Skip to main content

The Learning and Teaching of Linear Algebra Through the Lenses of Intellectual Need and Epistemological Justification and Their Constituents

  • Chapter
  • First Online:

Part of the book series: ICME-13 Monographs ((ICME13Mo))

Abstract

Intellectual need and epistemological justification are two central constructs in a conceptual framework called DNR-based instruction in mathematics. This is a theoretical paper aiming at analyzing the implications of these constructs and their constituent elements to the learning and teaching of linear algebra. At the center of these analyses are classifications of intellectual need and epistemological justification in mathematical practice along with their implications to linear algebra curriculum development and instruction. Two systems of classifications for intellectual need are discussed. The first system consists of two subcategories, global need and local need; and the second system consists of five categories of needs: need for certainty, need for causality, need for computation, need for communication, and formalization, and need for structure. Epistemological justification is classified into three categories: sentential epistemological justification (SEJ), apodictic epistemological justification (ASJ), and meta epistemological justification (MEJ).

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

Notes

  1. 1.

    APOS theory (Arnon et al., 2014; Dubinsky, 1991) will be used to provide conceptual bases for some of these observations. Given how widely this theory has been studied during the last three decades, there is no need to allocate more than a brief illustration to the four levels of conceptualizations, action, process, object, and schema offered by the theory and used in this paper. Briefly, consider the phrase “the coordinates of a vector of \(x\) with respect to a basis-matrix \(A\) in \(R^{n}\),” denoted by \(\left[ x \right]_{A}\). At the level of action conception, the learner might be able to deal with \(\left[ x \right]_{A}\) only in the context of a specific vector and a specific suitable basis-matrix, by following step-by-step instruction to compute the respective coordinate vector. At the level of process conception one is capable of imagining taking any vector \(x\) in \(R^{n}\), representing it as a linear combination of the columns of \(A\), and forming a column vector whose entries are the coefficient of, and are sequenced in the order they appear in, the combination. With this conceptualization, the learner is able to carry out this process in thought and with no restriction on the vector \(x\) considered. At the level of object conception, one is aware of the process of relating the two coordinate vectors as a totality, for example, in finding the relation between two coordinate vectors of \(x\), one with respect to a basis-matrix \(A_{1}\), \(\left[ x \right]_{{A_{1} }}\), and one with respect to a basis-matrix \(A_{2}\), \(\left[ x \right]_{{A_{2} }}\), whereby being able to express the relation in terms of a transition matrix \(S = A_{2}^{ - 1} A_{1}\) between the two vectors. Among the ways of thinking that are essential to cope with linear algebra, in particular, and mathematics, in general, are the abilities to construct concepts at the levels of process conception and object conception, as it is demonstrate throughout the paper. (See also Trigueros, this volume.)

  2. 2.

    A matrix whose columns form a basis for a subspace.

  3. 3.

    For the philosophical foundations of this premise, see Harel (2008c).

  4. 4.

    Questions concerning convergence are not discussed, though on rare occasions were raised by students.

  5. 5.

    Of course other contexts can be used as SEJ for the SVD Theorem. The problem of transmitting a digitized image is typically used in textbooks as an application of SVD; we, on the hand, used it as an intellectual motivation (see the distinction between “application” and “intellectual need” in Sect. 3).

References

  • Arnon, I., Cottrill, J., Dubinsky, E., Oktac, A., Roa, S. Trigueros, M., & Weller. K. (2014). APOS Theory—A Framework For Research And Curriculum Development In Mathematics Education, Springer New York, Heidelberg, Dordrecht, London, 2013.

    Google Scholar 

  • Brousseau, G. (1997). Theory of didactical situations in mathematics, Dordrecht: Kluwer.

    Google Scholar 

  • Confrey, J. (1991). Steering a course between Vygotsky and Piaget. Educational Researcher, 20(8), 28–32.

    Google Scholar 

  • Cooper, R. (1991). The role of mathematical transformations and practice in mathematical development. In L. Steffe (Ed.). Epistemological Foundations of Mathematical Experience. New York. Springer-Verlag.

    Google Scholar 

  • Dubinsky, E. (1991). Reflective Abstraction in Advanced Mathematical Thinking, in Advanced Mathematical Thinking (D. Tall, ed.), Kluwer, 95–126.

    Google Scholar 

  • Greeno, G. (1992). Mathematical and scientific thinking in classroom and other situations. In D. Halpern (Ed.), Enhancing Thinking Sills in Sciences and Mathematics (pp. 39–61). Hillsdale: Lawrence Erlbaum Associates.

    Google Scholar 

  • Harel, G. (2008a). DNR Perspective on Mathematics Curriculum and Instruction: Focus on Proving, Part I, ZDM—The International Journal on Mathematics Education, 40, 487–500.

    Google Scholar 

  • Harel, G. (2008b). DNR Perspective on Mathematics Curriculum and Instruction, Part II, ZDM—The International Journal on Mathematics Education, 40, 893–907.

    Google Scholar 

  • Harel, G. (2008c). What is mathematics? A pedagogical answer to a philosophical question. In B. Gold & R. Simons (Eds.), Proof and other dilemmas: Mathematics and philosophy (pp. 265–290). Washington, DC: Mathematical Association of America.

    Google Scholar 

  • Harel, G. (2013a). Intellectual Need. In Vital Direction for Mathematics Education Research, Leatham, K. (Ed.), Springer.

    Google Scholar 

  • Harel, G. (2013b). DNR-based curricula: The case of complex numbers. Journal of Humanistic Mathematics, 3 (2), 2–61.

    Google Scholar 

  • Harel, G. (2014). Common Core State Standards for Geometry: An Alternative Approach. Notices of the AMS, 61 (1), 24–35.

    Google Scholar 

  • Harel, G. (1998). Two Dual Assertions: The First on Learning and the Second on Teaching (Or Vice Versa). The American Mathematical Monthly, 105, 497–507.

    Google Scholar 

  • Harel, G. (2000). Three principles of learning and teaching mathematics: Particular reference to linear algebra—Old and new observations. In Jean-Luc Dorier (Ed.), On the Teaching of Linear Algebra, Kluwer Academic Publishers, 177–190.

    Google Scholar 

  • Harel. G. (in press a). The learning and teaching of linear algebra: Observations and generalizations. Journal of Mathematical Behavior.

    Google Scholar 

  • Harel. G. (in press b). Field-based hypotheses on advancing standards for mathematical practice. Journal of Mathematical Behavior.

    Google Scholar 

  • Harel, G., & Soto, O. (in press). Structural reasoning. International Journal of Research in Undergraduate Mathematics Education.

    Google Scholar 

  • Harel, G., & Soto, O. (2016). Structural reasoning. International Journal of Research in Undergraduate Mathematics Education, 3(1), 225–242.

    Google Scholar 

  • Harel, G., & Sowder, L. (1998). Students’ proof schemes. In E. Dubinsky, A. Schoenfeld, & J. Kaput (Eds.), Research on Collegiate Mathematics Education (Vol. III, pp. 234–283). AMS.

    Google Scholar 

  • Harel, G., & Sowder, L (2007). Toward a comprehensive perspective on proof, In F. Lester (Ed.), Second Handbook of Research on Mathematics Teaching and Learning, National Council of Teachers of Mathematics.

    Google Scholar 

  • Lay, D., Lay, S., & McDonald, J. (2016). Linear Algebra and its applications. Pearson, Boston.

    Google Scholar 

  • Li, H. (2008). Invariant algebras and geometric reasoning, World Scientific Publishing Co. Pte. Ltd.

    Google Scholar 

  • Piaget, J. (1985). The equilibration of cognitive structures: the central problem of intellectual development: Chicago: University of Chicago Press.

    Google Scholar 

  • Steffe, L. P., Cobb, P., & von Glasersfeld, E. (1988). Young children’s construction of arithmetical meanings and strategies. New York, NY: Springer-Verlag.

    Google Scholar 

  • Steffe, L. P. & Thompson, P. W. (2000). Interaction or intersubjectivity? A reply to Lerman. Journal for Research in Mathematics Education, 31, 191–209.

    Google Scholar 

  • Trigueros, M. (this volume). Learning linear algebra using of models and conceptual activities.

    Google Scholar 

  • Tucker, A. (1993). The growing importance of linear algebra in undergraduate mathematics. The College Mathematics Journal, 24, pp. 3–9.

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Guershon Harel .

Editor information

Editors and Affiliations

Rights and permissions

Reprints and permissions

Copyright information

© 2018 Springer International Publishing AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Harel, G. (2018). The Learning and Teaching of Linear Algebra Through the Lenses of Intellectual Need and Epistemological Justification and Their Constituents. In: Stewart, S., Andrews-Larson, C., Berman, A., Zandieh, M. (eds) Challenges and Strategies in Teaching Linear Algebra. ICME-13 Monographs. Springer, Cham. https://doi.org/10.1007/978-3-319-66811-6_1

Download citation

  • DOI: https://doi.org/10.1007/978-3-319-66811-6_1

  • Published:

  • Publisher Name: Springer, Cham

  • Print ISBN: 978-3-319-66810-9

  • Online ISBN: 978-3-319-66811-6

  • eBook Packages: EducationEducation (R0)

Publish with us

Policies and ethics