
  • Bächtold, M. (2018). How should energy be defined throughout schooling? Research in Science Education, 48(2), 345–367.
  • Banks, D. (2008). The development of scientific writing: Linguistic features and historical context. London: Equinox.
  • Barak, J., Gorodetsky, M., & Chipman, D. (1997). Understanding of energy in biology and vitalistic conceptions. International Journal of Science Education, 19(1), 21–30.
  • Bauman, R. P. (1992a). Physics that textbook writers usually get wrong. I work. The Physics Teacher, 30, 264–269.
  • Bauman, R. P. (1992b). Physics that textbook writers usually get wrong. II heat and energy. The Physics Teacher, 30, 353–356.
  • Bayir, E. (2014). Developing and playing chemistry games about elements, compounds and the periodic table: Elemental periodica, compoundica and groupica. Journal of Chemical Education, 91(4), 531–535.
  • Bearne, E. (1999). Use of language across secondary curriculum. London: Routledge.
  • Beck, I. L., McKeown, M. G., & Kucan, L. (2013). Bringing words to life: Robust vocabulary instruction. London: Guilford Press.
  • Beek, K. V., & Louters, L. (1991). Chemical language skills. Investigating the deficit. Journal of Chemical Education, 68(5), 389–392.
  • Beezer, G. (1940). Latin and Greek roots in chemical terminology. Journal of Chemical Education, 17(2), 63–66.
  • Bennet, J. M., & Sözbilir, M. (2007). A study of Turkish chemistry students’ understanding of entropy. Journal of Chemical Education, 84(7), 1204–1208.
  • Bensaude-Vincent, B. (1999). A language to order the chaos. Bulletin for the History of Chemistry, 23, 1–10.
  • Bensaude-Vincent, B. (2002). Language in chemistry. In M. J. Nye (Ed.), Cambridge history of science (Vol. 5, pp. 174–190). Cambridge: Cambridge University Press.
  • Ben-Zvi, R., Bat-Sheva, E., & Silberstein, J. (1986). Is an atom of copper malleable? Journal of Chemical Education, 63(1), 64–66.
  • Bergqvist, A., Drechsler, M., De Jong, O., & Rundgren, S.-N. C. (2013). Representations of chemical bonding models in school textbooks – Help or hindrance for understanding? Chemistry Education Research and Practice, 14, 589–606.
  • Bergqvist, W., & Heikkinen, H. (1990). Student ideas regarding chemical equilibrium. Journal of Chemical Education, 67(12), 1000–1003.
  • Best, R., Rowe, M., Ozuru, Y., & McNamara, D. (2005). Deep-level comprehension of science texts: The role of the reader and the text. Topics in Language Disorders, 25, 65–83.
  • Biber, D., Conrad, S., & Cortes, V. (2004). If you look at … : Lexical bundles in university teaching and textbooks. Applied Linguistics, 25(3), 371–405.
  • Bird, E., & Welford, G. (1995). The effect of language on performance of second-language students in science examinations. International Journal of Science Education, 17(3), 389–397.
  • Bleicher, R. E., Tobin, K. G., & McRobbie, C. J. (2003). Opportunities to talk science in a high school chemistry classroom. Research in Science Education, 33, 319–339.
  • Blown, E. J., & Bryce, T. G. K. (2017). Switching between everyday and scientific Language. Research in Science Education, 47, 621–653.
  • Bogaards, P., & Laufer, B. (2004). Vocabulary in a second language, selection, acquisition, and testing. Amsterdam: Benjamins Publishing Company.
  • Bradbury, L. U. (2014). Linking science and language arts: A review of the literature research which compares integrated versus non-integrated approaches. Journal of Science Teacher Education, 25, 465–488.
  • Brookes, D. T., & Etkina, E. (2006). Do our words really matter? Case studies from quantum mechanics. In 2005 physics education research conference, AIP conference proceedings (AIP), Salt Lake City.
  • Brookes, D. T., & Etkina, E. (2007). Using conceptual metaphor and functional grammar to explore how language used in physics affects student learning. Physical Review Physics Education Research, 3, 010105.
  • Brookes, D. T., & Etkina, E. (2015). The importance of language in students’ reasoning about heat in thermodynamic processes. International Journal of Science Education, 37(5–6), 759–779.
  • Brown, B. A. (2006). “It isn’t no slang that can be said about this stuff”: Language, identity, and appropriating science discourse. Journal of Research in Science Teaching, 43(1), 96–126.
  • Brown, B. A., & Ryoo, K. (2008). Teaching science as a language: A ‘content-first’ approach to science teaching. Journal of Research in Science Teaching, 45(5), 529–553.
  • Bucat, R. (2004). Pedagogical content knowledge as a way forward: Applied research in chemistry education. Chemistry Education Research and Practice, 5(3), 215–228.
  • Bulman, L. (1985). Teaching language and study skills in secondary science. London: Heinemann.
  • Bunch, G. C. (2013). Pedagogical Language knowledge: Preparing mainstream teachers for English learners in the new standards era. Review of Educational Research, 37, 298–341.
  • Byrne, M., Johnstone, A. H., & Pope, A. (1994). Reasoning in science: A language problem? School Science Review., 75(272), 103–107.
  • Çalık, M. (2005). A cross-age study of different perspectives in solution chemistry from junior to senior high school. International Journal of Science and Mathematics Education, 3(4), 671–696.
  • Canac, S., & Kermen, I. (2016). Exploring the mastery of French students in using basic notions of the language of chemistry. Chemistry Education Research and Practice, 17, 452–473.
  • Carnine, L., & Carnine, D. (2004). The interaction of reading skills and science content knowledge when teaching struggling secondary students. Reading & Writing Quarterly, 20(2), 203–218.
  • Carr, M. (1984). Model confusion in chemistry. Research in Science Education, 14, 97–103.
  • Carrier, S. J. (2013). Elementary preservice teachers’ science vocabulary: Knowledge and application. Journal of Science Teacher Education, 24, 405–425.
  • Carson, E. M., & Watson, J. R. (1999). Undergraduate students’ understanding of enthalpy change. University Chemistry Education, 3(2), 46–51.
  • Carson, E. M., & Watson, J. R. (2002). Undergraduate students’ understanding of entropy and Gibbs energy. University Chemistry Education, 6(1), 4–12.
  • Cassels, J. R. T., & Johnstone, A. H. (1980). Understanding of non-technical words in science. London: The Royal Society of Chemistry.
  • Cassels, J. R. T., & Johnstone, A. H. (1983). The meaning of words and the teaching of chemistry. Education in Chemistry, 20, 10–11.
  • Cassels, J. R. T., & Johnstone, A. H. (1984). The effect of language on student performance on multiple choice test in chemistry. Journal of Chemical Education, 61, 613–615.
  • Cassels, J. R. T., & Johnstone, A. H. (1985). Words that matter in science. London: The Royal Society of Chemistry.
  • Castillo, J., Ogaz, R., Merino, C., & Quiroz, W. (2016). An ontological and epistemological analysis of the presentation of the first law of thermodynamics in school and university textbooks. Chemistry Education Research and Practice, 17, 1041–1053.
  • Cervellati, R., & Perugini, D. (1981). The understanding of the atomic orbital concept by Italian high school students. Journal of Chemical Education, 58(7), 568–569.
  • Cervetti, G. N., Pearson, P. D., Greenleaf, C., & Moje, E. (2013). Science? Literacy? Synergy! In W. Banko, M. L. Grant, M. E. Jabot, A. J. McCormack, & T. O’Brien (Eds.), Science literacy and our nation’s future (pp. 99–124). Washington, DC: NSTA & STANYS.
  • Cervetti, G. N., Barber, J., Dorph, R., Pearson, P. D., & Goldschmidt, P. G. (2012). The impact of an integrated approach to science and literacy in elementary school classrooms. Journal of Research in Science Teaching, 49(5), 631–658.
  • Chamizo, J. A. (2013). A new definition of models and modelling in chemistry’s teaching. Science & Education, 22(7), 1613–1632.
  • Chandrasegaran, A. L., Treagust, D. F., Waldrip, B. G., & Chandrasegaran, A. (2009). Students’ dilemmas in reaction stoichiometry problem solving: Deducing the limiting reagent in chemical reactions. Chemistry Education Research and Practice, 10(1), 14–23.
  • Chang, H. (2012). Acidity: The persistence of everyday in the scientific. Philosophy of Science, 79, 690–700.
  • Chavkin, L. (2002). Readability and reading ease revisited: Stare-adopted science textbooks. The Clearing House, 70(3), 151–155.
  • Chi, M. T. H., & Roscoe, R. D. (2002). The processes and challenges of conceptual change. In M. Limon & L. Mason (Eds.), Reconsidering conceptual change: Issues in theory and practice (pp. 3–27). The Netherlands: Kluwer Academic Publishers.
  • Chi, M. T. H. (2005). Common-sense conceptions of emergent processes: Why some misconceptions are robust. The Journal of the Learning Sciences, 14(2), 161–199.
  • Chi, M. T. H., Roscoe, R. D., Slotta, J. D., Roy, M., & Chase, C. (2012). Misconceived causal explanations for emergent processes. Cognitive Science, 36, 1–61.
  • Chi, M. T. H., Slotta, J. D., & de Leeuw, N. (1994). From things to processes: A theory of conceptual change for learning science concepts. Learning and Instruction, 4, 27–43.
  • Childs, P. E., Markic, S., & Ryan, M. C. (2015). The role of language in the teaching and learning chemistry. In J. García-Martínez & E. Serrano-Torregrosa (Eds.), Chemistry education: Best practices, opportunities and trends (pp. 421–445). Weinheim: Wiley.
  • Childs, P. E., & O’Farrell, F. J. (2003). Learning science through English: An investigation of the vocabulary skills of native and non-native English speakers in international schools. Chemistry Education Research and Practice, 4(3), 233–247.
  • Chung, T. M., & Nation, P. (2003). Technical vocabulary in specialised texts. Reading in a Foreign Language., 15(2), 103–116.
  • Chung, T. M., & Nation, P. (2004). Identifying technical vocabulary. System, 32, 251–263.
  • Cink, R. B., & Song, Y. (2016). Appropriating scientific vocabulary in chemistry laboratories: A multiple case study of four community college students with diverse ethno-linguistic backgrounds. Chemistry Education Research and Practice, 17, 604–617.
  • Cintas, P. (2007). Tracing the origins and evolution of chirality and handedness in chemical Language. Angewandte Chemie International Edition, 46, 4016–4024.
  • Clerk, D., & Rutherford, M. (2000). Language as a confounding variable in the diagnosis of misconceptions. International Journal of Science Education, 22(7), 703–717.
  • Cooper, A. K., & Oliver-Hoyo, M. T. (2016). Argument construction in understanding noncovalent interactions: A comparison of two argumentation frameworks. Chemistry Education Research and Practice, 17(4), 1006–1018.
  • Cooper, M. M., Corley, L. M., & Underwood, S. M. (2013). An investigation of college chemistry students’ understanding of structure-property relationships. Journal of Research in Science Teaching, 50(6), 699–721.
  • Cooper, M. M., & Klymkowsky, M. W. (2013). The trouble with chemical energy: Why understanding bond energies requires an interdisciplinary systems approach. CBE-Life Science Education, 12, 306–312.
  • Cortes, V. (2004). Lexical bundles in published and student disciplinary writing: Examples from history and biology. English for Specific Purposes, 23, 397–423.
  • Coştu, B., & Ayas, A. (2005). Evaporation in different liquids: Secondary students’ conceptions. Research in Science and Technological Education, 23, 73–95.
  • Costu, B., Ayas, A., & Niaz, M. (2010). Promoting conceptual change in students' understanding of evapotation. Chemistry Education Research and Practice, 11(3), 5–16.
  • Coxhead, A. (2000). A new academic word list. TESOL Quarterly, 34(2), 213–238.
  • Coxhead, A., & Hirsh, D. (2007). A pilot science-specific word list. Revue Française de Linguistique Appliquée, 12, 65–78.
  • Coyle, D., Hood, P., & Marsh, D. (2010). CLIL. Content and Language Integrated Learning. Cambridge: Cambridge University Press.
  • Crosland, M. P. (2004). Historical studies in the language of chemistry. New York: Dover Phoenix Editions.
  • Crosson, A., & Lesaux, N. K. (2013). Connectives. Fitting another piece of the vocabulary instruction puzzle. The Reading Teacher, 67(3), 193–200.
  • Dalcq, A. E. (1987). Questions mal comprises ou mal possés? Un test de compétence linguistique. Langue Française, 75, 36–50.
  • Dalcq, A. E. (1996). Quelques balises pour une pédagogie de la rédaction en français scientifique. In ACFAS (Ed.), Le français et les langues scientifiques de démain (pp. 247–285). Montreal: Université de Québec.
  • Dalcq, A. E., Van Raemdonck, D., & Wilmet, B. (1989). Le français et les sciences. Paris: Duculot.
  • Davies, A. J. (1991). A model approach to teaching redox. Education in Chemistry, 28(5), 135–137.
  • Dávila, K., & Talanquer, V. (2010). Classifying end-of-chapter questions and problems for selected general chemistry textbooks used in the United States. Journal of Chemical Education, 87(1), 97–101.
  • Dawes, L. (2004). Talk and learning in classroom science. International Journal of Science Education, 26(6), 677–695.
  • de Berg, K. C. (2008). The concepts of heat and temperature: The problem of determining the content for the construction of an historical case study which is sensitive to nature of science issues and teaching-learning issues. Science & Education, 17, 75–114.
  • De Jong, O., & Treagust, D. (2002). The teaching and learning of electrochemistry. In J. K. Gilbert, O. De Jong, R. Justi, D. F. Treagust, & J. H. van Driel (Eds.), Chemical education: towards research-based practice (pp. 317–337). Dordrecht: Kluwer Academic Publishers.
  • De Jong, O., & Taber, K. (2007). Teaching and learning the many faces of chemistry. In S. Abell & N. Lederman (Eds.), Handbook of research on science education (pp. 631–652). Mahwah, NJ: Lawrence Erlbaum Publishers.
  • de Vos, W., & Pilot, A. (2001). Acids and bases in layers: The stratal structure of an ancient topic. Journal of Chemical Education, 78(4), 494–499.
  • Dickinson, V. L., & Young, T. A. (1998). Elementary science and language arts: Should we blur the boundaries? School Science and Mathematics, 98(6), 334–339.
  • Dimopoulos, K., Koulaidis, V., & Sklaveniti, S. (2005). Towards a framework of socio-linguistic analysis of science textbooks: The Greek case. Research in Science Education, 35(2), 173–195.
  • Doige, C. A., & Day, T. (2012). A typology of undergraduate textbook definitions of ‘heat’ across science disciplines. International Journal of Science Education, 34(5), 677–700.
  • Doménech, J. L., Gil-Pérez, D., Gras-Martí, A., Guisasola, J., Martínez-Torregrosa, J., Salinas, J., … Vilches, A. (2007). Teaching of energy issues: A debate proposal for a global reorientation. Science & Education, 16, 43–64.
  • Drechsler, M., & Schmidt, H.-J. (2005). Upper secondary school students’ understanding of models used in chemistry to define acids and bases. Science Education International, 16(1), 39–53.
  • Dreyfus, B. W., Gouvea, J., Geller, B. D., Sawtelle, V., Turpen, C., & Redish, E. F. (2014). Chemical energy in an introductory physics course for life science students. American Journal of Physics, 82, 403–411.
  • Driscoll, D. (1978). Comments on ionization. Journal of Chemical Education, 55, 465.
  • Duggleby, S. J., Tang, W., & Kwo-Newhouse, A. (2015). Does the use of connective words in written assessment predict high school students’ reading and writing achievement? Reading Psychology, 37, 511–532.
  • Ebenezer, J. V., & Erickson, G. L. (1996). Chemistry students’ conceptions of solubility: A phenomenography. Science Education, 80(2), 181–201.
  • El Masri, Y. H., Baird, J.-A., & Graesser, A. (2016). Language effects in international testing: The case of PISA 2006 science items. Assessment in Education: Principles, Policy & Practice, 23(4), 427–455.
  • Ellis, R. A. (2004). University student approaches to learning science through writing. International Journal of Science Education., 26(15), 1835–1853.
  • Evagorou, M., & Osborne, J. (2010). The role of language in the learning and teaching science. In J. Osborne & J. Dillon (Eds.), Good practice in science teaching. What practice has to say (pp. 135–157). Maidenhead: Open University Press.
  • Evrard, N., Huynen, A., & Van Der Borght, C. (1998). Communication of scientific knowledge in class - from verbalization to the concept of chemical equilibrium. International Journal of Science Education, 20(8), 883–900.
  • Fang, Z. (2005). Scientific literacy: A systemic functional linguistics perspective. Science Education, 89, 335–347.
  • Fang, Z. (2006). The language demands of science reading in middle school. International Journal of Science Education, 28(5), 491–520.
  • Fang, Z. (2014). Preparing content area teachers for disciplinary literacy instruction. Journal of Adolescent & Adult Literacy, 57(6), 444–448.
  • Fang, Z., & Schleppegrell, M. J. (2008). Reading in secondary content areas: A language-based pedagogy. Ann Arbor: University of Michigan Press.
  • Fang, Z., Schleppegrell, M. J., & Cox, B. E. (2006). Understanding the Language demands of schooling: Nouns in academic registers. Journal of Literacy Research, 38(3), 247–273.
  • Farrell, M., & Ventura, F. (1998). Words and understanding in Physics. Language and Education, 12(4), 243–253.
  • Fatonah, F. (2014). Students’ understanding of the realization of nominalizations in scientific text. Journal of Applied Linguistics, 4(3), 87–98.
  • Fazio, X., & Gallagher, T. L. (2019). Science and language integration in elementary classrooms: Instructional enactments and student learning outcomes. Research in Science Education, 49(4), 959–976.
  • Feez, S., & Quinn, F. (2017). Teaching the distinctive language of science: An integrated and scaffolded approach for pre-service teachers. Teaching and Teacher Education, 65, 192–204.
  • Fensham, P. J. (2004). Defining an identity. The evolution of science education as a field of research. Dordrecht: Kluwer.
  • Fensham, P. J., & Cumming, J. J. (2013). “Which child left behind”: Historical issues regarding equity in science assessment. Education Science, 3, 326–343.
  • Flores, F., Ulloa, N., & Covarrubias, H. (2015). The concept of entropy, from its origins to teachers. Revista Mexicana de Física, 61, 69–80.
  • Ford, A., & Peat, F. (1988). The role of language in science. Foundations of Physics, 18, 1233–1241.
  • Fradd, S. H., & Lee, O. (1999). Teachers’ roles in promoting science inquiry with students from diverse language backgrounds. Educational Researcher, 28, 14–42.
  • Frank, J. O. (1929). Valence as defined in high-school texts. Journal of Chemical Education, 6(4), 718.
  • Furió, C., Calatayud, M. L., & Bárcenas, S. L. (2007). Surveying students’ conceptual and procedural knowledge of acid‐base behaviour of substances. Journal of Chemical Education, 84, 1717–1724.
  • Furió, C., Calatayud, M. L., Guisasola, J., & Furió-Gómez, C. (2005). How are the concepts and theories of acid-base reactions presented? Chemistry in textbooks and as presented by teachers. International Journal of Science Education, 27(11), 1337–1358.
  • Galguera, T. (2011). Participant structures as professional learning tasks and the development of pedagogical language knowledge among preservice teachers. Teacher Education Quarterly, 38, 85–106.
  • Galili, I., & Lehavi, Y. (2006). Definitions of physical concepts: A study of physics teachers’ knowledge and views. International Journal of Science Education, 28(5), 521–541.
  • Gallagher, T., Fazio, X., & Ciampa, K. (2017). A comparison of readability in science-based texts: Implications for elementary teachers. Canadian Journal of Education, 40(1), 1–29.
  • Gardner, D., & Davies, M. (2014). A new academic vocabulary list. Applied Linguistics, 35(3), 305–327.
  • Gardner, P. L. (1971). Project SWNG—Scientific words: New Guinea. Melbourne: Faculty of Education, Monash University.
  • Gardner, P. L. (1972). ‘Words in science’: An investigation of non-technical vocabulary difficulties amongst form I, II, III and IV science students in Victoria. Melbourne: Australian Science Education Project.
  • Gardner, P. L. (1974). Language difficulties of science students. The Australian Science Teachers’ Journal, 20(1), 63–76.
  • Gardner, P. L. (1975). Logical connectives in science: A preliminary report. Research in Science Education, 5(1), 161–175.
  • Gardner, P. L. (1977). Logical connectives in science: A summary of the findings. Research in Science Education, 7(1), 9–24.
  • Gardner, P. L. (1980a). Difficulties with non-technical scientific vocabulary amongst secondary school students in the Philippines. The Australian Science Teachers’ Journal, 26(2), 82–90.
  • Gardner, P. L. (1980b). Identification of specific difficulties with logical connectives in science among secondary school students. Journal of Research in Science Teaching, 17(3), 223–229.
  • Gardner, P. L., Schafe, L., Thein, U. M., & Watterson, R. (1976). Logical connectives in science: Some preliminary findings. Research in Science Education, 6(1), 97–108.
  • Garnett, P. J., Garnett, P. J., & Hackling, M. A. W. (1995). Students’ alternative conceptions in chemistry: A review of research an implications for teaching and learning. Studies in Science Education, 25, 69–95.
  • Garraway, G. B. (1994). Language, culture and attitude in mathematics and science learning: A review of the literature. Journal of Research and Development in Education, 27(2), 102–111.
  • Gauchon, L., & Méheut, M. (2007). Learning about stoichiometry: From students’ preconceptions to the concepts of limiting reactant. Chemistry Education Research and Practice, 8(4), 362–365.
  • Gericke, N. M., & Hagberg, M. (2010). Conceptual incoherence as a result of the use of multiple historical models in school textbooks. Research in Science Education, 40(4), 605–623.
  • Gibbons, P. (2014). Scaffolding language, scaffolding learning: Teaching English language learners in the mainstream classroom. Portsmouth: Heinemann.
  • Gilbert, J. K. (2006). On the nature of ‘context’ in chemical education. International Journal of Science Education, 28(9), 957–976.
  • Giunta, C. J. (2015). The mole and amount of substance in chemistry and education: beyond official definitions. Journal Of Chemical Education, 92(10), 1393–1397. doi:10.1021/ed5007376
  • Giunta, C. J. (2016). What’s in a name? amount of substance, chemical amount, and stoichiometric amount. Journal Of Chemical Education, 93(4), 583–586. doi:10.1021/acs.jchemed.5b00690
  • Glen, N. J., & Dotger, S. (2009). Elementary teachers’ use of language to label and interpret science concepts. Journal of Elementary Science Education, 21(4), 71–83.
  • Glen, N. J., & Dotger, S. (2013). Writing like a scientist: Exploring elementary teachers’ understandings and practices of writing in science. Journal of Science Teacher Education, 24, 957–976.
  • Goedhart, M. J., & Kaper, W. (2002). From chemical energetics to chemical thermodynamics. In J. K. Gilbert, O. De Jong, R. Justi, D. F. Treagust, & J. H. Van Driel (Eds.), Chemical education: Towards research-based practice (pp. 339–363). Dordrecht: Kluwer Academic Publishers.
  • Gómez-Crespo, M. A., & Pozo, J. I. (2004). Relationships between everyday knowledge and scientific knowledge: Understanding how matter changes. International Journal of Science Education, 26(11), 1325–1343.
  • Green, C. (2019). Enriching the academic wordlists and secondary vocabulary list with lexicogrammar: Toward a pattern grammar of academic vocabulary. System, 87, 102158.
  • Green, C., & Lambert, J. (2019). Position vectors, homologous chromosomes and gamma rays: Promoting disciplinary literacy through secondary phrase lists. English for Specific Purposes, 53, 1–12.
  • Groves, F. H. (1995). Science vocabulary load of selected secondary science textbooks. School Science and Mathematics, 95(5), 231–235.
  • Gussarski, E., & Gorodetsky, M. (1988). On the chemical equilibrium concept: Constrained word associations and conceptions. Journal of Research in Science Teaching, 25(5), 319–333.
  • Gyasi, W. K. (2013). The role of readability in science education in Ghana: A readability index analysis of Ghana association of science teachers textbooks for senior high school. IOSR Journal of Research & Method in Education, 2(1), 9–19.
  • Gyllenpalm, J., Wickman, P.-O., & Holmgren, S.-O. (2010). Teachers’ language on scientific inquiry: Methods of teaching or methods of inquiry? International Journal of Science Education, 32(9), 1151–1172.
  • Ha, A. Y. H., & Hyland, K. (2017). What is technicality? A technicality analysis model for EAP vocabulary. Journal of English for Academic Purposes, 28, 35–49.
  • Haglund, J., Andersson, S., & Elmgren, M. (2015). Chemical engineering students’ ideas on entropy. Chemistry Education Research and Practice, 16, 537–551.
  • Haglund, J., Andersson, S., & Elmgren, M. (2016). Language aspects of engineering students’ view of entropy. Chemistry Education Research and Practice, 17, 489–508.
  • Haglund, J., Jeppson, F., & Strömdahl, H. (2010). Different senses of entropy-implications for education. Entropy, 12, 490–515.
  • Halliday, M. A. K., & Martin, J. R. (1993). Writing Science. London: The Falmer Press.
  • Harmon, J. M., & Hedrick, W. B. (2005). Research on vocabulary instruction in the content areas: Implications for struggling readers. Reading & Writing Quarterly, 21, 261–280.
  • Harrison, A. G., & Treagust, D. F. (2002). The particulate nature of matter: Challenges in understanding the submicroscopic world. In J. K. Gilbert, O. De Jong, R. Justi, D. F. Treagust, & J. H. Van Driel (Eds.), Chemical education: Towards research‐based practice (pp. 189–212). Dordrecht: Kluwer Academic Press.
  • Harrison, A. G., & Treagust, D. F. (1996). Secondary students’ mental models of atoms and molecules: Implications for teaching chemistry. Science Education, 80(5), 509–534.
  • Haug, B. S., & Ødegaard, M. (2014). From words to concepts: Focusing on word knowledge when teaching for conceptual understanding within an inquiry-based science setting. Research in Science Education, 44(5), 777–800.
  • Henderson, J., & Wellington, J. (1998). Lowering the language barrier in learning and teaching science. School Science Review, 79(288), 35–46.
  • Henderson, J. B., MacPherson, A., Osborne, J., & Wild, A. (2015). Beyond construction: Five arguments for the role and value of critique in learning science. International Journal of Science Education, 37(10), 1668–1697.
  • Hepler-Smith, E. (2015). “Just as the structural formula does”: Names, diagrams, and the structure of organic chemistry at the 1892 Geneva nomenclature congress. Ambix, 62(1), 1–28.
  • Herron, J. D. (1977). Are chemical terms well defined? Journal of Chemical Education, 54(12), 758.
  • Herron, J. D. (1978). Response to ‘Are chemical terms well defined’? Journal of Chemical Education, 55(6), 393.
  • Herron, J. D. (1979). Hey, watch your language! Journal of Chemical Education, 56(5), 330–331.
  • Herron, J. D. (1996). The chemistry classroom. Formulas for successful teaching. Washington: American Chemical Society.
  • Hesse, J. J., & Anderson, C. W. (1992). Students’ Conceptions of chemical change. Journal of Research in Science Teaching, 29(3), 277–299.
  • Hodson, D. (2009). Teaching and learning about science: Language, theories, methods, history, traditions and values. Rotherham: Sense.
  • Högström, P., Ottander, C., & Benckert, S. (2010). Lab work and learning in secondary school chemistry: The importance of teacher and student interaction. Research in Science Education, 40(4), 505–523.
  • Hyland, K., & Tse, P. (2007). Is there an ‘academic vocabulary’? TESOL Quarterly, 41(2), 235–254.
  • Hyland, K., & Tse, P. (2009). Approaches to English as a Foreign language reading comprehension: research and pedagogy. International Journal of English Studies, 9(2), 111–129.
  • Ibáñez, M., & Ramos, M. C. (2004). Physics textbooks presentation of the energy-conservation principle in hydrodynamics. Journal of Science Education and Technology, 13(2), 267–276.
  • Ingham, A. M., & Gilbert, J. K. (1991). The use of analogue models by students of chemistry at higher education level. International Journal of Science Education, 13, 193–202.
  • Itza-Ortiz, S. F., Rebello, N. S., Zollman, D., & Rodríguez-Achach, M. (2003). The vocabulary of introductory physics and its implications for learning physics. The Physics Teacher, 41(1), 41–46.
  • Jacobs, G. (1989). Word usage misconceptions among first-year university physics students. International Journal of Science Education, 11(4), 395–399.
  • Jasien, P. G. (2010). You said ‘neutral’, but what do you mean? Journal of Chemical Education, 88(1), 33–34.
  • Jasien, P. G. (2011). What do you mean that “strong” doesn’t mean “powerful”? Journal of Chemical Education, 88(10), 1247–1249.
  • Jasien, P. G. (2013). Roles of terminology, experience, and energy concepts in student conceptions of freezing and boiling. Journal of Chemical Education, 90, 1609–1615.
  • Jensen, W. B. (1996). Electronegativity from Avogadro to Pauling: Part 1: Origins of the electronegativity concept. Journal of Chemical Education, 73(1), 11–20.
  • Jensen, W. B. (2003). Electronegativity from Avogadro to Pauling: II. Late nineteenth- and early twentieth-century developments. Journal of Chemical Education, 80(3), 279–287.
  • Jensen, W. B. (2007). The origin of the oxidation-state concept. Journal of Chemical Education, 84(9), 1418–1419.
  • Jeppsson, F., Haglung, J., & Strömdalh, H. (2011). Exploiting language in teaching entropy. Journal of Baltic Science Education, 10(1), 27–35.
  • Johnson, B. E., & Zabrucky, K. M. (2011). Improving middle and high school students’ comprehension of science texts. International Electronic Journal of Elementary Education, 4(1), 19–31.
  • Johnstone, A. H. (1991). Why is science difficult to learn? Things are seldom what they seem. Journal of Computer Assisted Learning, 7, 75–83.
  • Johnstone, A. H., & Cassels, J. R. T. (1978). What’s in a word! New Scientist, 18, 432–434.
  • Johnstone, A. H., & Kellet, N. C. (1980). Learning difficulties in school science –Towards a working hypothesis. European Journal of Science Education, 2(2), 175–181.
  • Johnstone, A. H., & Selepeng, D. (2001). A language problem revisited. Chemistry Education Research and Practice, 2(1), 19–29.
  • Jones, H. L., Green, J. R., Prendergast, J., & Scott, J. (2019). Biology-specific vocabulary: Students’ understanding and learners’ expectations of student understanding. Journal of Biological Education, 53(4), 422–430.
  • Jung, K. G. (2019). Learning to scaffold science academic language: Lessons from an instructional coaching partnership. Research in Science Education, 49(4), 1013–1024.
  • Jung, K. G., & Brown, J. (2016). Examining the effectiveness of an academic language planning organizer as a tool for planning science academic instruction and supports. Journal of Science Teacher Education, 27, 847–872.
  • Justi, R., & Gilbert, J. (1999a). A cause of ahistorical science teaching: Use of hybrid models. Science Education, 83(2), 163–177.
  • Justi, R., & Gilbert, J. (1999b). History and philosophy of science through models: The case of chemical kinetics. Science and Education, 83(3), 287–307.
  • Justi, R., & Gilbert, J. (2000). History and philosophy of science through models: Some challenges in the case of the ‘atom’. International Journal of Science Education, 22(9), 993–1009.
  • Justi, R., & Gilbert, J. (2003). Teachers’ views on the nature of models. International Journal of Science Education, 25(11), 1369–1386.
  • Kahveci, A. (2010). Quantitative analysis of science and chemistry textbooks for indicators of reform: A complementary perspective. International Journal of Science Education, 32(11), 1495–1519.
  • Kallery, M., & Psillos, D. (2004). Anthropomorphism and animism in early years science: Why teachers use them, how they conceptualise them and what are their views on their use. Research in Science Education, 34, 291–311.
  • Kao, H.-L. (2007). A study of aboriginal and urban junior high school students’ alternative conceptions on the definition of respiration. International Journal of Science Education, 29(4), 349–371.
  • Kaya, E., & Erduran, S. (2013). Integrating epistemological perspectives on chemistry in chemical education: The case of concept duality, chemical language, and structural explanations. Science & Education, 22, 1741–1755.
  • Kearsey, J., & Turner, S. (1999). Evaluating textbooks: The role of genre analysis. Research in Science and Technological Education, 17, 35–43.
  • Kermen, I. (2018). Enseigner l’évolution des systemes chimiques au lycée. Savoirs et models, raisonnements d’élèves, pratiques enseignantes. Rennes: Presses Universitaires de Rennes.
  • Kidd, R., Ardini, R., & Anton, A. (1988). Evolution of the modern photon. American Journal of Physics, 57(1), 27–35.
  • Kirbulut, Z. D., & Beeth, M. E. (2013). Consistency of students’ ideas across evaporation, condensation, and boiling. Research in Science Education, 43(1), 209–232.
  • Knutton, S. (1983). Chemistry textbooks –Are they readable? Education in Chemistry, 20(3), 100–105.
  • Kotecha, P., Rutherford, M., & Starfield, S. (1990). Science, language or both? The development of a team teaching approach to English for science and technology. South African Journal of Education, 10(3), 212–221.
  • Kotsopoulos, D. (2007). Mathematics discourse: “It’s like hearing a foreign language”. Mathematics Teacher, 101(4), 301–305.
  • Koushatana, M., Demerouti, M., & Tsaparlis, G. (2005). Instructional misconceptions in acid-base equilibria: An analysis from a history and philosophy of science. Science & Education, 14(2), 173–193.
  • Laufer, B. (1988). The concept of ‘synforms’ (similar lexical forms) in vocabulary acquisition. Language and Education, 2(2), 113–132.
  • Lavoisier, A. (1789). Traité Elementaire de Chimie. Paris: Cuchet.
  • Lawrence, J. F., White, C., & Snow, C. E. (2010). The words students need. Educational Leadership, 68(2), 23–26.
  • Lazlo, P. (1999). Circulation of concepts. Foundations of Chemistry, 1, 225–238.
  • Lee, A. W. M., & Tse, C. L. (1994). Learning name reactions and name apparatuses through crossword puzzles. Journal of Chemical Education, 71(12), 1071–1072.
  • Lee, O., Fradd, S. H., & Sutman, F. X. (1995). Science knowledge and cognitive strategy use among culturally and linguistically diverse students. Journal of Research in Science Teaching, 32(8), 797–816.
  • Lee, O., Maaerten-Rivera, J., Buxton, C., Penfield, R., & Secada, W. G. (2009). Urban elementary teachers’ perspectives on teaching science to English language learners. Journal of Science Teacher Education, 20(3), 263–286.
  • Lemke, J. L. (1990). Talking science: Language, learning, and values. Norwood: Ablex Publishing Corporation.
  • Levy-Nahum, T., Hofstein, A., Mamlok-Naaman, R., & Bar-Dov, T. (2004). Can final examinations amplify students’ misconceptions in chemistry? Chemistry Education Research and Practice, 5(3), 301–325.
  • Lin, C., & Hu, R. (2003). Student’s understanding of energy flow and matter cycling in the context of the food chain, photosynthesis and respiration. International Journal of Science Education, 25(12), 1529–1544.
  • Llinares, A., Morton, T., & Whittaker, R. (2012). The roles of language in CLIL. Cambridge: Cambridge University Press.
  • Logan, S. R., & Logan, W. P. (1993). Scientifically speaking. Education in Chemistry, 30(2), 50–51.
  • López-Gay, R., Martínez-Sáez, J., & Martínez-Torregrosa, J. (2015). Obstacles to mathematization in physics: The case of the differential. Science & Education, 24, 591–613.
  • Love, K. (2009). Literacy pedagogical content knowledge in secondary teacher education: Reflecting on oral language and learning across the disciplines. Language and Education, 23(6), 541–560.
  • Loyson, P. (2009). Influences from Latin on chemical terminology. Journal of Chemical Education, 86(10), 1195–1199.
  • Loyson, P. (2010). Influences Ancient Greek on chemical terminology. Journal of Chemical Education, 87(12), 1303–1307.
  • Luder, W. F. (1948). Contemporary acid-base theory. Journal of Chemical Education, 25(10), 555–558.
  • Luzón, M. J. (1999). Procedural vocabulary: Lexical signalling of conceptual relations in discourse. Applied Linguistics, 20(1), 1–21.
  • Lynch, P. P., Benjamin, P., Chapman, T., Holmes, R., McCammon, R., Smith, A., & Symmons, R. (1979). Scientific language and the high school pupil. Journal of Research in Science Teaching, 16, 351–357.
  • Mammino, L. (2010). The essential role of language mastering in science and technology education. International Journal of Education and Information Technologies, 3(4), 139–148.
  • Markic, S., Broggy, J., & Childs, P. (2013). How to deal with linguistic issues in chemistry classes. In I. Eilks & A. Hofstein (Eds.), Teaching chemistry – A studybook (pp. 127–152). Rotterdam: Sense.
  • Markic, S. (2015). Chemistry teachers’ attitudes and needs when dealing with linguistic heterogeneity in the classroom. In M. Kahveci & M. Ogil (Eds.), Affective dimensions in chemistry education (pp. 279–296). Heidelberg: Springer.
  • Marshall, S., & Gilmour, M. (1990). Problematical words and concepts in physics education: A study of Papua New Guinean students’ comprehension of non-technical words used in science. Physics Education, 26, 330–337.
  • Marshall, S., Gilmour, M., & Lewis, D. (1991). Words that matter in science and technology. Research in Science & Technological Education, 9(1), 5–16.
  • Maskill, R. (1988). Logical language, natural strategies and the teaching of science. International Journal of Science Education, 10(5), 485–495.
  • Masrai, A., & Milton, J. (2017). Recognition vocabulary knowledge and intelligence as predictors of academic achievement in EFL context. TESOL International Journal, 12(1), 128–142.
  • McClure, G., & Cahnmann-Taylor, M. (2010). Pushing back against push-in: ESOL teacher resistance & the complexities of coteaching. TESOL Journal, 1(1), 101–129.
  • Méheut, M., Duprez, C., & Kermen, I. (2004). Approches historique et didactique de la réversibilité. Didaskalia, 25, 31–61.
  • Mercer, N., Dawes, L., & Wegerif, R. (2004). Reasoning as a scientist: Ways of helping children to use language to learn science. British Educational Research Journal, 30(3), 359–377.
  • Merzyn, G. (1987). The language of school science. International Journal of Science Education, 9(4), 483–489.
  • Meskill, C., & Oliveira, A. W. (2019). Meeting the challenges of English learners by pairing science and language educators. Research in Science Education, 49(4), 1025–1040.
  • Millar, R. (2014a). Teaching about energy: From everyday to scientific understandings. School Science Review, 96(354), 45–50.
  • Millar, R. (2014b). Towards a research-informed teaching sequence for energy. In R. F. Chen, A. Eisenkraft, D. Fortus, J. Krajcik, K. Neumann, J. C. Nordine, & A. Scheff (Eds.), Teaching and learning of energy in K-12 education (pp. 187–206). New York: Springer.
  • Miller, G. A. (1999). On knowing a word. Annual Review of Psychology, 50, 1–19.
  • Moje, E. B. (1995). Talking about science: An interpretation of the effects of teacher talk in a high school science classroom. Journal of Research in Science Teaching, 32(4), 349–371.
  • Moje, E. B. (2007). Developing socially just subject-matter instruction: A review of the literature on disciplinary literacy teaching. Review of Research in Education, 31, 1–44.
  • Moon, A., Gere, A. G., & Shultz, G. V. (2018). Writing in the STEM classroom: Faculty conceptions of writing and its role in the undergraduate classroom. Science Education, 102, 1007–1028.
  • Moon, A., Stanford, C., Cole, R., & Towns, M. (2017). Decentering: A characteristic of effective student-student discourse in inquiry-oriented physical chemistry classrooms. Journal of Chemical Education, 94(7), 829–836.
  • Mortimer, E., & Scott, P. (2000). Analyzing discourse in the science classroom. In R. Millar, J. Leach, & J. Osborne (Eds.), Improving science education: The contribution of research (pp. 126–142). Buckingham: Open University Press.
  • Mortimer, E., & Scott, P. (2003). Meaning making in secondary science classrooms. Philadelphia: Open University Press.
  • Munby, A. H. (1976). Some implications of language in science education. Science Education, 60(1), 115–124.
  • Myers, R. (2012). What are elements and compounds? Journal of Chemical Education, 89(7), 832–833.
  • Nagel, M. L., & Lindsey, B. A. (2015). Student use of energy concepts from physics in chemistry courses. Chemistry Education Research and Practice, 16(1), 67–81.
  • Nagy, W., & Townsend, D. (2012). Words as tools: Learning academic vocabulary as language acquisition. Reading Research Quarterly, 47, 91–108.
  • Nakhleh, M. B., & Krajcik, J. S. (1993). A protocol analysis of the influence of technology on students’ actions, verbal commentary, and thought processes during the performance of acid-base titrations. Journal of Research in Science Teaching, 30, 1149–1168.
  • Nakiboglu, C. (2003). Instructional misconceptions of Turkish prospective chemistry teachers about atomic orbitals and hybridization. Chemistry Education Research and Practice, 4, 171–188.
  • Nation, I. S. P. (2001). Learning vocabulary in another language. Cambridge: Cambridge University Press.
  • Nation, I. S. P. (2008). Teaching vocabulary. Strategies and techniques. Boston, MA: Heinle.
  • Nechamkin, H. (1958). Some interesting etymological derivations of chemical terminology. Science Education, 42, 463–474.
  • Nelson, P. G. (2003). Basic chemical concepts. Chemistry Education Research and Practice, 4(1), 19–24.
  • Newton, P. E., Driver, R., & Osborne, J. (1999). The place of argumentation in the pedagogy of school science. International Journal of Science Education, 21(5), 553–576.
  • Norris, S., & Phillips, L. (2003). How literacy in its fundamental sense is central to scientific literacy. Science Education, 87(2), 224–240.
  • Nyachwaya, J. M. (2016). General chemistry students’ conceptual understanding and language fluency: Acid-base neutralization and conductometry. Chemistry Education Research and Practice, 17, 509–522.
  • O’Rafferty, M. E. (1989). Pupil understanding of non-technical vocabulary in science textbooks. Irish Educational Studies, 8(2), 191–205.
  • O’Toole, M. (1996). Science, schools, children and books: Exploring the classroom interface between science and language. Studies in Science Education, 28(1), 113–144.
  • Olson, D. R., & Astington, J. W. (1990). Talking about text: How literacy contributes to thought. Journal of Pragmatics, 14, 705–721.
  • Osborne, J. (2002). Science without literacy: A ship without a sail? Cambridge Journal of Education, 32(2), 203–218.
  • Osborne, M. (1988). Access courses in mathematics, science and technology: Selected case studies. Journal of Access Studies, 3(2), 48–63.
  • Österlund, L. L., & Ekborg, M. (2009). Students’ understanding of redox reactions in three situations. Nordina, 5(2), 115–127.
  • Österlund, L. L., Berg, A., & Ekborg, M. (2010). Redox models in chemistry textbooks for upper secondary school: Friend or foe? Chemistry Education Research and Practice, 11(1), 182–192.
  • Oversby, J. (2000). Is it a weak acid or a weekly acidic solution? School Science Review, 81(297), 89–91.
  • Oyoo, S. O. (2009). An exploratory study of Kenyan physics’ teachers’ approaches to and perspectives on use of language during teaching. Electronic Journal of Literacy through Science, 8(2), 1–35.
  • Oyoo, S. O. (2012). Language in science classrooms: An analysis of physics teachers’ use of and beliefs about language. Research in Science Education, 42, 849–873.
  • Oyoo, S. O. (2017). Learner outcomes in South Africa: Role of the nature of learner difficulties with the language for learning and teaching science. Research in Science Education, 47(4), 783–804.
  • Özalp, D., & Kahveci, A. (2015). Diagnostic assessment of student misconceptions about the particulate nature of matter from ontological perspective. Chemistry Education Research and Practice, 16, 619–639.
  • Padilla, K., & Furió, C. (2008). The importance of history and philosophy of science in correcting distorted views of ‘amount of substance’ and ‘mole’ concepts in chemistry teaching. Science & Education, 17(4), 403–424.
  • Palmer, W. G. (1965). A history of the concept of valency to 1930. London: Cambridge University Press.
  • Park, J. Y. (2011). Revisiting the definitions and textbook descriptions of dissolution, diffusion and effusion. Journal of the Korean Association for Science Education, 31(6), 1009–1024.
  • Parkinson, J., Jackson, L., Kirkwood, T., & Padayachee, V. (2007). A scaffolded reading and writing course for foundation level science students. English for Specific Purposes, 26, 443–461.
  • Patterson, A., Román, D., Friend, M., Osborne, J., & Donovan, B. (2018). Reading for meaning: The foundational knowledge every teacher of science should have. International Journal of Science Education, 40(3), 291–307.
  • Pawan, J., & Ortloff, F. (2011).English-as-a-second-language, and content-area teachers. Teaching and Teacher Education, 27(2), 463–471.
  • Pearson, P. D., Moje, E., & Greenleaf, C. (2010). Literacy and science: Each in the service of the other. Science, 328, 459–463.
  • Pedrosa, M. A., & Dias, M. H. (2000). Chemistry textbook approaches to chemical equilibrium and student alternative conceptions. Chemistry Education Research and Practice, 1, 227–236.
  • Pekdag, B., & Azizoglu, N. (2013). Semantic mistakes and didactic difficulties in teaching the “amount of substance” concept: A useful model. Chemistry Education Research and Practice, 14, 117–129.
  • Penney, K., Norris, S. P., Phillips, L. M., & Clark, G. (2003). The anatomy of junior high school science textbooks: An analysis of textual characteristics and a comparison to media reports of science. Canadian Journal of Science, Mathematics and Technology Education, 3(4), 415–436.
  • Pickershill, S., & Lock, R. (1991). Student understanding of selected non-technical words in science. Research in Science & Technological Education, 9(1), 71–79.
  • Pinarbasi, T. (2007). Turkish undergraduate students’ misconceptions on acids and bases. Journal of Baltic Science Education, 6(1), 23–34.
  • Pisano, R., Abdelkader, A., Pellegrino, E. M., & Nagels, M. (2018). Thermodynamic foundations of physical chemistry: Reversible processes and thermal equilibrium into the history. Foundations of Chemistry. doi:10.1007/s10698-018-09324-1
  • Postman, N., & Weingartner, C. (1971). Teaching science as a subversive activity. London: Penguin.
  • Pritchard, H. O., & Skinner, H. A. (1955). The concept of electronegativity. Chemical Reviews, 55(4), 745–786.
  • Prophet, B., & Towse, P. (1999). Pupils’ understanding of some non-technical words in science. School Science Review, 81(295), 79–86.
  • Prophet, R. B., & Badede, N. B. (2009). Language and student performance in junior secondary science examinations: The case of second language learners in Botswana. International Journal of Science and Mathematics Education, 7, 235–251.
  • Pyburn, D. T., Pazicni, S., Benassi, V. A., & Tappin, E. (2013). Assessing the relation between language comprehension and performance in general chemistry. Chemistry Education Research and Practice, 14, 524–541.
  • Quílez, J. (2009). From chemical forces to chemical rates: A historical/philosophical foundation for the teaching of chemical equilibrium. Science & Education, 18(9), 1203–1251.
  • Quílez, J. (2012). First-year university chemistry textbooks’ misrepresentation of Gibbs energy. Journal of Chemical Education, 89(1), 87–93.
  • Quílez, J. (2019). A historical/epistemological account of the foundation of the key ideas supporting chemical equilibrium theory. Foundations of Chemistry, 21(2), 221–252.
  • Quinn, H., Lee, O., & Valdés, G. (2012). Language demands and opportunities in relation to next generation science standards for English language learners: What teachers need to know. Understanding language conference (pp. 32–43). Stanford, CA: Stanford University.
  • Ramos, K. A., & Foster, A. A. M. (2017). Bridging the gap: Building teachers’ capacity for developing language, literacy, and content learning. International Schools Journal, 37(1), 58–66.
  • Rappa, N. A., & Tang, K. S. (2018). Integrating disciplinary-specific genre structure in discourse strategies to support disciplinary literacy. Linguistics and Education, 43, 1–12.
  • Rees, S. W., Kind, V., & Newton, D. (2018). Can language focussed activities improve understanding of chemical language in non-traditional students? Chemistry Education Research and Practice, 19, 755–766.
  • Reid, N., & Yan, M. J. (2002). Open-ended problem-solving in school chemistry: A preliminary investigation. International Journal of Science Education, 24(12), 1313–1332.
  • Reynolds, J. A., Thais, C., Katkim, W., & Thomson, R. J. (2012). Writing-to-learn in undergraduate science education: A common-based, conceptually driven approach. CBE Life Sciences Education, 11(1), 17–25.
  • Ribeiro, M. G. T. C., Pereira, D. J. V. C., & Maskill, R. (1990). Reaction and spontaneity: The influence of meaning from everyday language on fourth year undergraduates’ interpretations of some simple chemical phenomena. International Journal of Science Education, 12(4), 391–401.
  • Rincke, K. (2011). It’s rather like learning a language: Development of talk and conceptual understanding in mechanics lessons. International Journal of Science Education, 33(2), 229–258.
  • Ringnes, V. (1995). Oxidation-reduction – Learning difficulties and choice of redox models. School Science Review, 77(279), 74–78.
  • Rodrigues, S., & Thompson, I. (2001). Cohesion in science lesson discourse: Clarity, relevance and sufficient information. International Journal of Science Education, 23(9), 929–940.
  • Romer, R. H. (2001). Heat is not a noun. American Journal of Physics, 69(2), 107–109.
  • Roon, P. H., van Sprang, H. F., & Verdonk, A. H. (1994). ‘Work’ and ‘Heat’: On a road towards thermodynamics. International Journal of Science Education, 16(2), 131–144.
  • Roth, W. M. (2005). Talking science. Language and learning in science classrooms. Lanham: Rowman & Littlefield.
  • Rusek, M., & Vojíř, K. (2019). Analysis of text difficulty in lower-secondary chemistry textbooks. Chemistry Education Research and Practice, 20(1), 85–99.
  • Ruthenberg, K., & Martínez-González, J. C. (2017). Electronegativity and its multiple faces: Persistence and measurement. Foundations of Chemistry, 19(1), 61–75.
  • Ryan, J. N. (1985). The language gap: Common words with technical meanings. Journal of Chemical Education, 62(12), 1098–1099.
  • Sager, J. C., Dungworth, D., & McDonald, P. F. (1980). English special languages: Principles and practice in science and technology. Wiesbaden: Brandstetter.
  • Sanabria-Ríos, D., & Bretz, S. L. (2010). Investigating the relationship between faculty cognitive expectations about learning chemistry and the construction of exam questions. Chemistry Education Research and Practice, 11, 212–217.
  • Sanger, M. J., & Greenbowe, T. J. (1999). An analysis of college chemistry textbooks as sources of misconceptions and errors in electrochemistry. Journal of Chemical Education, 76, 853–860.
  • Sanmartí, N., Izquierdo, M., & Watson, R. (1995). The substantialisation of properties in pupils’ thinking and in the history of science. Science & Education, 4, 349–369.
  • Sarma, N. S. (2004). Etymology as an aid to understand chemistry concepts. Journal of Chemical Education, 81(10), 1437–1439.
  • Savory, T. H. (1953). The language of science: Its growth, character and usage. London: Andre Deutsch.
  • Schizas, D., Katrana, E., & Stamou, G. (2013). Introducing network analysis into science education: Methodological research examining secondary school students’ understanding of ‘decomposition’. International Journal of Environmental & Science Education, 8(1), 175–198.
  • Schleppegrell, M. J. (2004). The language of schooling. A functional linguistics perspective. New Jersey: LEA.
  • Schleppegrell, M. J. (2007). The linguistic challenges of mathematics teaching and learning: A research review. Reading & Writing Quarterly: Overcoming Learning Difficulties, 23(2), 139–159.
  • Schmid, S., Youl, D. J., George, A. V., & Read, J. R. (2012). Effectiveness of a short, intense bridging course for scaffolding students commencing university-level study of chemistry. International Journal of Science Education, 34(8), 1211–1234.
  • Schmidt, H.-J. (1991). A label as a hidden persuader: Chemists’ neutralization concept. International Journal of Science Education, 13, 459–471.
  • Schmidt, H.-J. (1992). Conceptual difficulties with isomerism. Journal of Research in Science Teaching, 29(9), 995–1003.
  • Schmidt, H.-J. (1997). Students’ misconceptions – Looking for a pattern. Science Education, 81(2), 123–135.
  • Schmidt, H.-J. (1998). Does the periodic table refer to chemical elements? School Science Review, 80(290), 71–74.
  • Schmidt, H.-J., & Volke, D. (2003). Shift of meaning and students’ alternative concepts. International Journal of Science Education, 25(11), 1409–1424.
  • Schoerning, E. (2014). The effect of plain-English vocabulary on student achievement and classroom culture in college science instruction. International Journal of Science and Mathematics Education, 12(2), 307–327.
  • Schultz, E. (1997). Ionization or dissociation? Journal of Chemical Education, 74(7), 868–869.
  • Seah, L. H. (2016). Elementary teachers’ perception of language issues in science classrooms. International Journal of Science and Mathematics Education, 14(6), 1059–1078.
  • Seah, L. H., Clarke, D. J., & Eugene, C. (2014). Understanding the language demands of science students from an integrated science and language perspective. International Journal of Science Education, 36(6), 952–973.
  • Seah, L. H., Clarke, D. J., & Hart, C. E. (2011). Understanding students’ language use about expansion through analyzing their lexicogrammatical resources. Science Education, 95(5), 852–876.
  • Seah, L. H., Clarke, D. J., & Hart, C. E. (2015). Understanding middle school students’ difficulties in explaining density differences from a language perspective. International Journal of Science Education, 37(14), 2386–2409.
  • Seah, L. H., & Silver, R. E. (2018). Attending to science language demands in multilingual classrooms: A case study. International Journal of Science Education, 1–19. doi:10.1080/09500693.2018.1504177
  • Seçken, N. (2010). Identifying student’s misconceptions about SALT. Procedia - Social and Behavioral Sciences, 2(2), 234–245.
  • Seethaler, S., Czworkowski, J., & Wynn, L. (2018). Analysing general chemistry texts’ treatment of rates of change concepts in reaction kinetics reveals missing conceptual links. Journal of Chemical Education, 95(1), 28–36.
  • Sheppard, K. (2006). High school students’ understanding of titrations and related acid-base phenomena. Chemistry Education Research and Practice, 7(1), 32–45.
  • Shiland, T. W. (1997). Quantum mechanics and conceptual change in high school chemistry textbooks. Journal of Research in Science Teaching, 34(5), 535–545.
  • Shiram, G. (2007). Learning difficulties in chemistry: An overview. Journal of Turkish of Science Education, 4(2), 2–20.
  • Silberstein, T. P. (2011). Oxidation and reduction: Too many definitions? Journal of Chemical Education, 88(3), 279–281.
  • Silva, J. R. (2017). Diversos modos de pensar o conceito de substância química na história da ciência e sua visão relacional. Ciência & Educação, 23(3), 707–722.
  • Sima, J. (2013). Redox reactions: Inconsistencies in their description. Foundations of Chemistry, 15, 57–64.
  • Slater, T., & Mohan, B. (2010). Cooperation between science teachers and ESL teachers: A register perspective. Theory into Practice, 49(2), 91–98.
  • Slisko, J., & Dykstra, D. I. (1997). The role of scientific terminology in research and teaching: Is something important missing? Journal of Research in Science Teaching, 34(6), 655–660.
  • Smith-Walters, C., Mangione, K. A., & Smith-Bass, A. (2016). Science and language special issue: Challenges in preparing preservice teachers for teaching science as a second language. Electronic Journal of Science Education, 20(3), 59–71.
  • Snow, C. (2002). Reading for understanding. Toward an R&D program in reading comprehension. Santa Monica, CA: RAND.
  • Snow, C. (2010). Academic language and the challenge of reading for learning about science. Science, 328, 450–452.
  • Song, Y., & Carheden, S. (2014). Dual meaning vocabulary (DMV) words in learning chemistry. Chemistry Education Research and Practice, 15, 128–141.
  • Sözbilir, M. (2003). What students understand from entropy. Journal of Baltic Science Education, 2(1), 21–27.
  • Spencer, B. H., & Guillaume, A. M. (2006). Integrating curriculum through the learning cycle: Content-based reading and vocabulary instruction. The Reading Teacher, 60(3), 206–219.
  • Staver, J. R., & Lumpe, A. T. (1993). A content analysis of the presentation of the mole concept in chemistry textbooks. Journal of Research in Science Teaching, 30(4), 321–337.
  • Stoddart, T., Pinal, A., Latzke, M., & Canaday. (2002). Integrating inquiry science and language development for English language learners. Journal or Research in Science Teaching, 39, 664–687.
  • Stojanovska, M., Petruševski, V., Köller, H.-G., & Karlsenet, S. (2015). Students’ alternative conceptions and ways to overcome them. In I. Maciejowska & B. Byers (Eds.), A guidebook of good practice for the pre-service training of chemistry teachers (pp. 175–202). Krakow: Jagiellonian University.
  • Stojanovska, M., Petruševski, V. M., & Šoptrajanov, B. (2012). The concept of sublimation – Iodine as an example. Educación Química, 23(1), 171–173.
  • Stubbs, M. (1976). Language, schools and classrooms. London: Methuen.
  • Sumfleth, E. (1988). Knowledge of terms and problem-solving in chemistry. International Journal of Science Education, 10(1), 45–60.
  • Sutton, C. (1996). Beliefs about science and beliefs about language. International Journal of Science Education, 18(1), 1–18.
  • Sutton, C. R. (1980). Science, language and meaning. School Science Review, 218, 47–56.
  • Sutton, C. R. (1992). Words, science and learning. Buckingham: Open University Press.
  • Swartz, Y., Ben-Zvi, R., & Hofstein, A. (2006). The use of scientific literacy taxonomy for assessing the development of chemical literacy among high-school students. Chemistry Education Research and Practice, 7, 203–225.
  • Swendsen, R. H. (2011). How physicists disagree on the meaning of entropy. American Journal of Physics, 79(4), 342–348.
  • Syamal, A. (1985). Some improper terms in coordination chemistry. Journal of Chemical Education, 62(2), 143.
  • Taber, K. (2001). Building the structural concepts of chemistry: Some considerations from educational research. Chemistry Education Research and Practice in Europe, 2(2), 123–158.
  • Taber, K. (2002). Conceptualizing quanta: Illuminating the ground state of student understanding of atomic orbitals. Chemistry Education Research and Practice in Europe, 3(2), 145–158.
  • Taber, K. (2009a). College students’ conceptions of chemical stability: The widespread adoption of a heuristic rule out of context and beyond its range of application. International Journal of Science Education, 31(10), 1333–1358.
  • Taber, K. (2009b). Learning at the symbolic level. In J. K. Gilbert & D. Treagust (Eds.), Multiple representations in chemical education (pp. 75–105). Dordrecht: Springer.
  • Taber, K. S., & Coll, R. K. (2002). Bonding. In J. K. Gilbert, O. De Jong, R. Justi, D. F. Treagust, & J. H. Van Driel (Eds.), Chemical education: Towards research-based practice (pp. 213–234). Dordrecht: Kluwer.
  • Taber, K. S., & Watts, M. (1996). The secret life of the chemical bond: Students’ anthropomorphic and animistic references to bonding. International Journal of Science Education, 18(5), 557–568.
  • Taber, K. S., & Watts, M. (2000). Learners’ explanations for chemical phenomena. Chemistry Education Research and Practice, 1(3), 329–353.
  • Talanquer, V. (2007). Explanations and teleology in chemistry education. International Journal of Science Education, 29(7), 853–870.
  • Talanquer, V. (2010). Exploring dominant types of explanations built by general chemistry students. International Journal of Science Education, 32(18), 2393–2412.
  • Talanquer, V. (2013). When atoms want. Journal of Chemical Education, 90(11), 1419–1424.
  • Tan, M. (2011). Mathematics and science teachers’ beliefs and practices regarding the teaching of language in content learning. Language Teaching Research, 15(3), 325–342.
  • Tang, K. S. (2016a). How is disciplinary literacy addressed in the science classrooms? A Singaporean case study. Australian Journal of Language and Literacy, 39(3), 220–232.
  • Tang, K. S. (2016b). Constructing scientific explanations through Premise-Reasoning-Outcome (PRO): An exploratory study to scaffold students in structuring written explanations. International Journal of Science Education, 38(9), 1415–1440.
  • Tao, P. K. (1994a). Comprehension of non-technical words in science: The case of students using a ‘foreign language’ as the medium of instruction. Research in Science Education, 24, 322–330.
  • Tao, P. K. (1994b). Words that matter in science: A study of Hong Kong students’ comprehension of non-technical words in science. Educational Research Journal, 9(1), 15–23.
  • Thomas, P. L., & Schwenz, R. W. (1998). College physical chemistry students’ conceptions of equilibrium and fundamental thermodynamics. Journal of Research in Science Teaching, 25(10), 1151–1160.
  • Thonney, T. (2016). Analyzing the vocabulary demands of introductory college textbooks. The American Biology Teacher, 78(5), 389–395.
  • Tolbert, S., Knox, C., & Salinas, I. (2019). Framing, adapting, and applying: Learning to contextualize science activity in multilingual science classrooms. Research in Science Education, 49(4), 1069–1085.
  • Townsend, D. (2009). Building academic vocabulary in after school settings: Games for growth with middle school English learners. Journal of Adolescent & Adult Literacy, 53, 242–251.
  • Townsend, D., Brock, C., & Morrison, J. D. (2018). Engaging in vocabulary learning in science: The promise of multimodal instruction. International Journal of Science Education, 40(3), 328–347.
  • Townsend, D., Filippini, A., Collins, P., & Biancarosa, G. (2012). Evidence for the importance of academic word knowledge for the academic achievement of diverse middle school students. The Elementary School Journal, 112(3), 497–518.
  • Tsaparlis, G., & Papaphotis, G. (2009). High-school students’ conceptual difficulties and attempts at conceptual change: The case of basic quantum chemical concepts. International Journal of Science Education, 31(7), 895–930.
  • Ucelli, P., & Galloway, P. (2016). Academic language across content areas: Lessons from an innovative assessment and from students’ reflection about language. Journal of Adolescent & Adult Literacy, 60(4), 395–404.
  • Udu, T. T., Gyuse, E. Y., Samba, R. M. O., & Iortim, S. (2016). Readability characteristics of Nigerian science textbooks as a factor of students’ science achievement. ICSHER Journal, 2(2), 1–13.
  • Valipoury, L., & Nassaji, H. (2013). A corpus-based study of academic vocabulary in chemistry research articles. Journal of English for Academic Purposes, 12, 248–263.
  • Van Driel, J. H., & Verloop, N. (1999). Teachers’ knowledge of models and modelling in science. International Journal of Science Education, 21(11), 1141–1153.
  • Vladušić, R., Bucat, R., & Ožić, M. (2016). Understanding of words and symbols by chemistry university students in Croatia. Chemistry Education Research and Practice, 17, 474–488.
  • Vygotsky, L. S. (1962). Thought and language. Cambridge: MIT.
  • Vygotsky, L. S. (1978). Mind in society: The development of higher psychological processes. Cambridge: Harvard University Press.
  • Webb, P. (2010). Science education and literacy: Imperatives for the developed and developing world. Science, 328, 448–450.
  • Webb, P., & Treagust, D. F. (2006). Using exploratory talk to enhance problem-solving and reasoning skills in grade-7 science classrooms. Research in Science Education, 36, 381–401.
  • Weininger, S. J. (1998). Contemplating the finger: Visuality and the semiotics of chemistry. Hyle, 4(1), 3–27.
  • Wellington, J. (1994). Secondary science: Contemporary issues and practical approaches. London: Routledge.
  • Wellington, J. (2001). School textbooks and reading in science: Looking back and looking forward. School Science Review, 82(300), 71–81.
  • Wellington, J., & Osborne, J. (2001). Language and literacy in science education. Buckingham: Open University Press.
  • West, M. (1953). A general service list for english words. Green & Co:London: Longman.
  • Westbrook, S. L., & Marek, E. A. (1991). A cross-age study of student understanding of the concept of diffusion. Journal of Research in Science Teaching, 28(8), 649–660.
  • Widarti, H. R., Permanasari, A., & Mulyani, S. (2017). Undergraduate students’ misconception on acid-base and argentometric titrations: A challenge to implement multiple representation learning model with cognitive dissonance strategy. International Journal of Education, 9(2), 105–112.
  • Williams, H. T. (1999). Semantics in teaching introductory physics. American Journal of Physics, 67(8), 670–680.
  • Wilson, J. M. (1999). Using words about thinking: Content analyses of chemistry teachers’ classroom talk. International Journal of Science Education, 21(10), 1067–1084.
  • Wiswesser, W. J. (1985). Historic development of chemical notations. Journal of Chemical Information and Computer Sciences, 25, 258–263.
  • Wolfson, A. J., Rowland, S. L., Lawrie, G. A., & Wright, A. H. (2014). Student conceptions about energy transformations: Progression from general chemistry to biochemistry. Chemistry Education Research and Practice, 15(2), 168–183.
  • Wong, C. L., & Yap, K. C. (2012). Can definitions contribute to alternative conceptions?: A meta-study approach. Journal of the Korean Association for Science Education, 32(8), 1295–1317.
  • Yager, R. E. (1983). The importance of terminology in teaching K-12 science. Journal of Research in Science Teaching, 20(6), 577–588.
  • Yager, R. E., Akcay, H., Choi, A., & Yager, S. O. (2009). Student success in recognizing definitions of eight terms found in fourth grade science textbooks. Electronic Journal of Science Education, 13(2), 83–99.
  • Yong, B. C. S. (2010). Can students read secondary science textbooks comfortably? Brunei International Journal of Science and Mathematics Education, 2(1), 59–67.
  • Yore, L. D., Bisanz, G. L., & Hand, B. M. (2003). Examining the literacy component of science literacy: 25 years of language arts and science research. International Journal of Science Education, 25(6), 689–725.
  • Yore, L. D., Hand, B. M., Goldman, S. R., Hildebrand, G. M., Osborne, J. F., Treagust, D. F., & Wallace, C. S. (2004). New directions in language and science education research. Reading Research Quarterly, 39(3), 347–352.
  • Yore, L. D., & Treagust, D. F. (2006). Current realities and future possibilities: Language and science literacy – Empowering research and informing instruction. International Journal of Science Education, 28(2–3), 291–314.
  • Yun, E., & Park, Y. (2018). Extraction of scientific semantic networks from science textbooks and comparison with science teachers’ spoken language by text network analysis. International Journal of Science Education, 40(17), 2118–2136.
  • Yuriev, E., Capuano, B., & Short, J. L. (2016). Crossword puzzles for chemistry education: Learning goals beyond vocabulary. Chemistry Education Research and Practice, 17, 532–554.
  • Zhang, F., Lidbury, B., Richardson, A. M., Yates, B. F., Gardiner, M. G., Bridgeman, A., … Mate, K. E. (2012). Sustainable language support practices in science education. Technologies and solutions. Hersley: Information Science Reference.
  • Zhang, F., Lidbury, B., Schulte, J., Yates, B., Bridgeman, A., & Rodger, J. (2010). Integrating language learning practices in first year science disciplines. The International Journal of Learning, 17(4), 481–502.
  • Zwiers, J. (2014). Building academic language: Meeting common core standards across disciplines, grades 5-12. Newark: Jossey-Bass.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.